<<

arXiv:nucl-th/0402046v1 13 Feb 2004 Contents ∗ Structure Nuclear of View Unified as Model Shell The † ‡ § ¶ lcrncaddress: Electronic lcrncaddress: Electronic lcrncaddress: Electronic lcrncaddress: Electronic lcrncaddress: Electronic I.Teslto ftesclrpolmi finite a in problem secular the of solution The III. I h interaction The II. .Introduction I. .Peetpsiiiis20 possibilities Present G. space .The D. .Uieslt fteraitcitrcin 15 interactions realistic the of Universality D. .Aotti eiw h nfidve 6 view unified the review: this About D. .TeLnzsmto 16 method Lanczos The A. .Tetreplaso h hl oe 2 model shell the of pillars three The A. .Eetv neatos8 interactions Effective A. .TeGlasgow The C. .Tevlnespace valence The C. .TemlioeHmloin13 Hamiltonian multipole The C. .Tecoc ftebss17 basis the of choice The B. .Tecmeigveso ula tutr 3 structure nuclear of views competing The B. .Temnpl aitna 9 Hamiltonian monopole The B. .Teculdcd AHN 19 NATHAN. code coupled The E. .N oeselmdl20 model shell core No F. lhuhw r ul wr htmc fteceiiiyo th of credibility the of much that aware fully are spectrosc we specificitie “merely” although have the to and appear forces, calculations—that other nuclear conserving non isospin rcletv aue ewl loteti oedti h pr the detail some transiti Teller in Gamow the treat of also quenching of will origin We nu the the motion, interaction of nature. manifestations effective collective wh the the or results all of simultaneously those universality describe on to concentrate the will n approach: we a solve the review to present appears the s combination and In This saturation forces. bad the two-body correct istic to necessary terms monopole osrce htcnanbt w n he-oycontribut three-body ( and two potentials both realistic contain been that effectiv have new constructed methods Furthermore, approximation limitations. dimensionality new the many Besides, analyze. ned ti o osbet ignlz arcsi determ in matrices diagonalize to understa possible 10 our now to in is gains qualita it remarkable and in Indeed, quantitative resulted both have which witnessed ies, has decade last The 2018) 23, October (Dated: 4 3 2 1 .Caurier, E. eatmnod F´ısica Aut´onoma, Te´orica, de Universidad Departamento nttc´ aaaad eec sui Avan¸cats, Llu´ı Estudis i Recerca de Instituci´o Catalana ntttdEtdsEpcasd aauy,Eic eu,G Nexus, Edifici Catalunya, de Espacials d’Estudis Institut ntttd ehrhsSbtmqe,I23CR,Univers IN2P3-CNRS, Subatomiques, Recherches de Institut .Realistic 3. field Mean 2. Collective 1. .Clas vie 14 11 8 10 avoided Collapse 1. formation Shell forms Factorable 2. properties. Bulk 1. calculations and Theory 1. 9 m sn h aco rdaoa osrcin hs omla formal whose construction, tridiagonal Lanczos the using shm oeATIE18 ANTOINE code -scheme [email protected] [email protected] [email protected] [email protected] [email protected] 1, vs ∗ m vs. vs .Mart´ınez-Pinedo,G. hnmnlgcl4 Phenomenological shm oe18 code -scheme ignlztos4 Diagonalizations irsoi 3 Microscopic i.e. ossetwt w ulo aa.Telte incorporate latter The data). two with consistent , ,3, 2, † .Nowacki, F. 16 5 2 7 ulo-ulo neatoshv been have interactions nucleon-nucleon e ftevr eto ihnce.Many nuclei. rich very the of s opns2,E000Breoa Spain Barcelona, E-08010 23, Companys s pcitrs—r oce pnbriefly, upon touched interest—are opic la yaisete fsnl particle single of either dynamics clear 1, n,tedouble the ons, n h aaiyo h hl Model Shell the of capacity the and dn ftesrcueo h nucleus. the of structure the of nding me fhtet uzigproblems. puzzling hitherto of umber ‡ atbac,209 ard Spain Madrid, 28049, Cantoblanco, elfrainpoete ftereal- the of properties hell-formation ieporse nSelMdlstud- Model Shell in progresses tive nna pcso iesoaiyup dimensionality of spaces inantal II te ein n themes and regions Other VIII. os h omraedrvdfrom derived are former The ions. a ai` ,E004Breoa Spain Barcelona, E-08034 Capit`a 2, ran hl oe et nthem. on rests Model Shell e c lutaetegoa etrsof features global the illustrate ich I.Dsrpino eynurnrc nuclei rich neutron very of Description VII. .Poves, A. besascae ihrotational with associated oblems I peia hl oe ecito fnuclear of description model shell Spherical VI. V h aco basis Lanczos The IV. teLusPser -73 tabug France Strasbourg, F-67037 Pasteur, it´e Louis .Te0 The V. eeoe nodrt overcome to order in developed dnmrclapcsw will we aspects numerical nd rotations .Selmdlcluain nhairnce 52 nuclei heavier in calculations model Shell D. .N4;fo “magic” from N=40; D. .The D. .Temnpl rbe n h three-body the and problem monopole The A. .Rtr nthe in Rotors A. .Atohsclapiain 47 applications Astrophysical A. .Lvldensities Level A. .N=8; A. .Rdiiooesit nteClimiooe 51 the in shifts Radii C. .GmwTle n antcdpl tegh 32 strength. magnetic and Teller Gamow C. .LnzsSrnt ucin 25 Functions Strength Lanczos C. .N2:Vleaiiy44 Vulnerability N=28: C. .Hairnce:qaiped U3 38 SU(3) quasi+pseudo nuclei: Heavier C. .The B. .Quasi-SU(3). B. .Tegon state ground The B. B. .N=20; B. .Rno aitnas53 Hamiltonians Random E. .Rttoa ad funtrlprt 40 unnatural of bands Rotational E. .Temaigo h aec pc 32 space valence Quenching the of 2. meaning The 1. .Pr pcrsoy30 29 30 spectroscopy Pure forces conserving 3. non Isospin 2. factors Spectroscopic 1. interaction .MneCromtos24 23 24 methods Carlo Monte methods approximation 3. numerical Other method 2. exp(S) The 1. ββ 4, ~ -decays β § ~ ~ dcy,teeetof effect the -decays, ω pf 36 n .P Zuker P. A. and 11 rand Ar calculations 32 shell i halos Li: g eomditues43 intruders deformed Mg, pf 40 h pure the asprdfre ad 39 bands super-deformed Ca hl 35 shell 68 1, it deformed to Ni ¶ 64 r45 Cr 26 34 33 29 26 47 36 22 21 21 42 42 50 2

IX. Conclusion 53 As the number of and depart from the magic numbers it becomes indispensable to include in Acknowledgments 54 some way the “residual” two-body interaction, to break A. Basic definitions and results 54 the degeneracies inherent to the filling of orbits with two or more . At this point difficulties accu- B. Full form of Hm 55 mulate: “Jensen himself never lost his skeptical attitude 1. Separation of Hmnp and Hm0 56 towards the extension of the single-particle model to in- 2. Diagonal forms of Hm 57 clude the dynamics of several nucleons outside closed C. The center of mass problem 57 shells in terms of a residual interaction” (Weidenm¨uller, 1990). Nonetheless, some physicists chose to persist. One References 58 of our purposes is to explain and illustrate why it was worth persisting. The keypoint is that passage from the IPM to the interacting shell model, the shell model (SM) I. INTRODUCTION for short, is conceptually simple but difficult in practice. In the next section we succinctly review the steps in- In the early days of , the nucleus, com- volved. Our aim is to give the reader an overall view of posed of strongly interacting neutrons and protons con- the Shell Model as a sub-discipline of the Many Body fined in a very small volume, didn’t appear as a system to problem. “Inverted commas” are used for terms of the which the shell model, so successful in the , could nuclear jargon when they appear for the first time. The be of much relevance. Other descriptions—based on the rest of the introduction will sketch the competing views analogy with a charged liquid drop—seemed more natu- of and establish their connections with ral. However, experimental evidence of independent par- the shell model. Throughout, the reader will be directed ticle behavior in nuclei soon began to accumulate, such to the sections of this review where specific topics are as the extra binding related to some precise values of the discussed. number of neutrons and protons (magic numbers) and the systematics of spins and parities. The existence of shell structure in nuclei had already A. The three pillars of the shell model been noticed in the thirties but it took more than a decade and numerous papers (see Elliott and Lane, 1957, The strict validity of the IPM may be limited to closed for the early history) before the correct prescription was shells (and single particle –or hole– states built on them), found by Mayer (1949) and Axel, Jensen, and Suess but it provides a framework with two important compo- (1949). To explain the regularities of the nuclear proper- nents that help in dealing with more complex situations. ties associated to “magic numbers”—i.e., specific values One is of mathematical nature: the oscillator orbits de- of the number of protons Z and neutrons N—the authors fine a basis of Slater determinants, the m-scheme, in proposed a model of independent nucleons confined by a which to formulate the Schr¨odinger problem in the oc- surface corrected, isotropic harmonic oscillator, plus a cupation number representation (Fock space). It is im- strong attractive -orbit term1, portant to realize that the oscillator basis is relevant, not so much because it provides an approximation to the 1 individual nucleon wave-functions, but because it pro- U(r)= ~ω r2 + D ~l 2 + C ~l ~s. (1) 2 · vides the natural quantization condition for self-bound systems. In modern language this proposal amounts to assume Then, the many-body problem becomes one of diago- that the main effect of the two-body nucleon-nucleon nalizing a simple matrix. We have a set of determinantal interactions is to generate a spherical mean field. The † † states for A particles, ai1 ...aiA 0 = φI , and a Hamil- of the of a given nucleus is tonian containing kinetic and| potentiali | i energies then the product of one Slater determinant for the pro- K V tons and another for the neutrons, obtained by filling = a†a a†a†a a (2) H Kij i j − Vijkl i j k l the lowest subshells (or “orbits”) of the potential. This ij i≤j k≤l primordial shell model is nowadays called “independent X X particle model” (IPM) or, “naive” shell model. Its foun- that adds one or two particles in orbits i, j and removes one or two from orbits k, l, subject to the Pauli prin- dation was pioneered by Brueckner (1954) who showed † how the short range nucleon-nucleon repulsion combined ciple ( ai aj = δij ). The eigensolutions of the prob- lem Φ{ = } c φ , are the result of diagonalizing with the Pauli principle could to nearly independent | αi I I,α| I i particle motion. the matrix φI φI′ whose off-diagonal elements are hP |H| i either 0 or ijkl. However, the dimensionalities of the matrices—though±V not infinite, since a cutoff is inherent to a non-relativistic approach—are so large as to make 1 We assume throughout adimensional oscillator coordinates, i.e., the problem intractable, except for the lightest nuclei. r −→ (mω/~)1/2 r. Stated in these terms, the nuclear many body problem 3 does not differ much from other many fermion problems B. The competing views of nuclear structure in condensed matter, quantum liquids or cluster physics; with which it shares quite often concepts and techniques. Because Shell Model work is so computer-intensive, it The differences come from the interactions, which in the is instructive to compare its history and recent devel- nuclear case are particularly complicated, but paradox- opments with the competing—or alternative—views of ically quite weak, in the sense that they produce suf- nuclear structure that demand less (or no) computing ficiently little mixing of basic states so as to make the power. zeroth order approximations bear already a significant resemblance with reality.

This is the second far-reaching –physical– component 1. Collective vs. Microscopic of the IPM: the basis can be taken to be small enough as to be, often, tractable. Here some elementary def- The early SM was hard to reconcile with the idea of initions are needed: The (neutron and ) ma- the compound nucleus and the success of the liquid drop jor oscillator shells of principal p = model. With the discovery of rotational motion (Bohr, 0, 1, 2, 3 ..., called s, p, sd, pf... respectively, of energy 1952; Bohr and Mottelson, 1953) which was at first as ~ ω(p +3/2), contain orbits of total surprising as the IPM, the reconciliation became ap- j = 1/2, 3/2 ...p +1/2, each with its possible jz pro- parently harder. It came with the realization that jections, for a total degeneracy Dj = 2j + 1 for each collective rotors are associated with “intrinsic states” subshell, and Dp = (p + 1)(p + 2) for the major shells very well approximated by deformed mean field determi- (One should not confuse the p = 1 shell —p-shell in the nants (Nilsson, 1955), from which the exact eigenstates old spectroscopic notation—with the generic harmonic can be extracted by projection to good angular momen- ~ oscillator shell of energy ω(p +3/2)). tum (Peierls and Yoccoz, 1957); an early and spectacular When for a given nucleus the particles are restricted example of spontaneous symmetry breaking. By further to have the lowest possible values of p compatible with introducing nuclear superfluidity (Bohr et al., 1958), the the Pauli principle, we speak of a 0~ω space. When the unified model was born, a basic paradigm that remains many-body states are allowed to have components in- valid. volving basis states with up to N oscillator quanta more Compared with the impressive architecture of the uni- than those pertaining to the 0~ω space we speak of a fied model, what the SM could offer were some striking N~ω space (often refered as “no core”) but isolated examples which pointed to the soundness of a many-particle description. Among them, let us men- For nuclei up to A 60, the major oscillator shells pro- tion the elegant work of Talmi and Unna (1960), the ≈ ~ vide physically meaningful 0 ω spaces. The simplest pos- f n model of McCullen, Bayman, and Zamick (1964)— sible example will suggest why this is so. Start from the 7/2 probably the first successful diagonalization involving first of the magic numbers N = Z = 2; 4He is no doubt a both neutrons and protons—and the Cohen and Kurath closed shell. Adding one particle should produce a pair of (1965) fit to the p shell, the first of the classical 0~ω re- single-particle levels: p and p . Indeed, this is what 3/2 1/2 gions. It is worth noting that this calculation involved is found in 5He and 5Li. What about 6Li? The lowest 2 spaces of m-scheme dimensionalities, dm, of the order of configuration, p3/2, should produce four states of angular 100, while at fixed total angular momentum and isospin momentum and isospin JT = 01, 10, 21, 30. They are ex- d 10. JT ≈ perimentaly observed. There is also a JT = 20 level that The microscopic origin of rotational motion was found requires the p3/2 p1/2 configuration. Hence the idea of m by Elliott (1958a,b). The interest of this contribution choosing as basis p (p stands generically for p3/2 p1/2) was immediately recognized but it took quite a few years 16 for nuclei up to the next closure N = Z = 8, i.e., O. to realize that Elliott’s quadrupole force and the under- Obviously, the general Hamiltonian in (2) must be trans- lying SU(3) symmetry were the foundation, rather than formed into an “effective” one adapted to the restricted an example, of rotational motion (see Section VI). basis (the “valence space”). When this is done, the re- It is fair to say that for almost twenty years after its sults are very satisfactory (Cohen and Kurath, 1965). inception, in the mind of many physicists, the shell model The argument extends to the sd and pf shells. still suffered from an implicit separation of roles, which By now, we have identified two of the three “pillars” assigned it the task of accurately describing a few spe- of the Shell Model: a good valence space and an effec- cially important nuclei, while the overall coverage of the tive interaction adapted to it. The third is a shell model nuclear chart was understood to be the domain of the code capable of coping with the secular problem. In Sec- unified model. tion I.C we shall examine the reasons for the success of A somehow intermediate path was opened by the well the classical 0~ω SM spaces and propose extensions ca- known of Arima and Iachello pable of dealing with more general cases. The interaction (1975) and its developments, that we do not touch here. will touched upon in Section I.B.3 and discussed at length The interested reader can find the details in (Iachello and in Section II. The codes will be the subject of Section III. Arima, 1987). 4

