Paper

Halos and related structures

Published 28 January 2013 © 2013 The Royal Swedish Academy of Sciences
, , Citation K Riisager 2013 Phys. Scr. 2013 014001 DOI 10.1088/0031-8949/2013/T152/014001

1402-4896/2013/T152/014001

Abstract

The halo structure originated from nuclear physics but is now encountered more widely. It appears in loosely bound, clustered systems where the spatial extension of the system is significantly larger than that of the binding potentials. A review is given on our current understanding of these structures, with an emphasis on how the structures evolve as more cluster components are added and on the experimental situation concerning halo states in light nuclei.

Export citation and abstract BibTeX RIS

1. Introduction

The word 'halo' is used in many areas of human activity as a glance in an encyclopedia will show. A common feature of most of these meanings is an extended peripheral distribution, most often dilute, around a central object. This also holds for the object of this paper: the quantum halos found up to now are mainly in nuclear physics, but are also known to exist in molecular physics. A more precise definition follows in the next section; for now, a spatial extension significantly larger than that of otherwise similar systems will do as the characteristic property. They first came into focus through the measurements of nuclear matter radii carried out by Tanihata et al [1, 2] and obtained their name 25 years ago in the paper by Hansen and Jonson [3], where the key ingredient of their structure was identified. Halos were a major contributor to the growing interest in radioactive beams that followed shortly after that, and aspects of halos and their study have been covered in many review papers since then, for example [414].

It is only natural that halos will appear in many of the contributions to these proceedings. The aim here is to give an overview of what characterizes the halo systems and how they are probed experimentally. Section 2 focuses on the halo structure, what distinguishes it from other systems as well as its intimate relation to the Efimov effect. Section 3 gives a quick overview of the many types of experiments that have been used to study nuclear halos and also attempts to extract generic traits. Section 4 summarizes which pronounced halos are established currently and points to some of the as yet unsolved questions.

2. The structure of halos

2.1. Basics

The understanding of halos that has emerged after the first two decades of studies has changed very little since the latest major review [13], and I shall rely on that for much of the general description and refer to it for a more detailed exposition.

The defining feature of a halo was from the beginning understood to be a large spatial extension caused by neutrons tunneling out from a nuclear core. This picture, based on the first established cases such as 11Li, still prevails but must be refined in order to use the concept consistently also in other physics disciplines [13, 15]. It is not sufficient that a system be large; the tunneling that arises from the quantum wave nature must also be a prominent feature of the system in order that we obtain structures that are truly universal. As an example, Rydberg atoms are therefore excluded since most of the wavefunctions here reside in a classically allowed region; a long-range attractive potential will, in general, not give halos. The system should be dividable into a core and one or more halo particles that can tunnel out, making cluster models or few-body models a natural first choice for describing halos. Tunneling will occur in all quantum systems but it should be significant before one can expect a system to reach the regime of universal structure. One counter-example is the periphery of heavy nuclei where the neutron density will extend further out than the proton density—seen e.g. elegantly through antiprotonic 208Pb and 209Bi atoms [16] (the term 'halo' is, unfortunately, also employed traditionally in those studies, but the physics is, of course, quite different)—without having any significant dynamical effect on the overall system in contrast to light halo nuclei.

A few more technical comments may be useful before proceeding; a more complete discussion can again be found in [13]. A short-range potential is one that falls off with distance more rapidly than r−2, i.e. r2V (r) → 0 for r → . The obvious measure of the range of the potential, given a halo particle of a certain binding energy, is the classical turning point R of the particle where its potential energy equals its total energy. This definition assumes that we are dealing with a two-body system; the case of one two-body subsystem of a many-body system is not so straightforward. In practice, potentials fall off sufficiently rapidly that we may approximate R by using quantities that are more easily accessible experimentally. For two-body halo nuclei, one can use the equivalent square-well radius that is related to the mean-square radii of the two components by 3/5R2 = 〈r21 + 〈r22 + 3.3 fm2, where the last term reflects the finite range of the nuclear force. The good halos found early on in nuclear physics, such as 11Li and 11Be, have probabilities of being clustered and of having halo particles outside of R that are all above 50% and this value could be used as a yardstick for singling out well-developed halo systems.

Another question that will return in the next subsection is how does one, in practice, decide how many 'clusters' a halo state should be divided into? For most systems this is not a practical problem, but a challenging example is given by the hypertriton, 3ΛH = n + p + Λ, which is bound by about 0.14 MeV with respect to break-up into d + Λ but where the deuteron in turn is bound by only 2.2 MeV and therefore in itself is quite a good halo. Even though three-body models must be used to give a good description of 3ΛH, we shall see that it is most naturally considered a two-body halo.

Now, if a system is well clustered and can be described in terms of a short-range potential, a halo state will appear once the binding energy is sufficiently small. The tricky part of the question of when halos will form is therefore whether sufficient clustering is present. In general, a component can only be expected to remain inert if the energy required to excite it or break it apart is higher than the interaction energy between components. On the molecular scale, where atoms are the natural 'building blocks' this is almost automatically fulfilled and one can expect molecular physics to provide interesting halo test cases. On the atomic scale, where electrons are added as active ingredients, the electrons in atom components are easily perturbed, but one may look for halos in systems of the type alkali-atom + e + nobel-gas-atom or electrons bound to a molecule with permanent dipole moment. Atomic halos should be rare and no good case has so far been identified in experiment. Finally, on the nuclear scale, where nucleons could serve as halo particles around a core, the conditions for clustering are less clear and therefore in a sense more interesting. Nuclei may provide us not only with well-developed halo states but also with systems intermediate between normal nuclei and halo nuclei. The key question is to what extent the core will remain inert? This subject is closely related to nuclear cluster physics [17, 18] and we shall return to it in section 4. Further adding to their interest is that, as will also become clear in section 3, halo states in nuclei are characterized by their special dynamical behavior.

There is as yet no clear answer to where nuclear halo states will occur. Nevertheless, several calculations have been made, the most recent ones [1921] within Hatree–Fock–Bogoliubov theory. The criteria for defining halos in heavier nuclei are typically different from the ones given above, see the more extended discussion in [20], but at least for even–even nuclei it appears that halos will be less numerous (and on a relative scale less extended) than for light nuclei. Part of this is due to pairing that has an important, but intricate, effect on halos [19], in certain cases decreasing sizes by mixing orbitals, in others enhancing them. Some of the necessary general conditions have been identified and will be mentioned in the next subsection. An important restriction is the mixing with other states that cannot be avoided if the local density of states is too high. This limits the occurrence of halos in excited states; the estimates in [22] for an excitation energy E* use a level distance of $D_0\exp (-2\sqrt {aE^*})$ (with D0 = 7 MeV and a = A/7.5 MeV) and conclude that s-wave neutron halo states should have binding energy $B< 270\,\mathrm {keV}\,\left ( A/Z \right )^2\exp (-4\sqrt {aE^*})$ and that the restriction for p-waves is $Z < 0.44 A^{4/3}\exp (-2\sqrt {aE^*})$ . For now, the main inference from these estimates is that nuclear halos will predominantly occur in ground states or at low excitation energy and therefore is a dripline phenomenon.