2. Mean field vs Diagonalizations 3. Realistic vs Phenomenological

The first realistic2 matrix elements of Kuo and Brown Nowadays, many regions remain outside the direct (1966) and the first modern shell model code by French, reach of the Shell Model, but enough has happened in Halbert, McGrory, and Wong (1969) came almost simul- the last decade to transform it into a unified view. Many taneously, and opened the way for the first generation steps—outlined in the next Section I.D—are needed to of “large scale” calculations, which at the time meant substantiate this claim. Here we introduce the first: A dJT 100-600. They made it possible to describe the unified view requires a unique interaction. Its free pa- neighborhood≈ of 16O (Zuker, Buck, and McGrory, 1968) rameters must be few and well defined, so as to make and the lower part of the sd shell (Halbert et al., 1971). the calculations independent of the quantities they are However, the increases in tractable dimensionalities were meant to explain. Here, because we touch upon a dif- insufficient to promote the shell model to the status of ferent competing view, of special interest for shell model a general description, and the role-separation mentioned experts, we have to recall the remarks at the end of the above persisted. Moreover the work done exhibited seri- first paragraph of the preceding Section I.B.2: ous problems which could be traced to the realistic ma- The exciting prospect that realistic matrix elements trix elements themselves (see next Section I.B.3). could lead to parameter-free spectroscopy did not mate- However, a fundamental idea emerged at the time: rialize. As the growing sophistication of numerical meth- the existence of an underlying universal two-body in- ods allowed to treat an increasing number of particles teraction, which could allow a replacement of the uni- in the sd shell, the results became disastrous. Then, two fied model description by a fully microscopic one, based schools of thought on the status of phenomenological cor- on mean-field theory. The first breakthrough came rections emerged: In one of them, all matrix elements when Baranger and Kumar (1968) proposed a new form were considered to be free parameters. In the other, of the unified model by showing that Elliott’s quadrupole only average matrix elements (“centroids”)—related to force could be somehow derived from the Kuo-Brown ma- the bad saturation and shell formation properties of the trix elements. Adding a pairing interaction and a spher- two-body potentials—needed to be fitted (the monopole ical mean field they proceeded to perform Hartree Fock way). The former lead to the “Universal sd” (USD) in- Bogoliubov calculations in the first of a successful se- teraction (Wildenthal, 1984), that enjoyed an immense ries of papers (Kumar and Baranger, 1968). Their work success and for ten years set the standard for shell model could be described as shell model by other means, as it calculations in a large valence space (Brown and Wilden- was restricted to valence spaces of two contiguous major thal, 1988), (see Brown, 2001, for a recent review). The shells. second, which we adopt here, was initiated in Eduardo Pasquini PhD thesis (1976), where the first calculations This limitation was overcome when Vautherin and in the full pf shell, involving both neutrons and pro- Brink (1972) and Decharg´e and Gogny (1980) initi- tons were done3. Twenty years later, we know that the ated the two families of Hartree Fock (HF) calculations minimal monopole corrections proposed in his work are (Skyrme and Gogny for short) that remain to this day sufficient to provide results of a quality comparable to the only tools capable of giving microscopic descriptions those of USD for nuclei in f7/2 region. throughout the . They were later joined by However, there were still problems with the monopole the relativistic HF approach (Serot and Walecka, 1986). 56 (For a review of the three variants see Bender, Hee- way: they showed around and beyond Ni, as well as in nen, and Reinhard, 2003b). See also P´eru, Girod, and the impossibility to be competitive with USD in the sd Berger (2000) and Rodriguez-Guzm´an, Egido, and Rob- shell (note that USD contains non-monopole two-body ledo (2000) and references therein for recent work with corrections to the realistic matrix elements). The solu- the Gogny force, the one for which closest contact with tion came about only recently with the introduction of realistic interactions and the shell model can be estab- three-body forces: This far reaching development will lished (Sections VI.A and II.D). be explained in detail in Sections II, II.B and V.A. We will only anticipate on the conclusions of these sections Since single determinantal states can hardly be ex- through two syllogisms. The first is: The case for real- pected accurately to describe many body solutions, ev- istic two-body interactions is so strong that we have to erybody admits the need of going beyond the mean field. accept them as they are. Little is known about three- Nevertheless, as it provides such a good approximation body forces except that they exist. Therefore problems to the wavefunctions (typically about 50%) it suggests with calculations involving only two body forces must be efficient truncation schemes, as will be explained in Sec- blamed on the absence of three body forces. This argu- tions IV.B.1 and IV.B.2.

3 Two months after his PhD, Pasquini and his wife “disappeared” 2 Realistic interactions are those consistent with data obtained in in Argentina. This dramatic event explains why his work is only two (and nowadays three) nucleon systems. publicly available in condensed form (Pasquini and Zuker, 1978). 5 ment raises (at least) one question: What to make of all others so as to generate wavefunctions that can evolve to calculations, using only two-body forces, that give satis- the exact ones through low order perturbation theory. factory results? The answer lies in the second syllogism: The ultimate ambition of Shell Model theory is to get All clearly identifiable problems are of monopole origin. exact solutions. The ones provided by the sole valence They can be solved reasonably well by phenomenologi- space are, so to speak, the tip of the iceberg in terms cal changes of the monopole matrix elements. Therefore, of number of basic states involved. The rest may be so fitted interactions can differ from the realistic ones basi- well hidden as to make us believe in the literal valid- cally through monopole matrix elements. (We speak of ity of the shell model description in a restricted valence R-compatibility in this case). The two syllogisms become space. For instance, the magic closed shells are good va- fully consistent by noting that the inclusion of three-body lence spaces consisting of a single state. This does not monopole terms always improves the performance of the mean that a magic nucleus is 100% closed shell: 50 or forces that adopt the monopole way. 60% should be enough, as we shall see in Section IV.B.1. For the 0~ω spaces, the Cohen Kurath (Cohen and In section V.C.1 it will be argued that the 0~ω valence Kurath, 1965), the Chung Wildenthal (Wildenthal and spaces account for basically the same percentage of the Chung, 1979) and the FPD6 (Richter et al., 1991) inter- full wavefunctions. Conceptually the valence space may actions turn out to be R-compatible. The USD (Wilden- be thought as defining a representation intermediate be- thal, 1984) interaction is R-incompatible. It will be of tween the Schr¨odinger one (the operators are fixed and interest to analyze the recent set of pf-shell matrix ele- the wavefunction contains all the information) and the ments (GXPF1), proposed by Honma et al. (2002), but Heisenberg one (the reverse is true). not yet released. At best, 0~ω spaces can describe only a limited number of low-lying states of the same (“natural”) parity. Two contiguous major shells can most certainly cope with all C. The valence space levels of interest but they lead to intractably large spaces and suffer from a “Center of Mass problem” (analyzed in Appendix C). Here, a physically sound pruning of the The choice of the valence space should reflect a basic physical fact: that the most significant components of space is suggested by the IPM. What made the success of the model is the explanation of the observed magic the low lying states of nuclei can be accounted for by many-body states involving the excitation of particles in numbers generated by the spin-orbit term, as opposed to the HO (harmonic oscillator) ones (for shell forma- a few orbitals around the Fermi level. The history of the Shell Model is that of the interplay between experi- tion see Section II.B.2). If we separate a major shell as HO(p)= p r , where p is the largest subshell ment and theory to establish the validity of this concept. > ⊕ p > Our present understanding can be roughly summed up by having j = p +1/2, and rp is the “rest” of the HO p- saying that “few” mean essentially one or two contiguous shell, we can define the EI (extruder-intruder) spaces as major shells4. EI(p)= rp (p + 1)>: The (p + 1)> orbit is expelled from HO(p⊕+ 1) by the spin-orbit interaction and in- For a single major shell, the classical 0~ω spaces, exact trudes into HO(p). The EI spaces are well established solutions are now available from which instructive con- standards when only one “fluid” (proton or neutron) is clusions can be drawn. Consider, for instance,the spec- active: The Z or N = 28, 50, 82 and 126 isotopes or trum of 41Ca from which we want to extract the single- . These nuclei are “spherical”, amenable to ex- particle states of the pf shell. Remember that in 5Li we act diagonalizations and fairly well understood (see for expected to find two states, and indeed found two. Now, example Abzouzi et al., 1991). in principle, we expect four states. They are certainly in the data. However, they must be retrieved from a jun- As soon as both fluids are active, “deformation” ef- gle of some 140 levels seen below 7 MeV. Moreover, the fects become appreciable, leading to “coexistence” be- tween spherical and deformed states, and eventually to f7/2 ground state is the only one that is a pure single particle state, the other ones are split, even severely in dominance of the latter. To cope with this situation the case of the highest (f ) (Endt and van der Leun, we propose “Extended EI spaces” defined as EEI(p)= 5/2 r ∆ , where ∆ = p (p 2) ..., i.e., the ∆j =2 1990). (Section V.B.1 contains an interesting example of p ⊕ p+1 p >⊕ >− ⊕ the splitting mechanism). Thus, how can we expect the sequence of orbits that contain p>, which are needed to pf shell to be a good valence space? For few particles account for rotational motion as explained in Section VI. above 40Ca it is certainly not. However, as we report The EEI(1) (p1/2, d5/2, s1/2) space was successful in de- in this work, it turns out that above A = 46, the lowest scribing the full low lying spectra for A = 15-18 (Zuker m 5 et al., 1968, 1969). It is only recently that the EEI(2) (pf) configurations are sufficiently detached from all 40 (s1/2, d3/2, f7/2, p3/2) space (region around Ca) has become tractable (see Sections VI.D and VIII.C). EEI(3) is the natural space for the proton rich region centered 80 4 Three major shells are necessary to deal with super-deformation. in Zr. For heavier nuclei exact diagonalizations are not 5 A configuration is a set of states having fixed number of particles possible but the EEI spaces provide simple (and excel- in each orbit. lent) estimates of quadrupole moments at the beginning 6 of the well deformed regions (Section VI.C). the classical 0~ω calculations, some selected examples of As we have presented it, the choice of valence space pf shell spectroscopy are presented. Special attention is primarily a matter of physics. In practice, when is given to Gamow Teller transitions, one of the main exact diagonalizations are impossible, truncations are achievements of modern SM work. introduced. They may be based on systematic ap- Section VI. Another major recent achievement is the proaches, such as approximation schemes discussed in shell model description of rotational nuclei. The new Sections IV.B.1 and IV.B.2 or mean field methods men- generation of gamma detectors, Euroball and Gamma- tioned in Section I.B.2, and analyzed in Section VI.A. sphere has made it possible to access high spin states in medium mass nuclei for which full 0~ω calculations are available. Their remarkable harvest includes a large D. About this review: the unified view spectrum of collective manifestations which the spheri- cal shell model can predict or explain as, for instance, We expect this review to highlight several unifying as- deformed rotors, backbending, band terminations, yrast pects common to the most recent successful shell model traps, etc. Configurations involving two major oscillator calculations: shells are also shown to account well for the appearance a) An effective interaction connected with both the two of superdeformed excited bands. and three nucleon bare forces. Section VII. A conjunction of factors make light and b) The explanation of the global properties of nuclei medium-light nuclei near the neutron drip line specially via the monopole Hamiltonian. interesting: (a) They have recently come under intense c) The universality of the multipole Hamiltonian. experimental scrutiny. (b) They are amenable to shell d) The description of the collective behavior in the model calculations, sometimes even exact no-core ones laboratory frame, by means of the spherical shell model. (c) They exhibit very interesting behavior, such as ha- e) The description of resonances using the Lanczos los and sudden onset of deformation. (d) They achieve strength function method. the highest N/Z ratios attained. (e) When all (or most Let us see now how this review is organized along these of) the valence particles are neutrons, the spherical shell lines. References are only given for work that will not be model closures, dictated by the isovector channel of the mentioned later. nuclear interaction alone, may differ from those at the Section II. The basic tool in analyzing the interac- stability valley. These regions propose some exacting tion is the monopole-multipole separation. The for- tests for the theoretical descriptions, that the conven- mer is in charge of saturation and shell properties; it tional shell model calculations have passed satisfactorily. can be thought as the correct generalization of Eq. (1). For nearly unbound nuclei, the SM description has to Monopole theory is scattered in many references. Only be supplemented by some refined extensions such as the by the inclusion of three-body forces could a satisfac- shell model in the continuum, which falls outside the tory formulation be achieved. As this is a very recent scope of this review. References to the subject can be and fundamental development which makes possible a found in Bennaceur et al. (1999, 2000), and in Id Betan unified viewpoint of the monopole field concept, this sec- et al. (2002); Michel et al. (2002, 2003) that deal with tion is largely devoted to it. The “residual” multipole the Gamow shell model. force has been extensively described in a single reference Section VIII. There is a characteristic of the Shell which will be reviewed briefly and updated. The aim of Model we have not yet stressed: it is the approach to the section is to show how the realistic interactions can nuclear structure that can give more precise quantita- be characterized by a small number of parameters. tive information. SM wave functions are, in particular, Section III. The ANTOINE and NATHAN codes have of great use in other disciplines. For example: Weak made it possible to evolve from dimensionalities dm decay rates are crucial for the understanding of several 5 104 in 1989 to d 109, d 107 nowadays.∼ × m ∼ J ∼ astrophysical processes, and neutrinoless ββ decay is one Roughly half of the–order of magnitude–gains are due of the main sources of information about the neutrino to increases in computing power. The rest comes from masses. In both cases, shell model calculations play a algorithmic advances in the construction of tridiagonal central role. The last section deal with these subjects, Lanczos matrices that will be described in this section. and some others... Section IV. The Lanczos construction can be used to Appendix B contains a full derivation of the general eliminate the “black box” aspect of the diagonalizations, form of the monopole field. to a large extent. We show how it can be related to the notions of partition function, evolution operator and Two recent reviews by Brown (2001) and Otsuka et al. level densities. Furthermore, it can be turned into a pow- (2001b) have made it possible to simplify our task and erful truncation method by combining it to coupled clus- avoid redundancies. However we did not feel dispensed ter theory. Finally, it describes strength functions with from quoting and commenting in some detail important maximal efficiency. work that bears directly on the subjects we treat. Section V. After describing the three body mechanism Sections II, IV and the Appendices are based on un- which solves the monopole problem that had plagued published notes by APZ. 7

II. THE INTERACTION (2002) and Caurier et al. (2002), confirmed by Pieper, Varga, and Wiringa (2002), who also show that the prob- The following remarks from Abzouzi, Caurier, and lems can be remedied to a large extent by introducing Zuker (1991) still provide a good introduction to the sub- the new Illinois 3b potentials developed by Pieper et al. ject: (2001). “The use of realistic potentials (i.e., consistent with NN scattering data) in shell-model calculation was pi- oneered by Kuo and Brown (1966). Of the enormous body of work that followed we would like to extract two It is seen that the trouble detected with a 2b-only observations. The first is that whatever the forces (hard description—with binding energies, spin-orbit splittings or soft core, ancient or new) and the method of regu- and spectra—is always related to centroids, which, once larization (Brueckner G matrix (Kahana et al., 1969a; associated to operators that depend only on the num- Kuo and Brown, 1966), Sussex direct extraction (Elliott ber of particles in subshells, determine a “monopole” et al. et al. Hamiltonian m that is basically in charge of Hartree , 1968) or Jastrow correlations (Fiase , 1988)) H extraordinarily simi- Fock selfconsistency. As we shall show, the full can be the effective matrix elements are H lar et al. separated rigorously as = m + M . The multipole (Pasquini and Zuker, 1978; Rutsgi , 1971). The H H H et al. part M includes pairing, quadrupole and other forces most recent results (Jiang , 1989) amount to a vin- H dication of the work of Kuo and Brown. We take this responsible for collective behavior, and—as checked by similarity to be the great strength of the realistic inter- many calculations—is well given by the 2b potentials. actions, since it confers on them a model-independent status as direct links to the phase shifts. The second observation is that when used in shell- model calculations and compared with data these matrix The preceding paragraph amounts to rephrasing the elements give results that deteriorate rapidly as the num- two observations quoted at the beginning of this section ber of particle increases (Halbert et al., 1971) and (Brown with the proviso that the blame for discrepancies with and Wildenthal, 1988). It was found (Pasquini and experiment cannot be due to the—now nearly perfect— 2b potentials. Hence the necessary “corrections” to Zuker, 1978) that in the pf shell a phenomenologi- Hm cal cure, confirmed by exact diagonalizations up to must have a 3b origin. Given that we have no complaint with , the primary problem is the monopole contri- A=48 (Caurier et al., 1994), amounts to very simple mod- HM ifications of some average matrix elements (centroids) of bution to the 3b potentials. This is welcome because the KB interaction (Kuo and Brown, 1968).” a full 3b treatment would render most shell model cal- In Abzouzi et al. (1991) very good spectroscopy in the culations impossible, while the phenomenological study p and sd shells could be obtained only through more rad- of monopole behavior is quite simple. Furthermore it is quite justified because there is little ab initio knowledge ical changes in the centroids, involving substantial three- 6 body (3b) terms. In 1991 it was hard to interpret them of the 3b potentials . Therefore, whatever information as effective and there were no sufficient grounds to claim that comes from nuclear data beyond A=3 is welcome. that they were real. Nowadays, the need of true 3b forces has become ir- refutable: In the 1990 decade several two-body (2b) po- In Shell Model calculations, the interaction appears as tentials were developed—Nijmegen I and II (Stoks et al., matrix elements, that soon become far more numerous 1993), AV18 (Wiringa et al., 1995), CD-Bonn-(Machleidt than the number of parameters defining the realistic po- et al., 1996)—that fit the 4300 entries in the Ni- tentials. Our task will consist in analyzing the Hamilto- jmegen data base (Stoks et al.≈ , 1994) with χ2 1, and nian in the oscillator representation (or Fock space) so as none of them seemed capable to predict perfectly≈ the to understand its workings and simplify its form. In Sec- nucleon vector analyzing power in elastic (N, d) scatter- tion II.A we sketch the theory of effective interactions. In ing (the A puzzle). Two recent additions to the fam- y Section II.B it is explained how to construct from data a ily of high-precision 2b potentials—the charge-dependent minimal , while Section II.C will be devoted to extract “CD-Bonn” (Machleidt, 2001) and the chiral Idaho-A m from realisticH forces the most important contributions to and B (Entem and Machleidt, 2002)—have dispelled any . The basic tools are symmetry and scaling argu- hopes of solving the A puzzle with 2b-only interac- M y ments,H the clean separation of bulk and shell effects and tions (Entem et al., 2002). the reduction of to sums of factorable terms (Dufour Furthermore, quasi-exact 2b Green Function Monte and Zuker, 1996).H Carlo (GFMC) results (Pudliner et al., 1997; Wiringa et al., 2000) which provided acceptable spectra for A 8 (though they had problems with binding energies and≤ spin-orbit splittings), now encounter serious trouble in 6 Experimentally there will never be enough data to determine 10 the spectrum of B, as found through the No Core Shell them. The hope for a few parameter description comes from Model (NCSM) calculations of Navr´atil and Ormand chiral perturbation theory (Entem and Machleidt, 2002). 8

A. Effective interactions S = S1 + S2 + + Sk is a sum of k-body (kb) opera- tors. The decoupling··· condition ı α = 0 then to h |H| i The Hamiltonian is written in an oscillator basis as a set of coupled integral equations for the Si amplitudes. When the model space reduces to a single determinant, Γ + = + rstuZ ZtuΓ setting S = 0 leads to Hartree Fock theory if all other H K V rsΓ · 1 ≤ ≤ r s,X t u, Γ amplitudes are neglected. The S2 approximation con- Γ + tains both low order Brueckner theory (LOBT) and the + Z ZuvwΓ, (3) Vrstuvw rstΓ · random phase approximation (RPA) . In the presence r≤s≤t, u≤v≤w, Γ X of hard-core potentials, the priority is to screen them Γ where is the kinetic energy, rr′ the interaction matrix through LOBT, and the matrix elements contributing K + V r to the RPA are discarded. An important implementa- elements, ZrΓ ( ZrΓ) create (annihilate) pairs ( rs) or triples (r rst) of particles in orbits r, coupled≡ to tion of the theory was due to Zabolitzky, whose calcu- 4 16 40 Γ = JT . Dots≡ stand for scalar products. The basis and lations for He, O and Ca, included S3 (Bethe Fad- the matrix elements are large but never infinite7. deev) and S4 (Day Yacoubovsky) amplitudes (K¨ummel, The aim of an effective interaction theory is to reduce L¨uhrmann, and Zabolitzky, 1978). This “Bochum trun- the secular problem in the large space to a smaller model cation scheme” that retraces the history of nuclear mat- space by treating perturbatively the coupling between ter theory has the drawback that at each level terms that them, thereby transforming the full potential, and its re- one would like to keep are neglected. pulsive short distance behavior, into a smooth pseudopo- The way out of this problem (Heisenberg and Mi- tential. haila, 2000; Mihaila and Heisenberg, 2000) consists on In what follows, if we have to distinguish between a new truncation scheme in which some approximations large (N~ω) and model spaces we use , and for the are made, but no terms are neglected, relying on the fact pseudo-potential in the former and H,KH Kand VV for the that the matrix elements are finite. The calculations of 16 effective interaction in the latter. these authors for O (up to S3) can be ranked with the quasi-exact GFMC and NCSM ones for lighter nuclei. In the quasi-degenerate regime (many model states), 1. Theory and calculations the equations determine an effective in- teraction in the model space. The theory is much sim- The general procedure to describe an exact eigen- plified if we enforce the decoupling condition for a single state in a restricted space was obtained independently state whose exact wavefunction is written as by Suzuki and Lee (1980) and Poves and Zuker (1981a)8. It consists in dividing the full space into model ( i ) and ref = (1+ A1 + A2 + ... )(1 + B1 + B2 + ... ) ref | i | i | i external ( α ) determinants, and introducing a transfor- = (1+ C1 + C2 + ... ) ref , C = exp S, (6) | i | i mation that respects strict orthogonality between the where ref is a model determinant. The internal am- spaces and decouples them exactly: plitudes| associatedi with the B operators are those of an eigenstate obtained by diagonalizing Heff = (1 + A) ı = i + A α α = α A i (4) H | i | i iα| i | i | i− iα| i in the model space. As it is always possible to eliminate α i X X the A1 amplitude, at the S2 level there is no coupling ı α =0, Aiα is defined through ı α =0. (5) i.e., the effective interaction is a state independent G- h | i h |H| i matrix (Zuker, 1984), which has the advantage of pro- The idea is that the model space can produce one or viding an initialization for Heff . Going to the S3 level several starting wavefunctions that can evolve to exact would be very hard, and we examine what has become eigenstates through perturbative or coupled cluster eval- standard practice. uation of the amplitudes Aiα, which can be viewed as The power of CCT is that it provides a unified frame- matrix elements of a many body operator A. In cou- work for the two things we expect from decoupling: to pled cluster theory (CCT or exp S) (Coester and K¨ummel smooth the repulsion and to incorporate long range cor- (1960), K¨ummel et al. (1978)) one sets A = exp S, where relations. The former demands jumps of say, 50~ω, the latter much less. Therefore it is convenient to treat them separately. Standard practice assumes that G-matrix el- ements can provide a smooth pseudopotential in some 7 Nowadays, the non-relativistic potentials must be thought of as sufficiently large space, and then accounts for long range derived from an effective field theory which has a cutoff of about 1 GeV (Entem and Machleidt, 2002). Therefore, truly hard cores correlations through perturbation theory. The equation are ruled out, and H should be understood to act on a sufficiently to be solved is large vector space, not over the whole Hilbert space. ijαβ Gαβkl 8 The notations in both papers are very different but the pertur- Gijkl = ijkl V , (7) V − ǫα + ǫβ ǫi ǫj + ∆ bative expansions are probably identical because the Hermitean αβ − − formulation of Suzuki (1982) is identical to the one given in Poves X and Zuker (1981a). This paper also deals extensively with the where ij and kl now stand for orbits in the model space coupled cluster formalism. while in the pair αβ at least one orbit is out of it, ǫx 9 is an unperturbed (usually kinetic) energy, and ∆ a free Since contains two and three-body (2b and 3b) compo- parameter called the “starting” energy. Hjorth-Jensen, nents,H the separation is not mathematically clean. There- Kuo, and Osnes (1995) describe in detail a sophisticated fore, we propose the following partition that amounts to having two model spaces, one large and one small. = + , (8) H Hm HM In the NCSM calculations an N~ω model space is cho- sen with N 6-10. When initiated by Zheng et al. (1993) where m, the monopole Hamiltonian, contains and the pseudo-potential≈ was a G-matrix with starting en- all quadraticH and cubic (2b and 3b) forms in theK scalar † 10 ergy. Then the ∆ dependence was eliminated either by products of fermion operators arx asy , while the mul- arcane perturbative maneuvers, or by a truly interest- tipole contains all the rest. · HM ing proposal: direct decoupling of 2b elements from a Our plan is to concentrate on the 2b part, and intro- very large space (Navr´atil and Barrett, 1996), further im- duce 3b elements as the need arises. d plemented by Navr´atil, Vary, and Barrett (2000a), and m has a diagonal part, m, written in terms of num- extended to 3b effective forces (Navr´atil et al., 2000), H H† ber and isospin operators (arx ary ). It reproduces the (Navr´atil and Ormand, 2002). It would be of inter- average energies of configurations· at fixed number of par- est to compare the resulting effective interactions to the ticles and isospin in each orbit (jt representation) or, al- Brueckner and Bethe Faddeev amplitudes obtained in a ternatively, at fixed number of particles in each orbit and full exp S approach (which are also free of arbitrary start- each fluid (neutron-proton, np, or j representation). In ing energies). jt representation the centroids A most valuable contribution of NCSM is the proof that it is possible to work with a pseudo-potential in N~ω JT (2J + 1)[1 ( )J+T δ ] T = J stst st , (9a) spaces. The method relies on exact diagonalizations. As st V − −J+T V (2J + 1)[1 ( ) δst] they soon become prohibitive, in the future, CCT may P J − − 1 1 0 1 0 become the standard approach: going to S3 for N = 50 ast = (3Pst + st), bst = st st, (9b) (as Heisenberg and Mihaila did) is hard. For N = 10 it 4 V V V −V should be much easier. are associated to the 2b quadratics in number (ms) and Another important NCSM indication is that the exci- isospin operators (Ts), tation spectra converge well before the full energy, which validates formally the 0~ω diagonalizations with rudi- 1 9 m = m (m δ ), (10a) mentary potentials and second order corrections. st 1+ δ s t − st ~ st The 0 ω results are very good for the spectra. How- 1 3 ever, to have a good pseudopotential to describe energies Tst = (Ts Tt mst δst), (10b) 1+ δ · − 4 is not enough. The transition operators also need dress- st ing. For some of them, notably E2, the dressing mecha- to define the diagonal 2b part of the monopole Hamilto- ~ nism (coupling to 2 ω quadrupole excitations) has been nian (Bansal and French (1964), French (1969)) well understood for years (see Dufour and Zuker, 1996, for a detailed analysis), and yields the, abundantly tested d = d + (a m + b T )+ d3 , (11) and confirmed, recipe of using effective charges of 1.5e Hmjt K st st st st Vm ≈ s≤t for the protons and 0.5e for neutrons. For the Gamow X Teller (GT) transitions,≈ mediated by the spin-isospin The 2b part is a standard result, easily extended to in- στ operator the renormalization mechanism involves an ± clude the 3b term overall “quenching” factor of 0.7–0.8 whose origin is far subtler. It will be examined in Sections V.C.1 and V.C.2. d3 = (a m + b T ), (12) The interested reader may wish to consult them right Vm stu stu stu stu s t u away. X where m m m , or m (m 1)(m 2)/6 and stu ≡ st u s s − s − Tstu msTtu or (ms 2)Tsu. The extraction of the full B. The monopole Hamiltonian is≡ more complicated− and it will be given in detail Hm in Appendix B. Here we only need to note that m is A many-body theory usually starts by separating the closed under unitary transformations of the underlyingH Hamiltonian into an “unperturbed” and a “residual” fermion operators, and hence, under spherical Hartree- part, = 0 + r. The traditional approach consists Fock (HF) variation. This property explains the interest in choosingH H for H a one body (1b) single-particle field. of the + separation. H0 Hm HM

9 Even old potentials fit well the low energy NN phase shifts, the 10 r and s are subshells of the same parity and angular momentum, only that matter at 0~ω. x and y stand for neutrons or protons 10

1. Bulk properties. Factorable forms In Eq. (15) a single term strongly dominates all oth- ers. For the KLS interaction (Kahana et al., 1969b) and m must contain all the information necessary to pro- including the first eight major shells, the result in Fig. 1 duceH the parameters of the Bethe Weisz¨acker mass for- is roughly approximated by mula, and we start by extracting the bulk energy. The key step involves the reduction to a sum of factorable forms valid for any interaction (Dufour and Zuker, 1996). Its enormous power derives from the strong dominance 4. 0.5 (l l ) + (j l) of a single term in all the cases considered so far. Us − − h i − , (17) ≈ Dp For clarity we restrict attention to the full isoscalar centroids defined in Eqs. (B18a) and (B18c)11 p

3δstbst st = ast . (13) where p is the principal quantum number of an oscillator V − 4(4js + 1) shell with degeneracy Dp = s(2js +1)=(p + 1)(p + 2), Diagonalize st l = l l(2l + 1)/ l(2l + 1). The (unitary) U matrices V hhavei been affected by an arbitraryP factor 6 to have num- −1 d U U = = st = UskUtk k, ∴ (14) bers ofP order unity.P Operators of the form (D =2j +1) Vm E ⇒V E s s Xk d = U m U m m . (15) Vm Ek sk s tk t − Vss s s t s Xk X X X Ω=ˆ msΩs with DsΩs = 0 (18) For the 3b interaction, the corresponding centroids stu V s s are treated as explained in (Dufour and Zuker, 1996, X X Appendix B1b): The st pairs are replaced by a sin- gle index x. Let L and M be the dimensions of the x and s arrays respectively. Construct and diagonalize an vanish at closed shells and are responsible for shell effects. (L + M) (L + M) matrix whose non-zero elements are As l l ) and j l are of this type, only the the U part the rectangular× matrices and . Disregarding the p xs sx contributes− h i to the− bulk energy. contractions, the strictly 3bV part (3)V can then be written as a sum of factors V To proceed it is necessary to know how interactions d3 = (3) U (1)m U (2) m m . (16) depend on the oscillator frequency ω of the basis, related Vm Ek sk s tu,k t u k s x≡tu to the observed square radius (and hence to the density) X X X through the estimate (Bohr and Mottelson, 1969, Eq. (2- Full factorization follows by applying Eqs. (14,15) to the 157)): (2) Utu,k matrices for each k.