2.2. N-body systems

Several contributions to this symposium deal with the challenges posed by describing nuclear systems. For halo systems the requirements are to describe correlations in the systems well and to describe the large-distance behavior accurately. Not all theoretical frameworks can do this easily (as an example some work better in momentum space, whereas halos are described more naturally in configuration space). I shall focus here on lighter nuclei where few-body models can be used and start by commenting on the classification of few-body systems.

Knot theory in mathematics gives a well-established classification of linked systems including the famous Borromean rings: three linked rings where each pair is unlinked. A general system of N rings is called Borromean if each subsystem of two rings is unlinked and Brunnian if all subsystems with N − 1 rings is unlinked (this can be generalized further [23]). Following [24], this nomenclature was taken over in physics by replacing 'geometrically (un)bound in three dimension' with 'energetically (un)bound'. Although mainly used for three-body systems, one could therefore speak of N-body Borromean systems if all two-body subsystems are unbound. Examples of nuclear Borromean systems are 6He (α + n + n), 9Be (α + α + n), 11Li (9Li + n + n) and 45Fe (43Cr + p + p), but one can also find five-body Borromean nuclei such as 8He and 19B that consist of a core nucleus and four neutrons. Similar structures will be abundant for heavier dripline nuclei.

One can gain a first overview of the possible behaviors of an N-body system by going from the 3N spatial coordinates $\vec {r}_i$ to generalized hyperspherical coordinates [5, 25]. After separating out the three center-of-mass coordinates $\vec {r_{\mathrm {cm}}}$ , a single radial coordinate ρ is defined through

Equation (1)

where $ M = \sum _i m_i$ and m is chosen as a typical mass scale of the system (for nuclear halos: a nucleon mass). The remaining 3N − 4 coordinates are dimensionless and are generalized angles that incorporate the information on relative distances in the system. From the kinetic energy term one can extract a term depending only on ρ that appears in the same functional shape as a centrifugal barrier:

Equation (2)

Note that this contribution also appears for a system where all particles are in relative s-waves; the effective angular momentum given in table 1 therefore suggests that systems with more particles will be more confined. This can be understood from the following argument: if a wavefunction is allowed to extend away from origo it will spread out much faster in a higher-dimensional space (the volume increases more rapidly with distance) and the overlap with the binding potential close to the origo will therefore decrease; to counteract this and keep the total system bound it needs to stay small. The obvious loophole in this argument is that correlations in a subsystem may ensure binding; mathematically, this corresponds to having ρ large but one (or more) of the distances $\vec {r}_i-\vec {r}_k$ small (giving important contributions also from the 'angle' part of the Hamiltonian). I shall look at this in detail below for three-particle systems where one often introduces the hypermomentum K as a generalization of the two-body angular momenta, and where K = 0 corresponds to the most symmetric state with s-waves in all subsystems.

Table 1. The effective angular momentum for an N-body system.

N 2 3 4 5
ℓ* 0 3/2 3 9/2

There is no unique best way to measure the spatial extent of a halo. The tradition is to use the mean-square (or root-mean-square) radius; this is with the above choice of coordinates given as

Equation (3)

where the last term contains the contributions from the size 〈r2i of each of the components in the total system. This measure of extent emphasizes the tails and will therefore be quite sensitive to the dilute tails in halos. In several cases a more appropriate measure is the probability of finding a particle outside a certain distance (e.g. in a classically forbidden region) as we shall see when discussing the experimental probes of halos.

Since an N-body system without internal correlations (all particles in relative s-waves) appears as a two-body system with an effective angular momentum, we can start by considering two-body systems in the limit of a small binding energy B. This can be solved in the general case [26] and the expectation value of rn will go as

Equation (4)

i.e. it diverges in the zero-energy limit unless n < 2ℓ − 1. An angular momentum, of course, confines a system, but this can be compensated for by a higher weighting of the tail region. From the above analysis, we therefore deduce that the mean-square radius will go to infinity for vanishing binding for two-body systems with relative s- and p-waves and for three-body systems with K = 0 whereas all other systems remain finite. However, it should be noted that the only really pathological system is the two-body s-wave where the probability of being outside the binding potential can go to 1. This does not happen in any other case.

If a long-range repulsive interaction is present, e.g. the Coulomb field for nuclear proton halos, the tail of the wavefunction will be more suppressed at larger distances than for an angular momentum barrier. This makes halo formation more difficult and will prevent it altogether for nuclear charges above 10–20 (none of the identified one- or two-proton emitters will have halo character). On the other hand, it has been shown explicitly that nuclear deformation will not prevent the formation of halos [27].

We are interested in being able to compare also systems at finite binding. To do this across physics fields it is useful to introduce dimensionless scaling variables [15, 28]. For a two-body system we use the classical turning point that in practice is approximated by the R discussed above. The dimensionless mean-square radius is then simply 〈r2〉/R2. The momentum corresponding to a range R is of order ℏ/R giving a kinetic energy of order ℏ2/(2μR2), so a natural dimensionless binding energy is therefore μBR2/ℏ2. With this set of variables one has indeed scaling at low binding for two-body states with a good angular momentum ℓ. The resulting plot for identified halos will be shown later in figure 2.

It is less obvious how scaling appears in three-body systems. A more thorough discussion of this can be found in [13]; it suffices here to note that if a scaling radius ρ0 can be found one can introduce the dimensionless variables 〈ρ2〉/ρ20 and mBρ20/ℏ2. There are two slightly differing choices of ρ0; the first [15] proceeds in analogy to the way the radius ρ was introduced in equation (1) and makes use of two-body scaling radii Rik to define

Equation (5)

The second [28] builds on a deeper analysis of the case when all potentials are square wells and defines

Equation (6)

with μik being the reduced mass of the subsystem. If all masses and two-body scaling radii are the same, both definitions reduce to ρ0 = R.