3 Us ~ω 35.59 40 Up = = ~ω MeV (19) 2.5 (A)1/3 r2 ⇒ ≈ A1/3 h i 2 s

, U 1.5 p 3/2 U A δ force scales as (~ω) . A 2b potential of short range 1 is essentially linear in ~ω; for a 3b one we shall tentatively assume an (~ω)2 dependence while the Coulomb force 0.5 goes exactly as (~ω)1/2.

0 0 1 2 3 4 5 6 7 To calculate the bulk energy of we av- p erage out subshell effects through uniform filling m s ⇒ mp Ds/Dp. Though the Ω-type operators vanish, we have FIG. 1 The Us terms in Eq. (17). Arbitrary scale. kept them for reference in Eqs. (21,22) below. The lat- ter is an educated guess for the 3b contribution. The eigenvalue 0 for the dominant term in Eq. 15 is replaced by ~ω definedE so as to have U = 1. The subindex V0 p 11 They ensure the right average energies for T = 0 closed shells, m is dropped throughout. Then, using Boole’s factorial powers, e.g., p(3) = p(p 1)(p 2), we obtain the fol- which is not the case in Eq. (11) if one simply drops the bst − − terms. lowing asymptotic estimates for the leading terms (i.e., 11 disregarding contractions) nucleons only “see” the low energy part of the 2b poten- tial involving basically s wave scattering, which has been ~ω d = m (p +3/2) traditionally well fitted by realistic potentials. Which in K 2 p p turn explains why primitive versions of such potentials ~ X give results close to the modern interactions for G-matrix ω (3) ~ = (pf + 3) (pf +2) (20) elements calculated at reasonable ω values. This is a re- ⇒ 4 current theme in this review (see especially Section II.D) 2 m and Fig. 2 provides an example of particular relevance for d ~ω p + Ωˆ V ≈ V0 the study of shell effects in next Section II.B.2. Before p Dp ! X we move on, we insist on what we have learned so far: It = ~ω [p (p +p 4)]2, (21) may be possible to describe nuclear structure with soft ⇒ V0 f f 2 (2+3)b potentials consistent with the low energy data m m d3 (~ω)2β p + Ωˆ p + Ωˆ coming from the A = 2 and 3 systems. V ≈ V0 D 1 2 p p ! p Dp ! X X ~ 2 3 2 1.8 = ( ω) β 0pf (pf + 4) . p (22) ⇒ V 1.6 Finally, relate pf to A : pf (3) 1.4 2(pf ) mp = 2(p + 1)(p +2)= A = . (23) ⇒ 3 1.2 p p=0 s

X X U 1 Note that in Eq. (22) we can replace (~ω)2 by (~ω)1+κ and change the powers of Dp in the denominators accord- 0.8 ingly. d d 0.6 Assuming that non-diagonal and terms cancel Bonn K V KLS we can vary with respect to ~ω to obtain the saturation 0.4 energy 0 1 2 3 4 5 6 7 8 9 10 11 s

FIG. 2 The KLS (Kahana et al., 1969b) shell effects in Fig. 1 E = d + (1 βωκp ) d, s K − f V for the first 10 orbits compared to those produced by a Bonn ∂E d + d potential (Hjorth-Jensen, 1996) (to within an overall factor). =0 βωκp = K V ∂ω ⇒ e f (κ + 1) d Points in the same major shell are connected by lines and κ V ordered by increasing j. ∴ E = ( d + d) s κ +1 K V ~ω κ 3 4/3 = e (1 4 ) A . (24) 4 κ +1 − V0 2 2. Shell formation   The correct saturation properties are obtained by fix- Fig. 2 provides a direct reading of the expected sin- −1/3 ing 0 so that Es/A 15.5 MeV. The ~ωe 40A gle particle sequences produced by realistic interactions choiceV ensures the correct≈ density. It is worth noting≈ that above T = 0 closed shells. Unfortunately, they do not 2 c the same approach leads to VC 3e Z(Z 1)/5R (Duflo square at all with what is seen experimentally. In partic- and Zuker, 2002). ≈ − ular, the splitting between spin-orbit partners appears to It should be obvious that nuclear matter properties— be nearly constant, instead of being proportional to l. As derived from finite nuclei—could be calculated with tech- it could no longer be claimed that the 2b realistic forces niques designed to treat finite nuclei: A successful the- are “wrong”, the solution must come from the 3b oper- oretical mass table must necessarily extrapolate to the ators. The “educated guess” in Eq. (22) may prove of Bethe Weisz¨acker formula (Duflo and Zuker, 1995). It help in this respect but it has not been implemented yet. may be surprising though, that such calculations could What we propose instead is to examine an existing, strict d be conducted so easily in the oscillator basis. The merit (1+2)b, model of m that goes a long way in explaining goes to the separation and factorization properties of the shell formation. ThoughH we know that 3b ingredients are forces. necessary, they may be often mocked by lower rank ones, Clearly, Eq. (24) has no (or trivial) solution for κ = 0, in analogy with the kinetic energy, usually represented i.e., without a 3b term. Though 2b forces do saturate, as 1b, when in reality it is a 2b operator, as explained in they do it at the wrong place and at a heavy price be- Appendix C. Something similar seems to happen with cause their short-distance repulsion prevents direct HF the spin-orbit force which must have a 3b origin, while variation. The crucial question is now: Can we use real- a 1b picture is phenomenologically quite satisfactory, as istic (2+3)b potentials soft enough to do HF? Probably our study will confirm. We shall also find out that some we can. The reason is that the nucleus is quite dilute, and mechanisms have an irreducible 3b character. 12

d f m manifests itself directly in nuclear spectra through 5/2 theH doubly magic closures and the particle and hole states built on them. The members of this set, which we call p cs 1 are well represented by single determinants. Con- 1/2 figuration± mixing may alter somewhat the energies but, most often, enough experimental evidence exists to cor- p 3/2 rect for the mixing. Therefore, a minimal characteriza- zni ffi tion, ˜d , can be achieved by demanding that it repro- Hm f duce all the observed cs 1 states. The task was under- 7/2 49 ± 41 49 taken by Duflo and Zuker (1999). With a half dozen pa- Sc 41Ca Sc Ca rameters a fit to the 90 available data achieved an rmsd of some 220 keV. The differences in binding energies (gaps) 2BE(cs) BE(cs + 1) BE(cs 1), not included in the − − − fit, also came out quite well and provided a test of the 39 39 47 47 K Ca K Ca reliability of the results. The model concentrates on single particle splittings d3/2 which are of order A−1/3. The valence spaces involve a znc ffc s number of particles of the order of their degeneracy, i.e., 1/2 (p +3/2)2 (3A/2)2/3 according to Eq. (23). As a ≈ f ≈ π ν consequence, at midshell, the addition of single particle π or ν splittings generates shell effects of order A1/3. To avoid bulk contributions all expressions are written in terms of Ω-typeˆ operators [Eq. (18)] and the model is defined by d5/2

d ˜ (W 4K)+ Ω(ˆ ls)+ Ω(ˆ ll) 40 48 Hm ≡ − FIG. 3 Evolution of cs ± 1 states from Ca to Ca +Ω(ˆ ffi)+ Ω(ˆ zni)+ Ω(ˆ ffc)+ Ω(ˆ znc), (25)

Where: to produce the EI closures. Though zni and eei can be made to cure this serious defect, it is not clear how well 2 they do it, since the correct mechanism necessarily in- m [I] W 4K = p 2 m (p +3/2). (26) volves 3b terms, as will be demonstrated in Section V.A. − − p p Dp ! p Another problem that demands a three body mechanism X X is associated with the znc and ffc operators which al- p comes from d and d. Both terms lead to the same ways depress orbits of larger l [as in the hole spectra of gaps, and theK chosenV combination has the advantage of 47Ca and 47K in Fig. 3], which is not the case for light canceling exactly to orders A and A2/3. nuclei as seen from the spectra of 13C and 29Si. Only 3b [II] Ω(ˆ ls) 22 l s/A2/3 (very much the value in (Bohr operators can account for this inversion. ≈− · ˜d and Mottelson, 1969, Eq. (2-132)) and Ω(ˆ ll) 22[l(l + In spite of these caveats, m should provide a plausible ≈− H 1) p(p+3)/2]/A2/3 (in MeV), are Ω-typeˆ operators that view of shell formation, illustrated in Figure 4 for t = reproduce− the single-particle spectra above HO closed shells. 0 [III] Ω(ˆ ffi), Ω(ˆ zni), Ω(ˆ ffc) and Ω(ˆ znc)12 are Ω-typeˆ 2b operators whose effect is illustrated in Fig. 3: In going -20 40 48 from Ca to Ca the filling of the f7/2 (f for short) -40 49 49 47 neutron orbit modifies the spectra in Ca, Sc, Ca -60 47 and K. -80 As mentioned, realistic 2b forces will miss the particle 41 -100

spectrum in Ca. Once it is phenomenologically cor- E(t=0, Mev) rected by ls + ll such forces account reasonably well for -120 the zni and eei mechanism, within one major exception: -140 W-4K they fail to depress the f7/2 orbit with respect to the -160 W-4K+ll+ls Hm others. As a consequence, and quite generally, they fail -180 0 10 20 30 40 50 60 N

˜d 12 ff stands for “same fluid”, zn for “proton-neutron”, i for “intra- FIG. 4 The different contributions to Hm for N=Z nuclei shell” and c for “cross-shell” 13

-40 diction for a strongly rotational nucleus. However, there Hm0 Hm2 are no (or weak) spikes for the heavier isotopes except at -60 Hm4 90 96 Hm6 Zr and Zr, both definitely doubly magic. There are Hm8 many other interesting cases, and the general trend is the -80 following: No known closure fails to be detected. Con- versely, not all predicted closures are real. They may be -100 erased by deformation—as in the case of 80Zr—but this seldom happens, thus suggesting fairly reliable predictive E(t=0...8, Mev) -120 power for ˜d . Hm -140

-160 C. The multipole Hamiltonian 10 20 30 40 50 60 Z The multipole Hamiltonian is defined as M = - m. -80 H H H Hm10 As we are no longer interested in the full , but its re- Hm12 H -90 striction to a finite space, will be more modestly Hm14 HM Hm16 called HM , with monopole-free matrix elements given by -100 Hm18 -110 W JT = V JT δ δ V T . (27) rstu rstu − rs tu rs -120 We shall describe succinctly the main results of Du- -130 four and Zuker (1996), insisting on points that were not E(t=10...18, Mev) -140 stressed sufficiently in that paper (Section II.C.1), and adding some new information (Section II.D). -150 There are two standard ways of writing HM : -160 10 20 30 40 50 60 H = W Γ Z† Z , or (28) Z M rstu rsΓ · tuΓ ≤ ≤ r s,tXu,Γ ˜d FIG. 5 Hm for even t = N − Z = 10 . . . 18. 1/2 γ γ γ 0 HM = [γ] frtsu(SrtSsu) , (29) rstuXΓ N Z = 0 nuclei. ˜d is calculated by filling the oscillator − Hm γ γ orbits in the order dictated by ls + ll. The W 4K where frtsu = ωrtsu (1 + δrs)(1 + δtu)/4. The matrix − elements ωγ and W Γ are related through Eqs. (A9) term produces enormous HO closures (showing as spikes). rtsu prstu They are practically erased by the 1b contributions, and and (A10). 13 Replacing pairs by single indices rs x, tu y in it takes the 2b ones to generate the EI spikes. Note that ≡ ≡ the HO magicity is strongly attenuated but persists. The eq. (28) and rt a, su b in eq. (29), we proceed 1 3 ≡ ≡ Γ drift toward smaller binding energies that goes as A / is as in Eqs. (14) and (15) to bring the matrices Wxy Γ γ γ ≡ an artifact of the W 4K choice and should be ignored. Wrstu and fab frtsu, to diagonal form through unitary − ≡ Γ γ In Figs. 5 we show the situation for even t = N Z =0 transformations Uxk,uak to obtain factorable expressions to 18. Plotting along lines of constant t has the− advan- tage of detecting magicity for both fluids. In other words, H = EΓ U Γ Z† U Γ Z , (30) M k xk xΓ · yk yΓ the spikes appear when either fluid is closed, and are re- k,Γ x y X X X inforced when both are closed. The spikes are invari- ably associated either to the HO magic numbers (8, 20, 0 40) or to the EI ones (14, 28, 50), but the latter always γ γ γ γ γ 1/2 show magicity while the former only do it at and above HM = ek uakSa ubkSb [γ] , (31) a ! the double closures (Z,N)=(8,14),(20,28) and (40,50). Xk,γ X Xb For example, at Z = 20 40Ca shows as a weak closure. 42, 44, 46Ca are not closed (which agrees with experiment), which we call the E (or normal or particle-particle or 48Ca is definitely magic, and spikes persist for the heavier pp) and e (or multipole or particle-hole or ph) represen- 80 tations. Since m contains all the γ = 00 and 01 terms, isotopes. At Z = 40, Zr shows a nice spike, a bad pre- 00 H 01 for M , ωrstu = ωrstu = 0 (see Eq. (B11). There are no oneH body contractions in the e representation because 0τ they are all proportional to ωrstu. 13 Remember: EI valence spaces consist in the HO orbits shell p, The eigensolutions in eqs. (30) and (31) using the KLS except the largest (extruder), plus the largest (intruder) orbit interaction (Kahana et al., 1969b; Lee, 1969), yield the from shell p + 1. density of eigenvalues (their number in a given interval) 14

2000 5000 E-eigenvalue density e-eigenvalue density

1- 0 0- 0 0+ 1

number of states 1+ 0 3- 0 number of states - 1+ 0 1 0 2+ 0 2+ 0 + 1+ 1 2- 0 4 0 0 0 -4 -3 -2 -1 0 1 2 3 4 5 -10 -8 -6 -4 -2 0 2 4 Energy (MeV) Energy (MeV) FIG. 7 e-eigenvalue density for the KLS interaction in the FIG. 6 E-eigenvalue density for the KLS interaction in the pf+sdg major shells. Each eigenvalue has multiplicity [γ]. pf+sdg major shells ~ω = 9. Each eigenvalue has multiplicity The largest ones are shown by arrows. [Γ]. The largest ones are shown by arrows.

terms, in the E representation that is shown in Fig. 6 for a ~ ω 01 † † typical two-shell case. HP¯ = E (P p + P p+1) (P p + P p+1) (32) −~ω0 | | · It is skewed, with a tail at negative energies which is ~ ω 20 what we expect from an attractive interaction. Hq¯ = e (¯qp +¯qp+1) (¯qp +¯qp+1), (33) −~ω0 | | · The e eigenvalues are plotted in fig. 7. They are very symmetrically distributed around a narrow central group, turns out to be the usual pairing plus quadrupole Hamil- but a few of them are neatly detached. The strongest tonians, except that the operators for each major shell have γπ = 1−0, 1+1, 2+0, 3−0, 4+0. “If the corre- of principal quantum number p are affected by a nor- sponding eigenvectors are eliminated from H in” eq. (31) malization. E01 and e20 are the one shell values called and the associated H in eq. (30) is recalculated14, the E generically e in the discussion above. To be precise distribution becomes quite symmetric, as expected for a random interaction. ¯† † 1/2 1/2 Pp = Zrr01Ωr /Ωp , (34) If the diagonalizations are restricted to one major shell, r∈p negative parity peaks are absent, but for the positive par- X q¯ = S20q / , (35) ity ones the results are practically identical to those of p rs rs Np rs∈p Figs. 6 and 7, except that the energies are halved. This X point is crucial: 1 where Ω = j +1, q = r r2Y 2 s (36) r r rs 5 If up1 and up2 are the eigenvectors obtained in shells p1 r h k k i and p2, their eigenvalues are approximately equal ep1 1 2 2 5 4 ≈ and Ωp = Dp, p =Σqrs = (p +3/2) (37) ep2 = e. When diagonalizing in p1 +p2, the unnormalized 2 N ∼ 32π eigenvector turns out to be up1 + up2 with eigenvalue e. In the figures the eigenvalues for the two shell case are doubled, because they are associated with normalized 1. Collapse avoided eigenvectors. To make a long story short: The contribu- tion to HM associated to the large Γ = 01, and γ = 20 The pairing plus quadrupole (P + Q) model has a long and glorious history (Baranger and Kumar, 1968; Bes and Sorensen, 1969), and one big problem: as more shells are added to a space, the energy grows, eventually leading 14 Inverted commas are meant to call attention on an erratum to collapse. The only solution was to stay within lim- in (Dufour and Zuker, 1996). ited spaces, but then the coupling constants had to be 15