The main features of the classification of three-body systems [28] are reproduced in figure 1 (the second definition of ρ0 is used here). The results are shown for three model systems corresponding to 11Li, 3ΛH and a three-boson system where the interactions in all cases are varied to give different binding energies; the large triangle and the filled circle correspond to the physical 11Li and 3ΛH. The uncorrelated case with K = 0 shows good scaling properties. When correlations are allowed in the wavefunctions, the main effect is to increase the binding energy whereas the radii are less affected. The resulting curves therefore lie to the right (or, equivalently, above) the ones for K = 0. The closest curve is the one for Borromean systems. Here again a universal behavior is seen for three different systems. However, the physical hypertriton has a bound subsystem (the deuteron) and is further away. That radii increase (for a given three-body binding energy) as more and more subsystems become bound is likely to be a general effect [29]. The arrows indicate the positions where a two-body subsystem goes from being bound to unbound. If the neutron–proton interaction is weakened the hypertriton system (filled circles) eventually becomes Borromean and approaches the other Borromean systems. The physical 11Li is, of course, Borromean and when the neutron–9Li interaction is increased nothing spectacular happens to the ground state (filled triangles), but a very extended excited state (open triangles) appears. This is an Efimov state that appears along the dashed line just before the 10Li subsystem becomes bound, increases in binding energy and finally increases in size again when becoming unbound with respect to n + 10Li once the latter has become bound.

Figure 1.

Figure 1. Three-body scaling plot for various systems as marked in the legend. The contribution to the mean-square radius from the three-body radial coordinate is measured against the scaling radius from equation (6) and displayed versus the scaled binding energy. The arrows indicate where two-body subsystems become bound and the dashed line (- - - -) indicates where Efimov states can appear. See the text for details. Courtesy of E Garrido.

Standard image

The Efimov effect [30] appears when a two-body subsystem has a very large scattering length a. This induces an effective potential proportional to ρ−2 in the three-body system for distances between R and a, a potential where successive excited states will scale in energy as well as extension. More details can be found in [25, 31]. These extraordinary states have been searched in various systems, including nuclei and few-atom molecules, and were finally observed indirectly in cold atom gases a few years ago. Most observations rely on the increase of the recombination rate of the many-body system seen when the binding energy of the Efimov state goes to zero; see [32] for a recent overview, but in a few cases states have also been produced directly at finite binding energy ([33] and references therein). So far experiments have probed the binding energy systematics rather than the spatial extension of the states. The one case where experimental information on sizes of loosely bound molecules is available is the 4He dimer and trimer where a very elegant experiment [34] has succeeded in determining the sizes as 5.2(4) and 1.1(5) nm, respectively. An excited He trimer Efimov state has been predicted and looked for several decades, but still not identified. In principle, much larger molecular halo systems have already been produced via manipulation of atomic scattering lengths in cold atom gasses and more cases are expected to exist also as isolated systems.

To sum up, the only truly pathological behavior occurs in two-body systems where the scattering length a can be arbitrarily large (and, in the bound case, a halo state forms with arbitrarily small overlap between the two bodies). A very large a induces the Efimov effect where the two-body correlations in a three-body system give rise to excited states with universal scaling properties and sizes reaching up to a. In contrast to Efimov states that appear at the two-body threshold, the Borromean three-body states that occur at the three-body threshold are far more moderate in size for a given binding energy. Three-body systems that appear above the Borromean curve in the scaling plot therefore have substantial two-body correlations.

Going on now to N-body systems, recent calculations for N = 4,5,6 identical bosons [35] show that the systems remain finite even for vanishing binding energy. The most important correlations remain the two-body ones, even though higher-order correlations can also play a role for excited states. The overall sizes seem to decrease as N increases, but one can still observe systems that are significantly larger than the scale set by R. Similar results are likely to be present in the nuclear case and would be relevant for heavier nuclei along the dripline than the ones accessed currently. If the nuclei there are best described as N-body Borromean systems with N larger than three, one would expect halo formation to be quenched. This picture with several 'active' neutrons around a core is probably the reason, seen from few-body models, for the earlier mentioned decreased size of halos when pairing mixes orbitals.

The hypertriton was used above to clarify the classification of three-body systems and is best thought of as a Λ-halo around a deuteron. Hypernuclear physics has probably more halos to offer [36], but now with neutrons forming the halo rather than the Λ-particle. Among the interesting systems are 6ΛHe, where the outer neutron is expected to extend further out than in 6He, and 7ΛBe that with a binding energy of around 0.3 MeV below the 5ΛHe + p + p threshold probably is the best two-proton halo one could hope for. A recent experiment [37] indicated that 6ΛH is also bound and quite close to the 4ΛH + n + n threshold.

2.3. Halo excitations

The typical binding energy scale for nuclear halos is of the order of 100 keV (less than about 1 MeV in light nuclei, decreasing with the mass number as A−2/3) and for molecular halos 1 μeV. There have been very few studies of the dynamics of molecular halos, so the discussion will from now on concentrate on nuclei.

States with halo structure can be excited in many different ways; the ones of specific interest here are the ones that involve the halo degree of freedoms. Since the halo extension depends so crucially on binding energy, one cannot expect the structure to be maintained under e.g. rotation or isospin changes and we need to carefully look at which states can be reached in excitations.

Strong interactions typically require more dedicated treatments, whereas electroweak interactions can be treated perturbatively with operators that may change spin and/or isospin, and may contain factors rλ that enhance the tail. As will become obvious from the following section, halo excitations very often are dominated by transitions to the continuum and also often dominated by the tail properties. As will also be seen, in particular from the electromagnetic probes in section 3.3, these are not qualitatively new features in nuclear physics but their magnitude puts them on a quantitatively new level for halos.

Due to the importance of the continuum, the question of the continuum structures seen in excitations must be considered carefully. A good example is the E1 strength distribution of neutron halos that is now understood as being single-particle strength going mainly directly to the continuum [38] and not a semi-collective soft resonance, see also [39]. As shown in detail in [40] for the case of the 11Be E1 strength, one may choose to ascribe some of the strength to resonances in the continuum. This particular work followed the definition of Berggren [41] of a resonant state and his elucidation of how one can go from a description entirely in terms of a continuum to an equivalent description where resonances are included explicitly. An important outcome of Berggren's analysis is that one, in general, may expect a remaining non-resonant continuum contribution; this is very much the case for the 11Be E1 strength and can be expected also to be an important feature of other halo excitations. The research theme of discrete states embedded in a continuum is an important one in contemporary nuclear physics [42] and will also be covered in these proceedings [43].

3. Experimental probes of nuclear halos

This section will present some of the different experimental ways of testing and characterizing nuclear halo states. It is not possible to make a complete coverage of this large field that will also be the (partial) subject of quite a few of the other contributions to this symposium. An early overview of the different types of probes of halos was given in [26]. As remarked there, one may classify the probes according to how much they emphasize the tail, e.g. which power of r they correspond to. One should remember that the probability of 'staying inside the binding potential' remains finite except for s-wave neutrons at vanishing energy; the large spatial extent of halos is a result of quite a dilute very extended tail (cf the discussion after equation (4)) and some probes will also be quite dependent on core properties.

Combining information from different experiments gives a more accurate picture of a given halo state. This section will focus on what information may be obtained from the different experiments and whether there are any generic effects or signals of the halo structure. The properties of some individual states will be summarized in section 4.

3.1. Ground-state properties

Due to the extreme sensitivity to binding energies, it is important to know the mass of halo nuclei accurately. In most cases it is sufficient that the binding energy be determined with an uncertainty of 10 keV or less, but this is a rather challenging goal for nuclei close to the driplines. General overviews of the techniques used in mass measurements of radioactive nuclei can be found in [44, 45]. The masses of halo nuclei were at first measured via nuclear reactions or various time-of-flight measurements [46], but Penning traps have by now also been brought into use [47] and give an important step forward in accuracy as well as precision. The lightest halo nuclei have by now been measured with sufficient accuracy and the challenges lie in producing neutron dripline nuclei above Be in amounts that enable determination of the masses of heavier halo candidates. A particularly important challenge is the mass of 19C that at the moment is deduced from reaction measurement assuming it to be a good halo state, rather than from direct experiments.

Optical measurements on radioactive beams [48] give access to many nuclear ground-state properties, including charge radii, spin and electromagnetic moments. These can give very precise information, often on a level beyond what current nuclear theory can predict. The charge radii are of course very valuable for a direct determination of the sizes of proton halos, the experiments being sensitive to the mean-square radius of the total charge distribution. The best example of this is the charge radius of 17Ne [49] that was found to be clearly above the radius of heavier Ne isotopes although the magnitude of the effect, an increase in rms radius by less than 0.1 fm, is clearly below what is found for good neutron halo nuclei. This is, of course, consistent with the confining effect of the Coulomb barrier that already in Ne is very noticeable. Accurate measurements of the charge radius of light neutron halo nuclei have also appeared during the last few years [47]. Due to their high accuracy they are important in giving constraints on core modifications, they can furthermore for multi-neutron systems test the correlations in the overall system through the sensitivity of the total charge radius to the movement of the charged core around the center of mass.

Magnetic dipole moments and electric quadrupole moments have also been extracted for some halo nuclei (quite apart from the spin, an even more basic property). The accurate measurements must be coupled to model calculations to give the optimal information on the system; two good examples are the magnetic moment of 11Be that is μ = −1.6816(8)μN [50] and the electric quadrupole moment of 11Li whose ratio to that of 9Li, |Q/Qcore| = 1.088(15) [51], is similar to that of the rms charge radii for the two nuclei. In general, knowledge of the moments of a halo candidate and the 'bare core' nucleus can give important information on how inert the core is in the halo state.

3.2. Beta decay

The half-life of a radioactive nucleus also belongs to its ground-state properties, but will only in exceptional cases carry a clear signature of a halo structure. Beta decay probabilities depend on the overlap between the initial and the final state wavefunction and the effects of a halo on an overlap are in most cases below a factor of two, which means that one needs a good understanding of the core structure of the initial and final states in order to draw conclusions. However, beta decay brings information on halos in other ways (see [14, 52, 53] for more detailed reviews).

Firstly, beta decay may help in pinpointing the exact configuration of a halo. A good example is 11Li where already the total half-life indicates that the two last neutrons must reside partly in the sd-shell rather than only in the p-shell [54]. Individual decay branches, of course, can be used to strengthen this argumentation [52]. Secondly, there are indications that the halo and core beta decays decouple at least in some cases (it is not yet clear whether this will be a general feature for good halo states). The prime example of a core decay that reappears in the decay of the halo nucleus is that of 12Be where more than 99% of the decay goes to the 12B 1+ ground state and one correspondingly finds that most of the 14Be decays goes to a slightly unbound 1+ state in 14B [12, 52, 55].

An even more convincing case would be a decoupled halo decay. This appears to occur at least for two-neutron halo systems as beta-delayed deuteron emission. The current understanding of this decay mode is that it proceeds directly to the continuum [52, 53], a decay of the two halo neutrons directly to a deuteron in the periphery of the total system. Two cases have been established so far; for 6He, the intensity is low due to a cancellation of contributions from the inner and outer parts of the wavefunction, whereas 11Li, where the energy spectrum was measured recently [56], offers the possibility of a more stringent test once the final state interaction between the deuteron and the 9Li core is understood better. It will be interesting to see whether other halo nuclei will give similar unique decays. At least for the two known cases it is clear that the contribution from the non-resonant continuum is essential for describing the decays.

3.3. Electromagnetic processes

Electromagnetic transitions of type E(λ) or M(λ + 1) have operators that contain a factor rλ which enhances the tail behavior and makes these processes very sensitive to halo formation. It is worth noting that halos in excited states can be probed as well provided that one can single out transitions that involve these states. The emphasis on the wavefunctions at large radii was known also before reaction experiments indicated the existence of halo states, and several earlier papers identified the main features of what we now recognize as halo signatures.

For transitions between bound states the classic case is that of 11Be where the E1 transition between the first excited 1/2 state and the 1/2+ ground state could only be reproduced if the spatial extent of the wavefunctions was taken into account [57]. Transitions between a bound state and the continuum were known to be important for the deuteron, but the case of proton radiative capture into a spatially extended state was also clearly understood for 8B and the first excited state in 17F [58, 59]. In these cases, direct capture, i.e. contributions from the non-resonant continuum, are more important than resonance capture in line with the discussion in section 2.3. The proton radiative capture was investigated early on due to its importance for nuclear astrophysics, but a similar sensitivity can be expected in neutron radiative capture [26] and should again give a low-lying non-resonant peak [60].

Electromagnetic dissociation is the inverse reaction to radiative capture and will therefore be as sensitive to halo structures. It is now recognized as a key probe for nuclear halo states and an important dynamical consequence of the special structure, as was first pointed out in [3] and observed experimentally shortly afterwards [61], see also [39]. The prediction in [3] was based on sum rules for the E1 strength. These are easily generalized and give for a two-cluster model [62], where the clusters have charge and mass number Zi and Ai and distance r between them, a contribution to the energy-weighted sum of

Equation (7)

and for the non-energy-weighted sum

Equation (8)

Compared to a normal state, a halo state will therefore give more strength that appears at lower excitation energy than usual and if the specific contribution to the E1 strength from the halo degree of freedom can be extracted, one can derive its spatial extent. The E1 strength will carry the dominant signal for neutron halos, whereas higher orders will also be affected for proton halos. Concerning the strength distribution, the simple structure of halos implies that simple analytic models [6365] can be expected to give the main features. Simple analytic expressions for the photodissociation cross-sections valid for loosely bound nuclei have also been derived [66].