T readjusted on a case by case basis. The normalized ver- Similarly, OAB is the overlap for matrix elements with the sions of the operators presented above are affected by same T . OAA is a measure of the strength of an inter- universal coupling constants that do not change with the action. If the interaction is referred to its centroid (and ~ 2 number of shells. Knowing that ω0 = 9 MeV, they are a fortiori if it is monopole-free), OAA σA. The upper 01 ′ 20 ′ ≡ E /~ω0 = g = 0.32 and e /~ω0 = κ = 0.216 in |Eqs.| (32) and (33). | | 2 3 Introducing Amf (pf +3/2) , the total number of TABLE I The overlaps between the KLS(A) and Bonn(B) ≈ 3 interactions for the first ten oscillator shells (2308 matrix el- particles at the middle of the Fermi shell pf , the rela- tionship between g′, κ′, and their conventional counter- ements), followed by those for the pf shell (195 matrix ele- ments) for the BonnC (C) and KB (K) interactions. parts (Baranger and Kumar, 1968) is, for one shell15 0 1 ~ O = 17.56 O = 2.61 OAA = 6.49 0.32 ω 19.51 −1 AA AA 0 1 = 2/3 = G G0A , O = 11.84 O = 2.31 O = 4.78 Ωp ∼ 1/3 ≡ BB BB BB A Amf 0 1 OAB = 0.98 OAB = 0.99 OAB = 0.98 0.216~ω 1 216 χ′ χ′ = 0 A−5/3. (38) 2 ∼= 4/3 p 2 A1/3A 2 ≡ 2 0 1 N mf OCC = 3.71 OCC = 0.56 OCC = 1.41 0 1 OKK = 3.02 OKK = 0.46 OKK = 1.15 To see how collapse occurs assume m = O(Df ) = 0 1 O(A2/3) in the Fermi shell, and promote them to a higher OCK = 0.99 OCK = 0.97 OCK = 0.99 shell of degeneracy D. The corresponding pairing and quadrupole energies can be estimated as set of numbers in Table I contains what is probably the E = G 4m(D m +2)= G O(mD), (39) most important single result concerning the interactions: P −| | − −| | a 1969 realistic potential (Kahana et al. (1969b); Lee and E χ′ Q2 = χ′ O(m2D), (40) q ≈ −| | 0 −| | (1969)) and a modern Bonn one (Hjorth-Jensen, 1996) respectively, which become for the two possible scalings differ in total strength(s) but the normalized O(AB) are very close to unity. mD m EP (old) = O( )= EP (new) = O( ) (41) Now: all the modern realistic potentials agree closely A ⇒ A1/3 with one another in their predictions except for the bind- m2D m2 ing energies. For nuclear matter the differences are sub- E (old) = O( )= E (new) = O( ). (42) q A5/3 ⇒ q A1/3D stantial (Pieper et al., 2001), and for the BonnA,B,C If the m particles stay at the Fermi shell, all energies go potentials studied by Hjorth-Jensen et al. (1995), they as A1/3 as they should. If D grows both energies grow become enormous. However, the matrix elements given in the old version. For sufficiently large D the gain will in this reference for the pf shell have normalized cross become larger than the monopole loss O(mD1/2~ω) = overlaps of better than 0.998. At the moment, the overall O(D1/2A1/3). Therefore the traditional forces lead the strengths must be viewed as free parameters. system to collapse. In the new form there is no collapse: At a fundamental level the discrepancies in total strength stem for the degree of non-locality in the po- EP stays constant, Eq decreases and the monopole term provides the restoring force that guarantees that particles tentials and the treatment of the tensor force. In the old will remain predominantly in the Fermi shell. interactions, uncertainties were also due to the starting energies and the renormalization processes, which, again, As a model for M , P + Q is likely to be a reasonable first approximationH for many studies, provided it is sup- affect mainly the total strength(s), as can be gathered from the bottom part of Table I. plemented by a reasonable m, to produce the P +Q+m model. H The dominant terms of HM are central and table II collects their strengths (in MeV) for effective interac- tions in the pf-shell: Kuo and Brown (1968) (KB), the D. Universality of the realistic interactions potential fit of Richter et al. (1991) (FPD6), the Gogny force—successfully used in countless mean field studies— To compare sets of matrix elements we define the over- (Decharg´eand Gogny, 1980) and BonnC (Hjorth-Jensen laps et al., 1995). There is not much to choose between the different O = V V = V Γ V Γ (43) AB A · B rstuA rstuB forces which is a nice indication that overall nuclear rstuXΓ data (used in the FPD6 and Gogny fits) are consistent OAB with the NN data used in the realistic KB and BonnC OAB = . (44) √OAAOBB G-matrices. The splendid performance of Gogny de- serves further study. The only qualm is with the weak quadrupole strength of KB, which can be understood by looking again at the bottom part of Table I. In assess- 15 According to Dufour and Zuker (1996) the bare coupling con- ing the JT = 01 and JT = 10 pairing terms, it should stants, should be renormalized to 0.32(1+0.48) and 0.216(1+0.3) be borne in mind that their renormalization remains an 16 open question (Dufour and Zuker, 1996). very often only one) eigenstates of a given angular mo- Conclusion: Some fine tuning may be in order and 3b mentum (J) and isospin (T ) are needed. Secondly, the forces may (one day) bring some multipole news, but as matrices are very sparse. As can be seen in figure 8, the of now the problem is m, not M . number of NZME varies linearly instead of quadratically H H with the size of the matrices. For these reasons, itera- tive methods are of general use, in particular the Lanczos TABLE II Leading terms of the multipole Hamiltonian method (Lanczos, 1950). As an alternative, the Davidson method (Davidson, 1975) has the advantage of avoiding Interaction particle-particle particle-hole the storage of a large number of vectors, however, the JT=01 JT=10 λτ=20 λτ=40 λτ=11 large increase in storage capacity of modern computers has somehow minimized these advantages. KB -4.75 -4.46 -2.79 -1.39 +2.46 The Lanczos method consists in the construction of an FPD6 -5.06 -5.08 -3.11 -1.67 +3.17 orthogonal basis in which the Hamiltonian matrix (H) is GOGNY -4.07 -5.74 -3.23 -1.77 +2.46 tridiagonal. BonnC -4.20 -5.60 -3.33 -1.29 +2.70 A normalized starting vector (the pivot) 1 is chosen. The vector a = H 1 has necessarily the form| i | 1i | i a = H 1 + 2′ , with 1 2′ =0. (45) | 1i 11| i | i h | i III. THE SOLUTION OF THE SECULAR PROBLEM IN A Calculate H11 = 1 H 1 through 1 a1 = H11. Nor- ′ h | ′ |′ i1/2 h | i FINITE SPACE malize 2 = 2 / 2 2 to find H12 = 1 H 2 = 2′ 2′ 1/|2.i Iterate| i untilh | i state k has been found.h | | i The vectorh | i a = H k has necessarily| i the form Once the interaction and the valence space ready, it is | ki | i time to construct and diagonalize the many body secular ′ ak = Hk k−1 k 1 + Hk k k + (k + 1) . (46) matrix of the Hamiltonian describing the (effective) in- | i | − i | i | i ′ teraction between valence particles in the valence space. Calculate k ak = Hk k. Now (k + 1) is known. Nor- h | i | ′ i ′ 1/2 Two questions need now consideration: which basis to malize it to find Hk k+1 = (k + 1) (k + 1) . take to calculate the non-zero many-body matrix ele- The (real symmetric) matrixh is| diagonalizedi at each ments (NZME) and which method to use for the diag- iteration and the iterative process continues until all the onalization of the matrix. required eigenvalues are converged according to some cri- teria. The number of iterations depends little on the di- mension of the matrix. Besides, the computing time is A. The Lanczos method directly proportional to the number of NZME and for this reason it is nearly linear (instead of cubic as in the standard methods) in the dimension of the matrix. It de- pends on the number of iterations, which in turn depends on the number of converged states needed, and also on the choice of the starting vector. The Lanczos method can be used also as a projection method. For example in the m-scheme basis, a Lanc- zos calculation with the operator J 2 will generate states with well defined J. Taking these states as starting vec- tors, the Lanczos procedure (with H) will remain inside a fixed J subspace and therefore will improve the con- vergence properties. When only one converged state is required, it is convenient to use as pivot state the solu- tion obtained in a previous truncated calculation. For instance, starting with a random pivot the calculation of the ground state of 50Cr in the full pf space (dimension in m-scheme 14,625,540 Slater determinants) needs twice FIG. 8 m-scheme dimensions and total number of non-zero more iterations than if we start with the pivot obtained matrix elements in the pf-shell for nuclei with M = Tz = 0 as solution in a model space in which only 4 particles are allowed outside the 1f7/2 shell (dimension 1,856,720). In the standard diagonalization methods (Wilkinson, The overlap between these two 0+ states is 0.985. When 3 1965) the CPU time increases as N , N being the di- more eigenstates are needed (nc > 1), the best choice for mension of the matrix. Therefore, they cannot be used the pivot is a linear combination of the nc lower states in large scale shell model (SM) calculations. Nuclear SM in the truncated space. Even if the Lanczos method is calculations have two specific features. The first one is very efficient in the SM framework, there may be nu- that, in the vast majority of the cases, only a few (and merical problems. Mathematically the Lanczos vectors 17 are orthogonal, however numerically this is not strictly whose dimensions are much smaller. This is especially true due to the limited floating point machine precision. spectacular for the J = T = 0 states (see table III). Hence, small numerical precision errors can, after many It is often convenient to truncate the space. In the par- iterations, produce catastrophes, in particular, the states ticular case of the pf shell, calling f the largest subshell of lowest energy may reappear many times, rendering the (f7/2), and r, generically, any or all of the other subshells method inefficient. To solve this problem it is necessary of the p = 3 shell, the possible t-truncations involve the to orthogonalize each new Lanczos vector to all the pre- spaces ceding ones. The same precision defects can produce the appearance of non expected states. For example in a m- f m−m0 rm0 +f m−m0−1rm0+1+ +f m−m0−trm0+t, (48) ··· scheme calculation with a J = 4, M = 0 pivot, when many iteration are performed, it may happen that one where m0 = 0 if more than 8 neutrons (or protons) are present. For6 t = m m we have the full space (pf)m J = 0 and even one J = 2 state, lower in energy than − 0 the J = 4 states, show up abruptly. This specific prob- for A = 40+ m. lem can be solved by projecting on J each new Lanczos In the late 60’s, the Rochester group developed the vector. algorithms needed for an efficient work in the (J, T ) coupled basis and implemented them in the Oak-Ridge Rochester Multi-Shell code (French et al., 1969). The B. The choice of the basis structure of the calculation is as follows: In the first place, the states of ni particles in a given ji shell are defined: γ = (j )ni v J x (v is the seniority and x Given a valence space, the optimal choice of the basis | ii | i i i ii i i is related to the physics of the particular problem to be any extra quantum number). Next, the states of N par- solved. As we discuss later, depending on what states or ticles distributed in several shells are obtained by succes- properties we want to describe (ground state, yrast band, sive angular momentum couplings of the one-shell basic strength function,. . . ) and depending on the type of nu- states: cleus (deformed, spherical,. . . ) different choices of the Γ Γ3 k Γ2 basis are favored. There are essentially three possibilities [ γ1 γ2 ] γ3 ... γk . (49) depending on the underlying symmetries: m-scheme, J | i| i | i | i h i  coupled scheme, and JT coupled scheme. Compared to the simplicity of the m-scheme, the cal- culation of the NZME is much more complicated. It in- TABLE III Some dimensions in the pf shell volves products of 9j symbols and coefficients of fractional parentage (cfp), i.e., the single-shell reduced matrix ele- A 4 8 12 16 20 ments of operators of the form 6 8 9 9 M = Tz = 0 4000 2 × 10 1.10 × 10 1.09 × 10 2.29 × 10 6 7 7 † † λ † λ † † † λ j4 J = Tz = 0 156 41355 1.78 × 10 1.54 × 10 3.13 × 10 (aj1 aj2 ) , (aj1 aj2 ) , aj1 , ((aj1 aj2 ) aj3 ) . (50) J = T = 0 66 9741 3.32 × 105 2.58 × 106 5.05 × 106 This complexity explains why in the OXBASH code (Brown et al., 1985), the JT-coupled basis states As the m-scheme basis consists of Slater Determinants are written in the m-scheme basis to calculate the (SD) the calculation of the NZME is trivial, since they are NZME. The Oxford-Buenos Aires shell model code, equal to the decoupled two body matrix element (TBME) widely distributed and used has proven to be an invalu- up to a phase. This means that, independently of the size able tool in many calculations. Another recent code that of the matrix, the number of possible values of NZME is works in JT-coupled formalism is the Drexel University relatively limited. However, the counterpart of the sim- DUPSM (Novoselsky and Valli`eres, 1997) has had a much plicity of the m-scheme is that only J and T are good z z lesser use. quantum numbers, therefore all the possible (J, T ) states In the case of only J (without T ) coupling, a strong are contained in the basis and as a consequence the di- simplification in the calculation of the NZME can be mensions of the matrices are maximal. For a given num- achieved using the Quasi-Spin formalism (Ichimura and ber of valence neutrons, n , and protons, z the number v v Arima, 1966; Kawarada and Arima, 1964; Lawson and of different Slater determinants that can be built in the Macfarlane, 1965), as in the code NATHAN described valence space is : below. The advantages of the coupled scheme decrease also when J and T increase. As an example, in 56Ni the D D d = n p (47) ratio dim(M = J)/dim(J) is 70 for J = 0 but only 5.7 nv ! · zv ! for J = 6. The coupled scheme has another disadvantage compared to the m-scheme. It concerns the percentage + 50 where Dn and Dp are the degeneracies of the neutron and of NZME. Let us give the example of the state 4 in Ti proton valence spaces. Working at fixed M and Tz the (full pf space); the percentages of NZME are respectively et al. bases are smaller (d = M,Tz d (M,Tz)). The J or JT 14% in the JT -basis (see Novoselsky , 1997), 5% in coupled bases, split the full m-scheme matrix in boxes the J-basis and only 0.05% in m-scheme. For all these P 18

reasons and with the present computing facilities, we can the 2 subspaces will be associated only if M1 + M2 = M. conclude that the m-scheme is the most efficient choice A pictorial example is given in fig. 9. for usual SM calculations, albeit with some notable ex- + ceptions that we shall mention below. a m a n M M+1 M+2