As discussed above and in complete correspondence to radiative capture, the dissociation reactions mainly involve the non-resonant continuum. Their importance will be clearly seen in the next section; one good example is the identification of 31Ne as a halo candidate based on its large one-neutron removal cross-section [67].

3.4. Nuclear reactions

Most of the studies on halos have been performed using nuclear reactions. I can here give only a brief overview on this large field that is also covered in several other contributions to these proceedings, in particular [68]. Detailed reviews on the reaction theory of halo nuclei can be found in [4, 13, 6971].

It is customary to classify the reactions according to the beam energy, low-energy reactions taking place around the Coulomb barrier, intermediate energy reactions around the Fermi energy and above this the high-energy reactions (extending to relativistic energies), where the smaller nucleon–nucleon cross-section and the shorter interaction times make the reaction mechanisms simpler. The reaction mechanisms do evolve gradually as the energy increases, so a strict division into the three regions is not possible, but the rough dividing points are somewhat above 10 MeV u−1 and around 100 MeV u−1.

As realized essentially from the start, the 'halo-removal' channel (one-neutron removal for one-neutron halos, two-neutron removal for two-neutron halos, etc) will, for both strong and Coulomb interactions, have a significant cross-section and will furthermore be clearly influenced by the halo structure. The main focus has been on this channel; this is fully justified but has implied that possible information from other channels has often been neglected.

3.4.1. Interaction cross-section.

The experiments that triggered the interest in halos [1, 2] measured the total interaction cross-section for different isotopes and observed a clear enhancement for halo nuclei. The key point is that the strong interactions between nucleons make reactions so likely that nucleons 'shadow' each other so that compact systems with more shadowing will have reduced cross-sections. (This is in contrast to electromagnetic or weak interactions that only observe the total 'charge' of a nucleus.) As gradually became clear it is important to have a correct understanding of the reaction mechanism in order to extract reliable radii for the halos [72]. This is now the case for reactions at high energy; a compilation of matter radii extracted from such experiments can be found in [73]. There are many attempts at extending the theory also to intermediate and low energies, but for the moment extracted values from these energy ranges should be treated with more caution. Nevertheless, the large deduced radius for 22C [74] clearly identifies this nucleus as a new halo candidate even though the uncertainty of its radius is considerable.

For good halo systems it was pointed out in [75] that the 'halo removal' cross-section at high energy will be equal to the difference in reaction cross-sections for the halo nucleus and the core nucleus, quite a direct reflection of the decoupling of the halo system into halo and core parts. In line with this, the charge-changing cross-sections for halo nuclei and their core have been found to be essentially the same [76].

3.4.2. High to intermediate energy probes.

Many types of experiments have been developed for use at high and intermediate energy. Among the first were measurements of the transverse momentum distributions [77, 78] of reaction products from break-up reactions. Shortly after longitudinal momentum distributions of the core fragment [79] became available, these are less affected by the reaction and—if the reaction mechanism is disregarded—can be easily interpreted (a large spatial extension corresponding to a narrow momentum width). Surveys of longitudinal momentum distributions have been carried out for many neutron-rich nuclei [80, 81]. The reaction mechanism, of course, needs to be taken into account as well in order to extract quantitative information. On the experimental side, an often needed refinement is a more exclusive measurement where the final state of the core is detected (of course, only relevant when the core has several particle bound levels). A gating on specific final states is needed in order to get clean distributions as first shown for 11Be [82].

Most recent works have focused on core longitudinal momentum distributions in the halo-removal channel. Although not as widely appreciated, it is also valuable to look at neutron momentum distributions at high reaction energy for processes where the core is removed by break-up [83] or kicked out by a Coulomb field (this should, e.g., give a very clear signature in the neutron–neutron correlation [84]).

Experimental setups are now available at most facilities, allowing complete kinematics experiments to be carried out. Here, all outgoing fragments from the break-up are recorded combined with gamma-ray detection at the target position; in several cases, with proton targets the recoiling proton has also been detected. This allows reconstruction of the excitation energy in the final state as well as of various distributions in sub-systems; several examples will be given in the next section. This, of course, puts more stringent demands on the reaction models; see [13, 69] for the status for high-energy experiments and [85] for a recent comparison of different break-up models at intermediate energy.

A particular example is that of elastic scattering at high energy on a proton target. This well-established method has been employed in several experiments for the lightest halo nuclei and gives quite an accurate description of the density profile of the nuclei allowing for an independent determination of the matter radii. A recent example is the work on heavy Be isotopes leading up to a determination of the radius of 14Be [86].

3.4.3. Intermediate- to low-energy probes.

The already established many different types of lower-energy reactions have, essentially, all been applied to halo nuclei once they became available as secondary beams at lower energy. This includes in particular elastic scattering, fusion reactions, transfer reactions and break-up reactions; see [6971] for reviews. This field is still evolving and is still to some extent focused on the identification of halo signatures rather than extraction of halo properties. I will give just two different examples of this.

In elastic scattering at low energy the nucleus has time to adapt during the collision, giving e.g. unique polarization effects that do not appear at higher energy. One example is the elastic scattering of 11Be on Zn [87] where a careful comparison with the also measured elastic scattering of 9,10Be gave a clear halo signature.

Transfer reactions are a powerful tool for nuclear structure studies and secondary beam techniques have evolved so that it has now become possible to do e.g. (p,t) reactions on both 8He [88] and 11Li [89]. The basic theoretical procedures are still being critically discussed, see, e.g., [90]. One recent interesting suggestion is to focus on the ratio of angular distributions for elastic break-up and scattering [91] that appears less sensitive to the reaction mechanism giving more direct access to the halo properties.

4. Knowns and unknowns of nuclear halo states

The very first experiments on nuclear halo states gave only the rudimentary properties of the states, but developments in experimental techniques (both in production and in detection methods) have given an increasingly more detailed characterization of the lightest states such as 11Li and 6He. First observations of halo states typically take place at in-flight facilities, but ISOL-based facilities play an important role in refining our understanding in the later stages of experiments. The new generation of radioactive beam facilities [92], starting from RIBF in RIKEN, will allow studies to continue in new mass regions; several promising halo candidates have already been identified in recent years.

This section will first give an overview of what is currently known about nuclear halo states with a focus on the most pronounced cases. The discussion will hopefully indicate what is needed in order to establish future halos. It is very encouraging that the field has reached a stage where one can cross-check results, e.g. on matter radii, by comparing the results obtained by different methods. Nevertheless, there are still several unknowns. The last subsection gives a list (with a personal bias) of the open questions remaining in the field.

4.1. Established halo states

In order to use a few-body picture for the halos the amount of 'configuration mixing' should be limited, i.e. there should in the nuclear case be mainly one component in the wavefunction with one or two nucleons around a core (that very often, but not necessarily, will be in the ground state). This appears to be fulfilled for most of the nuclei considered below. Scaling plots updated with the latest experimental information are shown in figures 2 and 3. For the three-body halos, equation (5) is used to define ρ0; the alternative definition in equation (6) gives a ρ20 that for the nuclear two-neutron halos is smaller by about a factor of 1.7. It may be interesting to note that systems where the probability of being outside of the potential range is higher than 50% correspond to a value just below 2 for the scaled square radii. As can be seen in the figures, this limits the number of pronounced halos.

Figure 2.

Figure 2. Scaling plot for two-body halo systems. The filled circle denotes the deuteron, the filled squares denote nuclei where radii were extracted from experimental interaction cross-sections, the open squares are simple model estimates and the open circles are theoretical calculations. See the text for details.

Standard image
Figure 3.

Figure 3. Scaling plot for three-body halo systems. The filled squares denote nuclei where radii were extracted from experimental interaction cross-sections and the open circles are the results of theoretical calculations. See the text for details.

Standard image

For easier reference, some properties of selected halo systems are collected in tables 2 and 3. The tables list the root mean square radii for the total system as well as for the core nucleus, the components of the system and the binding energy [93] for the halo particle(s). Except where discussed explicitly, in the following the experimental values for radii are taken from [73, 94]; in the cases when several analyses have been performed, the 'few-body' results (that take correlations in the system into account) are used. For two-body halos the angular momentum of the halo particle is given; for three-body halos the possible angular momenta for a halo neutron with respect to the core is given. The deduced mean square radius of the halo is listed in the end. Since selections between different data had to be done in several cases, the following detailed discussion should be consulted to obtain a complete overview for a specific state. The assumptions made are also discussed there.

Table 2. Properties of some two-body halos.

System Composition B (MeV) rrmscore (fm) rrmstot (fm) r2〉 (fm2) r2〉/R2
d n+p 0 2.225 1.9754(9) 15.61(2) 4.2
8B 7Be+p 1 0.136(1) 2.31(5) 2.50(4) 14(3) 0.93(18)
11Be 10Be+n 0 0.502 2.28(2) 2.90(5) 44(4) 2.9(3)
11Be* 10Be+n 1 0.182 2.28 40 2.6
15C 14C+n 0 1.218(1) 2.30(7) 2.50(8) 21(8) 1.4(5)
17F* 16O+p 0 0.105 2.72 28 1.5
19C 18C+n 0 0.58(9) 2.82(4) 3.23(8) 58(11) 2.9(6)
31Ne 30Ne+n 1 0.3(2)a ? ? 60a 2.3a
35Mg 34Mg+n ? 1.0(2) 3.23(13) 3.40(24) 50 2(2)

aSee the text.

Table 3. Properties of some three-body halos.

System Composition B (MeV) rrmscore (fm) rrmstot (fm) ρ2〉 (fm2) ρ2〉/ρ20
6He 4He+n+n 1 0.975 1.58(4) 2.54(4) 28.6(1.3) 1.84(10)
11Li 9Li+n+n 0, 1 0.369(1) 2.30(2) 3.53(10) 89(8) 3.4(3)
14Be 12Be+n+n 0, 1, 2 1.26(13) 2.59(6) 3.10(15) 54(13) 1.7(4)
17B 15B+n+n 0, 2 1.34(17) 2.59(3) 2.90(6) 42(6) 1.3(2)
19B 17B+n+n 2 ? 1.0(4) 2.90(6) 3.11(13) 41(16) 1.1(5)
17Ne 15O+p+p 0, 2 0.933 2.44(4) 2.75(7) 39(7) 1.3(3)
22C 20C+n+n 0 ? 0.4(3)a 2.98(5) 5.4(9) 460(210) 12(5)

aSee the text.

4.1.1. The deuteron.

The deuteron is obviously composed of a proton and a neutron and in this sense it has no core. Historically it is, as remarked in [8], 'the forerunner of all halo states' and it is where the main reaction mechanisms for halos were discovered. (In the words of Gregers Hansen [95] it is 'the mother of all halo nuclei and certainly the only one with both a proton and a neutron halo'.) The rms radius has been measured accurately from optical measurements [96] and is 1.97535(85) fm. Its exact position in figure 2 depends on the choice of R, which is of the order of 1.9–2.0 fm. Recent reaction work on deuterons has focused on relativistic scattering experiments, but a review of low-energy deuteron break-up can be found in [97].

4.1.2. 11Li.

This was among the first identified halo states [2] and is still the archetype of a two-neutron halo. Much of the existing data have been summarized and compared to a three-body model in [98]. The following brief discussion mainly illustrates the large number of data available on 11Li.

The total matter radius is extracted [94] as 3.53(10) fm from the interaction cross-section and 3.71(20) fm from high-energy elastic scattering [99]. The two radii determinations are consistent, but since the radii of 9Li also differ slightly from the two methods the deduced values of 〈ρ2〉/ρ20 are essentially identical: 3.4(3) and 3.5(6), respectively. The former is the value included in table 3.

The magnetic dipole moment is slightly larger than for the core 9Li (closer to the single-particle value); as mentioned above this is also the case for the electric quadrupole moment. The charge radius data may indicate the need to include core excitations [100]; a more extensive discussion is given in [101]. The charge radius is basically consistent with the radii determined from reaction experiments, but the simple picture of an inert 9Li core and two halo neutrons coupled to angular momentum zero must be modified slightly to make all the data fit [98, 101].

A naive filling of orbits from a simple shell model scheme would have placed the two halo neutrons in p-waves. The need for s-wave components was deduced already from the beta-decay half-life [54], but also arises from an analysis of the reaction experiments [102]. Rather direct evidence for different parity states was obtained from the observation of anisotropic angular distribution between fragments [103]. The s- and p-wave components have roughly equal weight, but smaller contributions of e.g. d-wave could also be present.

In the halo break-up channel, 9Li has a longitudinal momentum distribution with full-width at half-maximum (FWHM) of the order of 40–50 MeV/c depending on the target and the beam energy [104]. An analysis of two-neutron interferometry [105] gave an average distance between the two neutrons of 6.6(1.5) fm. The E1 strength distribution measured in detail [106] led to a deduced value for the rms distance between the core and the center of mass of the two halo neutrons of 5.0(3) fm. All in all, the combined data indicate that the two halo neutrons tend to be on the same side of the core, an asymmetry made possible by the configuration mixing of s- and p-waves.

Through a 11Li(p,n)11Be* reaction of 64 MeV u−1 the isobaric analogue state of 11Li was observed to have a smaller Coulomb displacement energy than for other Li isotopes [107], consistent with calculations including the halo [108]. See also [109] for a general discussion of the relevant isospin multiplets.

4.1.3. 6,8He.