i 1 2 3 4 5 6 7 89 10 11 12 13 14 15 . . . Jzp C. The Glasgow m-scheme code

The steady and rapid increase of computer power in α . . . J recent years has resulted in a dramatic increase in the 1 2 3 4 5 6 7 8 9 10 11 zn dimensionality of the SM calculations. A crucial point −M −M−1 −M−2

nowadays is to know what are the limits of a given com- a + a µ puter code, their origin and their evolution. As far as the λ NZME can be calculated and stored, the diagonalization FIG. 9 Schematic representation of the basis with the Lanczos method is trivial. It means that the fundamental limitation of standard shell model calcula- tions is the capacity to store the NZME. This is the origin It is clear that for each i state the allowed α states | i | i of the term “giant” matrices, that we apply to those for run, without discontinuity, between a minimum and a which it is necessary to recalculate the NZME during the maximum value. Building the basis in the total space by diagonalization process. The first breakthrough in the means of an i-loop and a nested α-loop, it is possible to treatment of giant matrices was the SM code developed construct numerically an array R(i) that points to the by the Glasgow group (Whitehead et al., 1977). Let us I state17: | i recall its basic ideas: It works in the m-scheme and each SD is represented in the computer by an integer word. I = R(i)+ α (51) Each bit of the word is associated to a given individual state nljmτ . Each bit has the value 1 or 0 depending For example, according to figure 9, the numerical values on whether| thei state is occupied or empty. A two-body of R are: R(1) = 0, R(2) = 6, R(3) = 12,...The re- † † lation (51) holds even in the case of truncated spaces, operator ai aj akal will select the words having the bits i,j,k,l in the configuration 0011, say, and change them provided we define sub-blocks labeled with M and t (the to 1100, generating a new word which has to be located truncation index defined in (48)). Before the diagonal- in the list of all the words using the bi-section method. ization, all the calculations that involve only the pro- ton or the neutron spaces separately are carried out and the results stored. For the proton-proton and neutron- neutron NZME the numerical values of R(i), R(j), Wij D. The m-scheme code ANTOINE and α, β, Wαβ, where i H j = Wij and α H β = Wαβ, are pre-calculated andh | stored.| i Therefore,h | in| thei Lanc- The shell model code ANTOINE16 (Caurier and zos procedure a simple loop on α and i generates all Nowacki, 1999) has retained many of these ideas, while the proton-proton and neutron-neutron NZME, WI,J = improving upon the Glasgow code in the several aspects. I H J . For the proton-neutron matrix elements the To start with, it takes advantage of the fact that the di- situationh | | i is slightly more complicated. Let’s assume that mension of the proton and neutron spaces is small com- the i and j SD are connected by the one-body operator † | i | i pared with the full space dimension, with the obvious ex- aqar (that in the list of all possible one-body operators ception of semi-magic nuclei. For example, the 1,963,461 appears at position p), with q = nljm and r = n′l′j′m′ SD with M = 0 in 48Cr are generated with only the 4,865 and m′ m = ∆m. Equivalently, the α and β SD SD (corresponding to all the possible M values) in 44Ca. are connected− by a one-body operator whose| i position| i is The states of the basis are written as the product of two denoted by µ. We pre-calculate the numerical values of SD, one for protons and one for neutrons: I = i, α . R(i), R(j),p and α,β,µ. Conservation of the total M We use, I, J, capital letters for states in the| fulli space,| i implies that the proton operators with ∆m must be as- i, j, small case Latin letters for states of the first subspace sociated to the neutron operators with ∆m. Thus we (protons or neutrons), α, β, small case Greek letters for could draw the equivalent to figure 9 for− the proton and states of the second subspace (neutrons or protons). The neutron one-body operators. In the same way as we did Slater determinants i and α can be classified by their M before for I = R(i)+ α, we can now define an index values, M1 and M2. The total M being fixed, the SD’s of K = Q(p)+ µ that labels the different two-body matrix

16 A version of the code can be downloaded from the URL http: 17 We use Dirac’s notation for the quantum mechanical state |ii, //sbgat194.in2p3.fr/~theory/antoine/main.html the position of this state in the basis is denoted by i. 19 elements (TBME). Then, we denote V (K) the numeri- matrix of 52Fe with 6 protons and 6 neutrons in the pf cal value of the proton-neutron TBME that connects the shell is 108. If now we consider 132Ba with also 6 protons states i, α and j, β . Once R(i), R(j),Q(p) and α,β,µ and 6 neutron-holes in a valence space of 5 shells, four | i | i are known, the non-zero elements of the matrix in the of them equivalent to the pf shell plus the 1h11/2 orbit, full space are generated with three integer additions: the dimension reaches 2 1010. Furthermore, when many × I = R(i)+ α, J = R(j)+ β, K = Q(p)+ µ. (52) protons and neutrons are active, nuclei tend to get de- formed and their description requires even larger valence The NZME of the Hamiltonian between states I and | i spaces. For instance, to treat the deformed nuclei in the J is then: 80 | i vicinity of Zr, the ”normal” valence space, 2p3/2,1f5/2, I H J = J H I = V (K). (53) 2p1/2, 1g9/2, must be complemented with, at least, the h | | i h | | i orbit 1d . This means that our span will be limited to The performance of the code is optimal when the 5/2 nuclei with few particles outside closed shells (130Xe re- two subspaces have comparable dimensions. It becomes mains an easy calculation) or to almost semi-magic, as for less efficient for asymmetric nuclei (for semi-magic nuclei instance the Tellurium isotopes. These nuclei are spheri- all the NZME must be stored) and for large truncated cal, therefore the seniority can provide a good truncation spaces. No-core calculations are typical in this respect. scheme. This explains the interest of a shell model code If we consider a N = Z nucleus like 6Li, in a valence in a coupled basis and quasi-spin formalism. space that comprises up to 14~ω configurations, there In the shell model code NATHAN (Caurier and are 50000 Slater determinants for protons and neutrons Nowacki, 1999) the fundamental idea of the code AN- with M =1/2 and t = 14. Their respective counterparts TOINE is kept, i. e. splitting the valence space in two can only have M = 1/2 and t = 0 and have dimen- parts and writing the full space basis as the product of sionality 1. This situation− is the same as in semi-magic states belonging to these two parts. Now, i and α are nuclei. states with good angular momentum. They| arei built| i with As far as two Lanczos vectors can be stored in the RAM the usual techniques of the Oak-Ridge/Rochester group memory of the computer, the calculations are straightfor- (French et al., 1969). Each subspace is now partitioned ward. Until recently this was the fundamental limitation with the labels J and J . The only difference with the of the SM code ANTOINE. It is now possible to overcome 1 2 m-scheme is that instead of having a one-to-one associa- it dividing the Lanczos vectors in segments: tion (M1 + M2 = M), for a given J1 we now have all the (k) Ψ = Ψ . (54) possible J2, Jmin J2 Jmax, with Jmin = J0 J1 f f ≤ ≤ | − | and Jmax = J0 + J1. The continuity between the first Xk The Hamiltonian matrix is also divided in blocks so that state with Jmin and the last with Jmax is maintained and the action of the Hamiltonian during a Lanczos iteration consequently the fundamental relation I = R(i)+ α still can be expressed as: holds. The generation of the proton-proton and neutron- neutron NZME proceeds exactly as in m-scheme. For the (k) (q,k) (q) Ψf = H Ψi (55) proton-neutron NZME the one-body operators in each † q space can be written as Oλ = (a a )λ. There exists X p j1 j2 The k segments correspond to specific values of M (and a strict analogy between ∆m in m-scheme and λ in the occasionally t) of the first subspace. The price to pay for coupled scheme. Hence, we can still establish a relation the increase in size is a strong reduction of performance of K = Q(p)+ ω. The NZME read now: the code. Now, I H J and J H I are not generated h | | i h | | i I H J = J H I = h h W (K), (56) simultaneously when I and J do not belong at the h | | i h | | i ij · αβ · same k segment of the| vectori and| i the amount of disk use with h = i Oλ j , h = α Oλ β , and increases. As a counterpart, it gives a natural way to par- ij h | p | i αβ h | ω | i allelize the code, each processor calculating some specific (k) i αJ Ψ . This technique allows the diagonalization of matri- f W (K) V (K) j βJ , (57) ces with dimensions in the billion range. All the nuclei of ∝ ·   the pf-shell can now be calculated without truncation.  λ λ 0  Other m-scheme codes have recently joined ANTOINE   in the run, incorporating some of its algorithmic findings where V (K) is a TBME. We need to perform –as in the as MSHELL (Mizusaki, 2000) or REDSTICK (Ormand m-scheme code– the three integer additions which gener- and Johnson, 2002). ate I, J, and K, but, in addition, there are two floating point multiplications to be done, since hij and hαβ, which in m-scheme were just phases, are now a product of cfp’s E. The coupled code NATHAN. and 9j symbols (see formula 3.10 in French et al. (1969)). Within this formalism we can introduce seniority trun- As the mass of the nucleus increases, the possibilities of cations, but the problem of the semi-magic nuclei and performing SM calculations become more and more lim- the asymmetry between the protons and neutrons spaces ited. To give an order of magnitude the dimension of the remains. To overcome this difficulty in equalizing the 20 dimension of the two subspaces we have generalized the ical parameter ∆ used to define G-matrix starting en- code as to allow that one of the two subspaces contains ergy. A truly ab initio formulation of the formalism was a mixing of proton and neutron orbits. For example, for presented by Navr´atil and Barrett (1998), where conver- heavy or Titanium isotopes, we can include gence to the exact solutions was demonstrated for the the neutron 1f7/2 orbit in the proton space. For the A = 3 system. The same was later accomplished for the N = 126 isotones we have only proton shells. We then A = 4 system (Navr´atil and Barrett, 1999), where it was take the 1i13/2 and 1h9/2 orbits as the first subspace, also shown that a three-body effective interaction can be while the 2f7/2, 2f5/2, 3p3/2 and 3p1/2 form the second introduced to improve the convergence of the method. one. The states previously defined with Jp and Jn are The capability to derive a three-body effective interac- now respectively labelled A1,J1 and A2,J2, A1 and A2 tion and apply it in either relative-coordinate (Navr´atil being the number of particles in each subspace. In hij et al., 2000) or Cartesian-coordinate formalism (Navr´atil and hαβ now appear mean values of all the operators in and Ormand, 2002) together with the ability to solve a Eq.(50) three-nucleon system with a genuine three-nucleon force The fact that in the coupled scheme the dimensions in the NCSM approach (Marsden et al., 2002) now opens are smaller and that angular momentum is explicitly en- the possibility to include a realistic three-nucleon force forced makes it possible to perform a very large number in the NCSM Hamiltonian and perform calculations us- of Lanczos iterations without storage or precision prob- ing two- and three-nucleon forces for the p-shell nuclei. lems. This can be essential for the calculation of strength Among successes of the NCSM approach was the first functions or when many converged states are needed, for published result of the of 4He with the example to describe non-yrast deformed bands. The cou- CD-Bonn NN potential (Navr´atil and Barrett, 1999), the pled code has another important advantage. It can be near-converged results for A = 6 using a non-local Hamil- perfectly parallelized without restriction on the number tonian (Navr´atil et al., 2001), the first observation of the of processors. The typical working dimension for large incorrect ground-state spin in 10B predicted by the realis- matrices in NATHAN is 107 (109 with ANTOINE). It tic two-body nucleon-nucleon interactions (Caurier et al., means that there is no problem to define as many fi- 2002). This last result, together with the already known nal vectors as processors are available. The calculation problems of under-binding, confirms the need of realistic of the NZME is shared between the different proces- three-nucleon forces, a conclusion that also stems from sors (each processor taking a piece of the Hamiltonian, the Green Function Monte Carlo calculations of Wiringa (k) H = k H ) leading to different vectors that are added and Pieper (2002) (see also Pieper and Wiringa, 2001, to obtain the full one: for a review of the results of the Urbana-Argonne collab- P oration for nuclei A 10). (k) (k) (k) ≤ Ψf = H Ψi, Ψf = Ψi . (58) Recently, a new version of the shell model code AN- Xk TOINE has been developed for the NCSM applications (Caurier et al., 2001b). This new code allows to perform calculations in significantly larger basis spaces and makes F. No core shell model it possible to reach full convergence for the A = 6 nuclei. Besides, it opens the possibility to investigate slowly con- The ab initio no-core shell model (NCSM) (Navr´atil verging intruder states and states with unnatural parity. et al. , 2000a,b) is a method to solve the nuclear many The largest bases reached with this code so far are, ac- body problem for light nuclei using realistic inter-nucleon cording to the number of HO excitations, the 14~ω space forces. The calculations are performed using a large but for 6Li and, according the matrix dimension, the 10~ω finite harmonic-oscillator (HO) basis. Due to the basis calculations for 10C. In the latter case, the m-scheme truncation, it is necessary to derive an effective interac- matrix dimension exceeds 800 millions. tion from the underlying inter-nucleon interaction that is appropriate for the basis size employed. The effective interaction contains, in general, up to A-body compo- G. Present possibilities nents even if the underlying interaction had, e.g. only two-body terms. In practice, the effective interaction is The combination of advances in computer technology derived in a sub-cluster approximation retaining just two- and in algorithms has enlarged the scope of possible SM or three-body terms. A crucial feature of the method is studies. The rotational band of 238U seems to be still very that it converges to the exact solution when the basis far from the reach of SM calculations, but predictions for size increases and/or the effective interaction clustering 218U are already available. Let us list some of the present increases. opportunities. In a first phase, their applications were limited to the use of realistic two-nucleon interactions, either G-matrix- i) No-core calculations. One of the major problems based two-body interactions (Zheng et al., 1994), or de- of the no-core SM is the convergence of the results rived by the Lee-Suzuki procedure (Suzuki and Lee, 1980) with the size of the valence space; For 6Li we can for the NCSM (Navr´atil and Barrett, 1996). This re- handle excitations until 14~ω and at least 8~ω for sulted in the elimination of the purely phenomenolog- all the p-shell nuclei. 21

ii) The pf-shell. This where the shell model has been still hears questions such as: Who is interested in di- most successful and exact diagonalizations are now agonalizing a large matrix? As an answer we propose to possible throughout the region Beyond 56Ni, as the examine Fig. 10 showing the two point-functions that de- 1f7/2 orbit becomes more and more bound, trun- fine the diagonal, Hii, and off-diagonal elements, Hi i+1 cated calculations are close to exact, for instance, in a typical Lanczos construction. The continuous lines in 60Zn (Mazzocchi et al., 2001) the wave-functions are fully converged when 6p-6h excitations are in- 4 cluded. Hi i+1(exact) Hi i+1(nib) 3 Hi i(exact) H (log) iii) The r3 g valence space. We use the notation rp i i for the set of the nlj orbits with 2(n 1) + l = p 2 − excluding the orbit with maximum total angular 1 momentum j = p +1/2. This space describes nu- clei in the region 28

v) Heavy nuclei. All the semi-magic nuclei, for in- Γ(N + 1) N stance the N = 126 isotones, can be easily studied ρ (x,N,p,S)= pxN qxN¯ d , (59) b Γ(xN + 1)Γ(¯xN + 1) S and the addition of few particles or holes remains tractable. Some long chains of Tellurium and Bis- where x is an a-dimensional energy,x ¯ = 1 x, N the muth isotopes have been recently studied (Caurier number of valence particles, p an asymmetry− parameter, et al. , 2003). p + q = 1, and S the span of the spectrum (distance between lowest and highest eigenstates). Introducing the energy scale ε, S, the centroid E , the variance σ2 and x IV. THE LANCZOS BASIS c are given by There exists a strong connection between the 2 2 E Lanczos algorithm, the partition function, Z(β) = S = Nε, Ec = Npε, σ = Np(1 p)ε , x = . (60) i exp( β ) i , and the evolution operator exp(i t). − S ih | − H | i H In the three cases, powers of the Hamiltonian determine Note that ρ (x,N,p,S) reduces to a discrete binomial, P b the properties of the systems. The partition function N if x = n/N = nε/S, with integer n. i.e. n can be written as Z(β) = E ρ(E) exp( βE) , the N, p and S are calculated using the moments of the Laplace transform of the density of states,− a quantity  P Hamiltonian , i.e., averages given by the traces of readily accessible once the Hamiltonian has been fully K , to be equatedH to the corresponding moments of diagonalized. The evolution operator addresses more H ρb(x,N,p,S), which for low K are the same as those of a specifically the problem of evolving from a starting vec- discrete binomial. The necessary definitions and equali- tor into the exact ground state. We shall discuss both ties follow. questions in turn. d−1tr( K )= K , E = 1 , = ( E )K H hH i c hH i MK h H− c i A. Level densities 2 2 K q p σ = , K = M , γ1 = 3 = − M M σK M √Npq For many, shell model is still synonymous with diag- 1 6pq γ = 3= − ; d = d (1 + p/q)N . (61) onalizations, in turn synonymous with black box. One 2 M4 − Npq 0 22

These quantities also define the logarithmic and inverse 30 binomial (nib) forms of Hii and Hi i+1 in Fig. 10. Note that the corresponding lines are almost impossible to dis- 25 tinguish from those of the exact matrix. The associated 20 level densities are found in Fig. 11. 15

1200 10 Level Density 1000 ρ(fpgd J=6) 5 Binomial ρ(bin) Bethe 60 0 Ni (exp) 800 -5 600 0 10 20 30 40 50 Energy 400 (States per MeV)

ρ FIG. 12 Binomial, Bethe and experimental level densities for 60 200 Ni.