The nucleus 6He was also among the first halo states to be identified [1]. Its structure is well understood with the two neutrons being in p-orbits outside the alpha particle core. It is often used for detailed checks of our understanding of reaction mechanisms and other effects. It has been studied extensively in complete kinematics experiments at GSI (Helmholtzzentrum für Schwerionenforschung), and very complete data on its behavior in reactions are now available [110].

Adding two neutrons to 6He gives 8He that is considerably more bound. It appears to be better thought of as four neutrons outside an alpha particle core rather than two neutrons outside 6He. Its spatial extension is not as large, and it is probably better to consider it a skin nucleus [111].

Also here there is agreement between total matter radii extracted from interaction cross-sections and high-energy elastic scattering. The matter radius of 8He is only slightly larger than that of 6He and the charge radius is actually smaller; this is understood in terms of the correlations between the halo neutrons in 6He [112].

For both 6He and 8He, many experiments have looked at (in)elastic scattering, fusion reaction and transfer reactions.

4.1.4. 11Be.

Following the deutron, this is the first and most studied one-neutron halo. A particular feature is that the first excited state also is a halo, but with the neutron in a p-wave rather than an s-wave (a simple estimate for the radius of the excited state is given in table 2). This special structure is reflected in the strong E1 transition between the two states [57].

The matter radius is extracted as 2.90(5) fm from interaction cross-sections [94] or as 2.91(5) fm with the alternative analysis in [73]. The deduced rms value for the distance between the core and halo neutrons is then 6.6(3) fm, which can be compared with 6.4(7) fm extracted from Coulomb dissociation at 72 MeV u−1 [113]. Later determinations from other Coulomb dissociation experiments tend to give lower values for the distance [114, 115], but the charge radius measurement rather indicates a distance of 7.0 fm [116].

The longitudinal momentum of 10Be in the one-neutron removal channel has a FWHM around 45 MeV/c [104]. There are several excited states in 10Be and by gating on observed gamma rays, detailed studies have been performed of the one-neutron removal reaction leading to specific final states [82, 114].

The core is not completely inert and the wavefunction will contain a component with a 2+ excited core. Several experiments have addressed this issue and by now a consistent understanding of the size of this effect has been obtained [117]; about 70% of the wavefunction will be the 'classical' one with a 10Be ground state.

4.1.5. 8B.

The large spatial extension of the wavefunction of the p-wave proton in the ground state of 8B was, as mentioned above in section 3.3, already noted early in connection with studies of proton radiative capture [58]. The 7Be(p,γ)8B remains astrophysically interesting which has given rise to much experimental activity on this nucleus.

Due to the combined effect of the Coulomb and centrifugal barriers its radius is not as pronounced as for 11Be even though the binding energy is low. Several detailed reaction studies have been carried out, e.g. a complete kinematics experiment [118] that determines the core excited state component to be 13(2)%. One can also mention the measurement [119] of the electric quadrupole moment of +64.5(1.4) mb, a value that seems to be sensitive to the extended proton tail.

4.1.6. 17F*.

The very large spatial extension of the wavefunction of the s-wave proton in the first excited 1/2+ state in 17F was, as for 8B, noted from proton radiative capture; see in particular [59] for a very clear exposition and [120] for a later calculation within the continuum shell model. The recognition that this corresponded to a halo state [26] prompted investigations of possible effects in beta decay into this level. A clear deviation of mirror symmetry was seen, but the theoretical interpretation also depends on the description of the mother nucleus, 17Ne, and gives only indirect information on the state, see the review in [52]. The radius quoted in table 2 is a simple model estimate. The special status of this state was apparently rediscovered in [121].

4.1.7. 17Ne.

This two-proton halo candidate most likely has a well-defined core in 15O. The beta decay data may indicate changes of orbit occupation compared to the mirror decay of 17N, see the discussion in [52] and the detailed calculation within the shell model embedded in the continuum approach [122].

As already mentioned, its charge radius has been measured [49] and is in agreement with the matter radius of 2.75(7) fm extracted from the interaction cross-section [73]. This is probably the best candidate for a two-proton halo in normal nuclei, but has a moderate spatial extension.

4.1.8. 14Be.

The binding energy of this two-neutron halo candidate is above 1 MeV. Its matter radius has been extracted as 3.10(15) fm from interaction cross-sections [73] and 3.25(11) fm from high-energy elastic scattering [86], again in good agreement.

It seems increasingly likely that this system has parallels to 8He in the sense that 12Be cannot be considered an inert core. No experiment has so far reported on core excitations in two-neutron removal reactions, but 12Be is known to have bound 2+, 0+ and 1 states and in particular the excited state 0+ state is very interesting in this respect. More experimental information is needed to clarify our understanding of this interesting system in between normal nuclei and the pronounced two-neutron halos.

4.1.9. 15C.

The one-neutron binding energy is here just above 1 MeV, but the last neutron seems to be a good single-particle s-state and the system is therefore a good example of a state in between normal nuclei and well-developed one-neutron halos. This is reflected both in its interaction cross-section and in the width of the longitudinal momentum distribution after one-neutron removal; note that this nucleus and 22N appear to have quite a similar behavior in the general surveys [80, 81].

4.1.10. 19C.

The first indications of a halo in this nucleus came from the longitudinal momentum distribution after one-neutron removal with an FWHM of only 44(6) MeV/c [123]. The direct measurements of the binding energy have too low precision and the currently used value of 0.58(9) MeV is derived e.g. from an analysis of Coulomb dissociation in the one-neutron removal channel assuming the state to be a good halo [124]. A direct precise measurement of the binding energy would be very helpful.

All reaction data are consistent with this system being a good single-particle halo. The observation of two gamma rays [125] from excited states in 19C indicates an interesting low-lying structure (compatible with expectations from shell models) and incidentally puts a lower limit on the binding energy.

4.1.11. 17,19B.

The matter radii of the two heaviest boron isotopes were measured in [126]. A few more reaction studies have been carried out on 17B (e.g. demonstrating the presence of a bound excited state just below 1.1 MeV), but very little is currently known about 19B. Two different values have been extracted for the matter radius of 17B: 2.90(6) fm from the optical limit and 2.99(9) fm from a few-body approach [73]. The former is used in table 3; if the latter is used the value of 〈ρ2〉/ρ20 will increase to 1.6(3). It could well be that 17B is a more extended system than 19B in spite of being more bound. However, one should note that it was suggested [126] that 19B is better thought of as having four valence neutrons around a core, so the values derived from a three-body model could well be inapplicable. More data are clearly needed on 19B, but it is striking that we in both 8He and 19B, the two nuclei investigated so far where both one- and three-neutron removal lead to an unbound system (five-body Borromean nuclei), observe indications of a four-neutron structure rather than a two-neutron structure built upon the intermediate nucleus. (A similar situation could actually be present also in 14Be.) This is, of course, relevant for an extrapolation to much heavier neutron dripline nuclei where five-body Borromean (and seven-body, etc) nuclei will be very common.

4.1.12. 22C.

This N = 16 nucleus is quite an interesting system. The radius of 5.4(9) fm extracted from intermediate energy reaction cross-sections [74] is very large and would make this by far the largest nuclear halo known to date, see table 3. However, it seems likely that the 5.4 fm is an overestimate (note that the error is quite significant), since momentum distributions from two-neutron removal on a carbon target [127] do not show the same extreme result. Future experiments should be able to sort this out: a radius above 5 fm should give rise to a very large Coulomb dissociation cross-section unless there happens to be a strong anti-correlation between the two halo neutrons (so that the core and the center of mass of 22C almost coincide). On the other hand, should this large radius turn out to be correct it would indicate that there is a very strong core-neutron substructure, cf the discussion around figure 1. The currently known binding energy is much too uncertain, and the value used in table 3 and figure 3 is a guess reflecting the prejudice that the true value must be below 1 MeV.

4.1.13. 31Ne.

This nucleus was first observed at RIKEN [128], and it is also there that one recently, after the start of RIBF, has been able to determine its Coulomb cross-section [67] and interaction cross-section [129]. Both show the typical enhancement for a good halo state. Most properties of this nucleus have not yet been measured with a sufficient accuracy, but it is believed (from systematics and the analysis of the Coulomb break-up) that the halo neutron is in a p-wave. Several theoretical calculations on this interesting system are already available.

Its binding energy has an uncertainty of more than 1 MeV, but it should be within a few hundreds of keV of the threshold to fit the available reaction data, so I have arbitrarily put it at 0.3(2) MeV. The order of magnitude of the one-neutron Coulomb dissociation cross-section indicates a halo radius of the order of that in 19C and a scaled extension that is slightly lower (due to the larger core nucleus). I have used this for the rough estimates in table 2.

4.1.14. Heavier nuclear states.

There is by now sufficient experimental information available to conclude that we have identified all possible pronounced ground-state halos up to about mass 30. Above this mass we do not expect proton halos to be very extended, and current information on the neutron dripline candidates is quite limited. However, a recent reaction experiment [130] has pointed to 35Mg (extracted radius of 3.40(24) fm with a core radius of 3.23(13) fm) as a possible interesting nucleus. The precision must be improved in order to draw a definite conclusion based on interaction cross-sections alone.

The effect of the halo neutron(s) becomes lesser and lesser on a relative scale as we go to higher masses, which explains the greater requirement for precision on the cross-sections. It may be easier to detect halo states in the halo-removal channel via the Coulomb dissociation, see section 3.3 above and [39]. Some of the interesting nuclei that could be studied in the future in this way are the heavy Mg and Si isotopes.

4.1.15. Halos in excited nuclear states.

It is more difficult to extract information on halos in excited states, but as mentioned above it is possible in electromagnetic transitions (going to or from the state). The other possibility is to look for signals in nuclear reactions that populate the state. Recently, it was suggested [131] that a general signal may be an enhanced diffraction radius in intermediate energy scattering. Transfer reactions into a candidate state can, of course, also provide information if sufficient knowledge is available about the reaction partners.

Apart from the excited 1/2 state in 11Be and the 1/2+ state in 17F mentioned earlier, other interesting halo candidates are the 1, 2 states in 10Be [132] if their configuration is an s-wave neutron around a 9Be core. Two-neutron halos in excited states may also exist, e.g., in 12Be [133], but for most of these candidates there is insufficient experimental evidence. More candidates are listed in tables I and II in [13].

One special state where initial experiments have been performed is the 0+, T = 1 excited state in 6Li [134] that should have a neutron–proton halo and has been probed via pionic fusion [135] as well as 1H(6He,6Li)n reaction [136]. It is the isobaric analogue state of the ground state of 6He.

4.1.16. Halos in other systems.

Theoretical predictions for hypernuclear halos have also been performed in several cases. Predictions for the hypertriton [137] are included in figure 2. Experimental information is very hard to obtain and at the moment only binding energies are known, with considerable uncertainties, and nothing has been established concerning the spatial extensions.

Theoretical calculations for the 4He dimer and three trimer states [138] are included in the scaling figures. The other few-atom molecular states that appear in current cold atom gas experiments can have even larger sizes that furthermore can be tuned via an external applied magnetic field. They have not yet been isolated as single molecules, but would in the scaling plots occur along the ℓ = 0 line in figure 2 and along the dashed diagonal in figure 1.

4.2. Open questions

Halo physics has made impressive progress during the 25 years since they were discovered. Nevertheless, there are still several open questions we need to address. The following incomplete list contains the main ones that came to my mind when preparing this paper.

  • Where will nuclear halos heavier than the currently known ones occur in the nuclear chart? Somewhat coupled to this is the question of how much single-particle character remains in heavier nuclei?
  • Are there dynamical characteristics of skin nuclei?
  • Does closeness to continuum promote (or stabilize) halo cluster structures? For the general case of clusters this may be the case [139]; recent work [140142] is starting to establish quantitative criteria for clustering. The findings so far that most halo candidate states (with low angular momenta) end up having a good halo structure may be a 'selection bias' since all the cases are among the light nuclei, but the continuum could also play a more active role.
  • When is the concept of a resonant continuum useful? As seen above, the concept of a non-resonant continuum is essential in order to understand all aspects of electromagnetic processes and beta decays involving halos. Nuclear processes are less transparent, but the final state structures there should also be treated with great care. (Even though a resonance fit may describe the data it will not necessarily correspond to the physical reaction mechanism [84].)
  • Can we find ways to experimentally study the hypernuclear halos?
  • It would be very instructive to have a detailed experimental characterization of one of the halos in atomic and/or molecular physics.

The first decade or so of halo physics established the main concepts and identified the main cases occurring among light nuclei. Apart from a consolidation and refinement, reflected e.g. in the cross-checks on the spatial extension that have now been successfully made in many cases, the intervening years have also seen the concept being established in molecular physics and the links to the Efimov effect clarified. During the last few years, first experiments at a new generation of radioactive beam facilities have yielded new nuclear cases and more experimental information will soon be available. This may allow us to check our understanding of the conditions for halo formation and thereby complete our picture of the phenomenon.

Acknowledgments

I am grateful to the many colleagues who have during the past decades improved my understanding of halos and delivered the impressive experimental results reviewed here. I thank in particular Aksel Jensen and Björn Jonson for many enlightening discussions.

Note added in proof:

The universal aspects of halo nuclei have been the subject of a recent review [143].

Please wait… references are loading.
10.1088/0031-8949/2013/T152/014001