0 0 2 4 6 8 10 12 14 16 state position. The problem also exists for the binomial18 E(MeV) and the result in the figure (Zuker, 2001) solves it phe- FIG. 11 Exact and binomial level densities for the matrix in nomenologically. The Shell Model Monte Carlo method Fig. 10. provides the only parameter-free approach to level den- sities (Dean et al., 1995; Langanke, 1998; Nakada and The mathematical status of these results is somewhat Alhassid, 1997) whose reliability is now established (Al- mixed. Mon and French (1975) proved that the total hassid et al., 1999). The problem is that the calculations density of a Hamiltonian system is to a first approxima- are hard. tion a Gaussian. Zuker (2001) extended the approxima- As the shape of the level density is well reproduced tion to a binomial, but the result remains valid only in by a binomial except in the neighborhood of the ground some neighborhood of the centroid. Furthermore, one state, to reconcile simplicity with full rigor we have to does not expect it to hold generally because binomial examine the tridiagonal matrix at the origin. thermodynamics is trivial and precludes the existence of phase transitions. In the example given above, we do not deal with the total density, which involves all states, B. The ground state ρ = JT (2J + 1)(2T + 1)ρJT but with a partial ρJT at fixed quantum numbers. In this case Zuker et al. (2001) conjecturedP that the tridiagonal elements given by the Obviously, some dependence on the pivot should exist. We examine it through the J = 1 T = 3 pf states in logarithmic and nib forms are valid, and hence describe 48 19 the full spectrum. The conjecture breaks down if a dy- Sc . The matrix is 8590-dimensional, and we calcu- namical symmetry is so strong as to define new (approx- late the ground state with two pivots, one random (ho- imately) conserved quantum numbers. mogeneous sum of all basic states), the other variational (lowest eigenstate in f 8 space). A zoom on the first ma- Granted that a single binomial cannot cover all situa- 7/2 tions we may nonetheless explore its validity in nuclear trix elements in Fig. 13 reveals that they are very differ- physics, where the observed level densities are extremely ent for the first few iterations, but soon they merge into well approximated by the classical formula of Bethe the canonical patterns discussed in the preceding section. (1936) with a shift ∆, The ground state wavefunction is unique of course, but it takes the different aspects shown in Fig. 14. In both cases the convergence is very fast, and it is not difficult to √2π e√4a(E+∆) show in general that it occurs for a number of iterations ρB(E, a, ∆) = 1/4 5/4 . (62) 12 (4a) (E + ∆) of order N log N for dimensionality d = 2N . However, the variational pivot is clearly better if we are interested Obviously, if binomials are to be useful, they must in the ground state: If its overlap with the exact solu- reproduce—for some range of energies—Eq. (62). They tion exceeds 50%, all other contributions are bound to do indeed, as shown in Fig. 12, where the experimen- tal points for 60Ni (Iljinov et al., 1992) are also given. Though Bethe and binomial forms are seen to be equiv- alent, the latter has the advantage that the necessary 18 The energies are referred to the mean (centroid) of the distribu- parameters are well defined and can be calculated, while tion, while they are measured with respect to the ground state. in the former, the precise meaning of the a parameter 19 They are reached in the 48Ca(p,n)(48)Sc reaction, which will be is elusive. The shift ∆ is necessary to adjust the ground our standard example of Gamow Teller transitions. 23

30 the full wavefunction takes the form

25 0¯ = (1+ c P + c P + + c P ) 0 , (63) | i 1 1 2 2 ··· I I | i 20 where Pm is a polynomial in H that acting on the pivot 15 produces orthogonal unnormalized vectors in the Lanczos basis: Pm 0 = m . Eq. (46) becomes 10 | i | i m = V m 1 + E m + m +1 . (64) 5 m m Tridiagonal elements H| i | − i | i | i Hii, Variational 0 Hii Random To relate with the normalized version (m m¯ ); Divide Hii+1, Variational by m m 1/2, then multiply and divide m ⇒1 and m+1 Hii+1, Random -5 h | i | − i | i 0 20 40 60 80 100 by their norms to recover Eq. (46), and obtain Em = 2 i Hm¯ m¯ , Vm = m m / m 1 m 1 = Hm¯ m¯ −1. The secular equationh (| i hE) −0¯ =| 0− leadsi to the recursion FIG. 13 Tridiagonal matrix elements after 100 iterations for H− | i a random and a variational pivot. c − + (E E) c + V c =0, (65) m 1 m − m m+1 m+1 whose solution is equivalent to diagonalizing a matrix be smaller and in general they will decrease uniformly20. with Vm+1 in the upper diagonal and 1 in the lower one, We shall see how to exploit this property of good pivots which is of course equivalent to the symmetric problem. to simplify the calculations. However, here we shall solve the recursion by transform- ing it into a set of non-linear coupled equations for the cm amplitudes. The profound reason for using an unnor- malized basis is that first term in Eq. (65), E = E0 +V1c1 1. The exp(S) method implies that once c1 is known the problem is solved. Call- ing εm = Em E0 and replacing E = E0 +V1c1 in Eq (65) leads to −

0.8 Random cm−1 + (εm V1c1) cm + Vm+1 cm+1 =0. (66) Variational − 0.7 Now introduce 0.6

0.5 cmPm = exp( SmPm), (67)

0.4 X X Expanding the exponential and equating terms 0.3

Wavefunction(s) 1 c = S , c = S + S2, (68) 0.2 1 1 2 2 2 1 0.1 1 3 c3 = S3 + S1S2 + S1 , etc. (69) 0 3! 0 5 10 15 20 25 30 35 40 i Note that we have used the formal identification Pm Pn = Pn+m as a heuristic way of suggesting Eqs. (68). In- FIG. 14 The ground state wavefunction in the Lanczos basis serting in Eq. (66) and regrouping, a system of coupled for a random, and a variational pivot. The plotted values are equations obtains. The first two are the squared overlaps of the successive Lanczos vectors with the final wave-function 1 0=S V + S2( V V )+ S ε + 1 (70) 2 2 1 2 2 − 1 1 1 In Section II.A.1 we have sketched the coupled-cluster 0=S3V3 + S2ε2 + S1S2(V3 V1)+ (or exp(S)) formalism. The formulation in the Lanczos − 1 2 3 1 1 basis cannot do justice to the general theory, but it is a S1 ε2 + S1 ( V3 V1)+ S1. (71) 2 3! − 2 good introduction to the underlying ideas. Furthermore, it turns out to be quite useful. Hence, instead of the usual cm-truncations we can use The construction is as in Eqs. (45,46), but the succeed- Sm-truncations, which have the advantage of providing a ing vectors—except the pivot—are not normalized. Then, model for the wavefunction over the full basis. At the first iteration Eq. (70) is solved neglecting S2. The resulting value for S1 is inserted in Eq. (71) which is solved by neglecting S3. The resulting value for S2 is reinserted 20 There are some very interesting counter-examples. One is found in Eq. (70) and the process is repeated. Once S1 and in Fig. 24, where the natural pivot is heavily fragmented. S2 are known, the equation for S3 (not shown) may be 24

m m shell nuclei, for which the f [or eventually (f7/2p3/2) ] 7 7/2 subspaces provide a good pivot. The same argument ap- 6.5 plies for other regions. For well deformed nuclei Hartree 6 Fock should provide good determinantal pivots, and 5.5 hence enormous gains in dimensionality, once projection 5 to good angular momentum can be tackled efficiently. 4.5 As we shall see next, the Lanczos and exp S procedures Energies provide a convenient framework in which to analyze other 4 -Ec(J=0) -Ec(J=2) approaches. 3.5 E(J=2)-E(J=0) exp. fits 3 2.5 2. Other numerical approximation methods 2 4 6 8 10 12 14 16 truncation t The exponential convergence method introduced in the FIG. 15 Energy convergence for ground and first previous section was first described by Horoi et al. (1999), 56 in Ni as a function of the truncation level t (Caurier et al., under a different but equivalent guise. In later work, a 1999c). hierarchy of configurations determined by their average energy and width was proposed. Successive diagonaliza- tions make it possible to reach the exact energy by expo- incorporated, and so on. Since the energy depends only nential extrapolation. The method has been successfully on S1, convergence is reached when its value remains applied to the calculation of the binding energies of the constant from step to step. With a very good pivot, pf shell nuclei (Horoi et al., 2002) and to calculation of Eq. (70) should give a fair approximation, improved by the excitation energies of the deformed 0+ states of 56Ni incorporating Eq. (71), and checked by the equation for and 52Cr (Horoi et al., 2003). S3. The work of Mizusaki and Imada (2002, 2003), is based If all amplitudes except S1 are neglected, the differ- on the fact that the width of the total Hamiltonian in a ence scheme (65) is the same that would be obtained for 0 + truncated space tends to zero as the solution approaches an harmonic εS + V (S + S−). Successive ap- H ≡ the exact one. They have devised different extrapolation proximations amount to introduce anharmonicities. An methods and applied them to some pf-shell nuclei. equivalent approach—numerically expedient—consists in Andreozzi, Lo Iudice, and Porrino (2003) have recently diagonalizing the matrices at each c truncation level. m proposed a factorization method that allows for impor- Then, after some iteration, the energies converge expo- tance sampling to approximate the exact eigenvalues and nentially: Introduce transition matrix elements. exp( ai) 1 A totally different approach which has proven its power conv(i,a,e0,i0)= e0 − − , (72) in other fields, the density matrix renormalization group, exp( ai ) 1 − 0 − has been proposed (Dukelsky and Pittel, 2001; Dukelsky et al. which equals e0 at point i = i0. Choose a so as to yield , 2002), and Papenbrock and Dean (2003) have de- conv(i = i0 +1) = e(i0 + 1) i.e., the correct energy veloped a method based in the optimization of product at the next point . Check that e(i0 + 2) is well repro- wavefunctions of protons and neutrons that seems very duced. Fig. 15 provides an example of the efficiency of promising. the method for the lowest two states in 56Ni. The index Still, to our knowledge, the only prediction for a truly i is replaced by the truncation level i t/2 [Eq. (48, large matrix that has preceded the exact calculation re- ≡ 56 with m = 16, m0 = 0]: We set i0 = 1, fix a so as to mains that of Ni, J = 2 in Fig. 15. It was borne out reproduce the energy at the second iteration and check when the diagonalization became feasible two years later. that the curve gives indeed the correct value at the third point. The results reproduce those of the S2 truncation, confirming that “exponential convergence” and exp(S) 3. Monte Carlo methods are very much the same thing. In this example, the closed shell pivot is particularly good, and the exponen- Monte Carlo methods rely on the possibility to deal tial regime sets in at the first iteration. In general, it will with the imaginary-time evolution operator acting on always set in for some value of i that may be quite large some trial wavefunction exp( β ) 0 which tends to the 0 − H | i (Fig. 14 suggests i 25 for the random pivot). Fortu- exact ground state 0¯ as β . This is very much what 0 | i ⇒ ∞ nately, it is almost always≈ possible to find good pivots, the Lanczos algorithm does, but no basis is constructed. and the subject deserves a comment. Instead, the energy (or some observable Ω)ˆ is calculated In the case of 56Ni the good pivot is a closed shell. As through a consequence, the first iterations are associated to trun- −βH/2 −βH/2 cated spaces of much smaller dimensionalities than the 0 e Ωˆ e 0 β→∞ 0¯ Ωˆ 0¯ 9 h | | i = h | | i (73) total one (dm 10 ). This also happens for lighter pf 0 e−βH 0 ⇒ 0¯ 0¯ ≈ h | | i h | i 25 which is transformed into a quotient of multidimensional integrals evaluated through Monte Carlo methods with importance sampling. A “sign problem” arises because the integrands are not definite positive, leading to enor- mous precision problems. The Green Function Monte Carlo studies mentioned at the beginning of Section II are conducted in coordinate space. The Shell Model Monte Carlo (SMMC) variant is formulated in Fock space, and hence directly amenable to comparisons with standard SM results (see Koonin, Dean, and Langanke, 1997a, for a review). The approach relies on the Hubbard- Stratonovich transformation, and the sign problem is circumvented either by an extrapolation method or by choosing Hamiltonians that are free of it while remain- ing quite realistic (e.g. pairing plus quadrupole). At present it remains the only approach that can deal with much larger valence spaces than the standard shell model. It should be understood that SMMC does not lead to detailed spectroscopy, as it only produces ground state averages, but it is very well suited for finite tempera- ture calculations. The introduction of Monte Carlo tech- niques in the Lanczos construction is certainly a tempting project.... The quantum Monte Carlo diagonalization method (QMCD) of Otsuka, Honma, and Mizusaki (1998) con- sists in exploring the mean field structure of the valence space by means of Hartree-Fock calculations that break the symmetries of the Hamiltonian. Good quantum num- bers are enforced by projection techniques (Peierls and Yoccoz, 1957). Then the authors borrow from SMMC to select an optimal set of basic states and the full Hamil- tonian is explicitly diagonalized in this basis. More basic states are iteratively added until convergence is achieved. For a very recent review of the applications of this method see (Otsuka et al., 2001b). A strong con- nection between mean field and shell model techniques is also at the heart of the Vampir approach (Petrovici et al., 1999), (Schmid, 2001) and of the projected shell model (PSM) (Hara and Sun, 1995).

C. Lanczos Strength Functions

The choice of pivot in the Lanczos tridiagonal con- FIG. 16 Evolution of the Gamow-Teller strength function of struction is arbitrary and it can be adapted to special 48Ca as the number of Lanczos iterations on the doorway state problems. One of the most interesting is the calculation increases. of strength functions (Bloom, 1984; Whitehead, 1980): If Ui j is the unitary matrix that achieves diagonal form, its first column Ui 0 gives the amplitudes of the ground state wavefunction in the tridiagonal basis while the first row U0 j determines the amplitude of the pivot in the j-th 2 eigenstate. U0 j plotted against the eigenenergies Ej is called the “strength function” for that pivot. In practice, given a transition operator , act with it on a target state t to define a pivot 0′ T = t that exhausts the sum rule| i 0′ 0′ for . Normalize| i T it. | i Then, h | i T 26 by definition the result for 50 iterations, and 250 keV widths for the non converged states. The profiles become almost iden- t tical. 0 = T | i = U0j j , whose moments, (74) | i ′ ′ | i 0 0 j h | i X p0 k 0 = U 2 Ek, are those of the (75) h |H | i 0j j j X strength function S(E)= δ(E E )U 2 , (76) − j 0j j 0~ω X V. THE CALCULATIONS As the Lanczos vector I is obtained by orthogonaliz- | i In this section we first revisit the p, sd, and pf shells to explain how a 3b monopole mechanism solves prob- lems hitherto intractable. Then we propose a sample of pf shell results that will not be discussed elsewhere. Fi- nally Gamow Teller transitions and strength functions are examined in some detail.

A. The monopole problem and the three-body interaction

The determinant influence of the monopole interaction was first established through the mechanism of Bansal and French (1964). Its efficiency in cross-shell calcula- tions was further confirmed by Zamick (1965), and the success of the ZBM model (Zuker, Buck, and McGrory, 1968) is implicitly due to a monopole correction to a real- istic force. Zuker (1969) identified the main shortcoming of the model as due to what must be now accepted as a 3b effect21. Trouble showed in 0~ω calculations somewhat later simply because it takes larger matrices to detect it FIG. 17 Strength functions convoluted with gaussians: Upper in the sd shell than in the ZBM space (dimensionalities panel; 50 iterations, Bottom panel; 1000 iterations. of 600 against 100 for six particles). Up to 5 particles the results of Halbert et al. (1971) with a realistic interaction ing I 0 to all previous Lanczos vectors i , i

− 9/2 0 − − 7/2− 1f7/2 BC 3/2− + 3/2− 5/2 7/2 4 − 5/2− − − 4 2p3/2 5/2− (1/2 ,3/2 ) 5/2 − 9/2 5/2 2p1/2 − + -5 7/2− 9/2 1f5/2 3/2− − 3 5/2 7/2 3 -10

− − 7/2 9/2− − 5/2 1/2 2 − − − 2

E (MeV) 7/2− 1/2 1/2 -15 1/2 − E (in Mev) ∆ 5/2 − 3/2 − 1 7/2 1 -20

− − − − 0 3/2 3/2 3/2 3/2 0 -25 40 50 60 70 80 KB KB' KB3 Expt. A 0 FIG. 18 The spectrum of 49Ca for the KB, KB’ and KB3 1f7/2 KB3 2p3/2 interactions compared to experiment -5 2p1/2 1f5/2 49 Fig. 18 gives an idea of what happens in Ca with -10 the KB interaction (Kuo and Brown, 1968): there are six states below 3 MeV, where only one exists. In Pasquini -15 (1976) and Pasquini and Zuker (1978) the following mod- E (in Mev) ifications were proposed (f f7/2, r f5/2, p3/2, p1/2), ≡ ≡ -20

V T (KB1) = V T (KB) ( )T 300 keV, fr fr − − -25 V 0 (KB1) = V 0 (KB) 350 keV, (77) 40 50 60 70 80 ff ff − V 1 (KB1) = V 1 (KB) 110 keV. A ff ff − FIG. 19 Effective single particle energies in the pf-shell along The first line defines KB’ in Fig. 18. The variants KB2 the N = Z line, computed with the BC and KB3 interactions. and KB3 in Poves and Zuker (1981b) keep the KB1 cen- troids and introduce very minor multipole modifications. KB3 was adopted as standard23 in successful calculations the Vfr ones Poves et al. (2001) defined a KB3G inter- in A = 47-50 that will be described in the next sec- action that brings interesting improvements over KB3 in tions (Caurier et al., 1994; Mart´ınez-Pinedo et al., 1996b, A = 50-52, but around 56Ni there are still some problems, 1997). though not as severe as the ones encountered in the sd For higher masses, there are some problems, but noth- and p shells. Compared to KB3, the interaction FPD6 56 ing comparable to the serious ones encountered in the sd has a better gap in Ni, however, the orbit 1f5/2 is defi- shell, where modifications such as in Eq. (77) are ben- nitely too low in 57Ni. This produces problems with the eficial but (apparently) insufficient. The effective single description of Gamow-Teller processes (see Borcea et al., particle energies in Fig. 19 are just the monopole values 2001, for a recent experimental check in the of the particle and hole states at the subshell closures. of 56Cu). They give an idea of what happens in a 2b description. The classic fits of Cohen and Kurath (1965) (CK) de- At the origin, in 41Ca, the spectrum is the experimental 57 fined state of the art in the p shell for a long time. As they one. At Ni there is a bunching of the upper orbits that preceded the realistic G-matrices, they also contributed the realistic BonnC (BC) and KB describe reasonably to hide the fact that the latter produce in 10B a catastro- well, but they fail to produce a substantial gap. Hence phe parallel to the one in 22Na. The work of Navr´atil and the need of KB1-type corrections. At the end of the shell 41 Ormand (2002), and Pieper, Varga, and Wiringa (2002) BC and KB reproduce an expanded version of the Ca acted as a powerful reminder that brought to the fore the spectrum, which is most certainly incorrect. The indi- 3b nature of the discrepancies. cation from ˜d in Eq. (II.B.2) is that the bunching of m Once this is understood, the solution follows: Eq. (77) the upper orbitsH should persist. By incorporating this is assumed to be basically sound but the corrections are hint—which involves the V ′ centroids—and fine-tuning rr taken to be linear in the total number of valence particles m (the simplest form that a 3b term can take).

Using f (p3/2, d5/2, f7/2) generically in the (p, sd, pf) shells≡ respectively, and r = p and r 23 The multipole changes—that were beneficial in the perturbative 1/2 ≡ treatment of Poves and Zuker (1981b)—had much less influence d3/2,s1/2 for the p and sd shells, Zuker (2003) proposes in the exact diagonalizations. (κ = κ + (m m ) κ ) 0 − 0 1 28

28 8 around Si it must be substantially smaller. The spectra NO 7 KLS in Fig. 22 are obtained with κ(m)=0.9 0.05(m 6), CK − − 6 KLS κ=1.1 and now the spectra are as good as those given by USD. EXP 5 4 7 3 KB BC E (MeV) 2 6 USD KB κ=0.80 1 EXP 5 0 -1 4 -2

E (MeV) 3 0 1 2 3 4 J 2

10 FIG. 20 The excitation spectrum of B for different interac- 1 tions. 0 0 1 2 3 4 J 5 KB V T (R) = V T (R) ( )T κ USD fr ⇒ fr − − (78) KB κ=0.55 V T (R) = V T (R) 1.5 κδ 4 EXP ff ⇒ ff − T0 where R stands for any realistic 2b potential. The results 3 for 10B are in Fig. 20. The black squares (NO) are from

(Navr´atil and Ormand, 2002, Fig 4, 6~Ω). E (MeV) 2

4.5 1 KB 4 USD KB κ=0.90 0 3.5 EXP 1 3 5 3 2J 2.5 24 29 2 FIG. 22 The excitation spectra of Mg and Si for different

E (MeV) 1.5 interactions.

1

0.5 To determine a genuine 3b effect (i.e., the linearity of κ) a sufficiently large span of A values is necessary. 0 In the p shell, corrections to Vrr′ become indispensable -0.5 0 1 2 3 4 5 6 very soon and must be fitted simultaneously, as done J successfully by Abzouzi et al. (1991), so we do not dwell

22 on the subject. As we have seen, in the pf shell KB3, FIG. 21 The excitation spectrum of Na for different inter- which is nearly perfect in A = 47-50, must be turned into actions. KB3G for A = 50-52, which also does well at the lower masses. To find a real problem with a 2b R-interaction The black circles correspond to the bare KLS G-matrix 24 ~ (i.e., compatible with NN data) we have to move to ( ω =17 MeV), the white squares to the same with κ = 56,58Ni. In particular the first B(E2)(2 0) transition 1.1, and the pentagons to the CK fit. NO and KLS give in 58Ni falls short of the observed value−→ (140 e2fm4) by a quite similar spectra, as expected from the discussion factor 0.4 with any monopole corrected KB interaction. in Sections II.A and II.D. The κ correction eliminates The problem≈ can be traced to weak quadrupole strength the severe discrepancies with experiment and give values and it is not serious: as explained in Section II.D it is due close to CK. 22 to a normalization uncertainty, and Table I shows that In Na the story repeats itself: BC and KB are very with equal normalizations BC and KB in the pf shell are close to one another, the ground state spin is again J =1 instead of J = 3 and the whole spectrum is awful. The κ correction restores the levels to nearly correct positions, though USD still gives a better fit. This simple cure 24 It may be possible to fit the data with a purely 2b set of matrix el- was not discovered earlier because κ is not a constant. ements: USD is the prime example, but it is R-incompatible (Du- In 24Mg we could still do with the value for 22Na but four and Zuker, 1996). 29 as close as in the sd shell. within 100 200 keV and to give a fairly good descrip- In (Zuker, 2003) BC was adopted. Small modifications tion of the∼ highly deformed excited band of this doubly of the Vfr matrix were made to improve the (already rea- magic nucleus (Mizusaki et al., 2002, 1999; Otsuka et al., sonable) r-spectrum in 57Ni (see Fig. 19). The particular 1998). Other applications to the pf shell can be found mixture in Eq. (78) was actually chosen to make possi- in (Honma et al., 1996). The existence of excited collec- ble, in the simplest way, a good gap in 48Ca and a good tive bands in the N=28 isotones is studied in (Mizusaki single-particle spectrum in 49Sc. It was found that for et al., 2001). Very recently, a new interaction for the A = 48, κ(m = 8) 0.43. For A = 56, truncated calcu- pf shell (GXPF1) has been produced by a Tokyo-MSU lations yielded κ(16)≈ / 0.28. The B(E2)(2 0) in 58Ni collaboration (Honma et al., 2002), following the fitting indicated convergence to the right value. → procedures that lead to the USD interaction. The fit starts with the G-matrix obtained from the Bonn-C nu- cleon nucleon potential (Hjorth-Jensen et al., 1995). The 16 fit privileges the upper part of the pf shell, as seen by 14 the very large difference in single particle energies be- 12 tween the 1f7/2 and 12p3/2 orbits (3 MeV instead of the standard 2 MeV). In the calcium isotopes i.e., when only 10 the T=1 neutron neutron interaction is active) this new J 8 interaction retains the tendency of the bare G-matrices 6 (KB, Bonn-C, KLS) to produce large gaps at N=32 and N=34, contrary to FPD6 or KB3G that only predict a 4 BC κ=0.43 large gap at N=32. Some early spectroscopic applications BC κ=0.28 2 KB3 of GXPF1 to the heavy isotopes of Titanium, EXP 0 and Chromium can be found in (Janssens et al., 2002) 500 1000 1500 2000 2500 3000 3500 and (Mantica et al., 2003), where the N=32-34 gaps are ∆(E ) (keV) J explored. FIG. 23 Backbending in 48Cr. See text. Each nucleus has its interest, sometimes anecdotic, sometimes fundamental. The agreement with experiment With κ(m = 8), BC reproduces the yrast spectrum of for the energies, and for the quadrupole and magnetic 48Cr almost as well as KB3, but not with κ(m = 16) moments and transitions is consistently good, often ex- (Fig. 23), indicating that the three-body drift is needed, cellent. These results have been conclusive in establish- though not as urgently as in the sd shell. Another indica- ing the soundness of the minimally-monopole-modified tion comes from the T=0 spectrum of 46V, the counter- realistic interaction(s). There is no point in reproducing part of 22Na and 10B in the pf-shell. The realistic inter- them here, and we only present a few typical examples actions place again the J=1 and J=3 states in the wrong concerning spectroscopic factors, isospin non-conserving order, the correct one is re-established by the three-body forces and “pure spectroscopy”. monopole correction.

B. The pf shell

Systematic calculations of the A=47-52 isobars can be 1. Spectroscopic factors found in (Caurier et al., 1994, A = 48, KB3), (Mart´ınez- Pinedo et al., 1997, A = 47 and 49, KB3) and (Poves et al., 2001, A = 50-52, KB3G). Among the other The basic tenet of the independent particle model is ~ † full 0 ω calculations let’s highlight the following: The that addition of a particle to a closed shell ar cs pro- SMMC studies using either the FPD6 interaction (Al- duces an eigenstate of the cs +1 system. Nowadays| i we hassid et al., 1994) or the KB3 interaction (Langanke know better: it produces| a doorwayi that will be frag- et al., 1995). A comparison of the exact results with mented. If we choose cs = 48Ca and cs +1 = 49Sc , the SMMC can be found in (Caurier et al., 1999b). For the four pf orbits provide| i the| doorways.i | Thei lowest,| i a recollection of the SMMC results in the pf shell see f7/2, leads to an almost pure eigenstate. The middle also (Koonin et al., 1997b). The recent applications of ones, p3/2, 1/2, are more fragmented, but the lowest level the exponential extrapolation method (Horoi et al., 2002, still has most of the strength and the fragments are scat- 2003). The calculations of Novoselsky et al. (1997) for tered at higher energies. Fig. 24 shows what happens to 51 51 Sc and Ti using the DUPSM code (see also the erra- the f5/2 strength: it remains concentrated on the door- tum in Novoselsky et al., 1998b) and for 52Sc and 52Ti way but splits locally. The same evolution with energy (Novoselsky et al., 1998a). The extensive QMCD calcu- of the quasi-particles (Landau’s term for single particle lation of the spectrum of 56Ni have been able to repro- doorways) was later shown to occur generally in finite duce the exact result for the ground state binding energy systems (Altshuler et al., 1997). 30

2.0 A=46 isospin triplet in (Garrett et al., 2001). Th. Exp. 1.0 3. Pure spectroscopy

The first nucleus we have chosen is also one that has

S 0.0 been measured to complete the mirror band (Bentley et al., 2000) in A = 51, whose MED are as well described as those of A = 49 in Fig. 25. Such calculations need very −1.0 good wavefunctions, and the standard test they have to pass is the “purely spectroscopic” one. Work on the subject usually starts with a litany. Such −2.0 as: Quadrupole effective charges for neutrons qν =0.5 and s s 0.0 5.0 10.0 15.0 20.0 protons qπ=1.5 and bare g-factors gπ=5.5857 µN , gν =- Energy (MeV) l l 3.3826 µN , gπ=1.0 µN and gν =0.0 µN in M1 transitions and moments are used. Except when the M1 transitions FIG. 24 Spectroscopic factors, (2j +1)S(j, tz), corresponding are fully dominated by the spin term, the use of effective to stripping of a particle in the orbit 1f5/2 (Mart´ınez-Pinedo g-factors does not modify the results very much due to et al., 1997). the compensation between the spin and orbital modifica- tions. 51 2. Isospin non conserving forces Then some spectrum: The yrast band of Mn, calcu- lated in the full pf-shell space, is compared in Fig. 26 Recent experiments have identified several yrast bands with the experimental data. The first part of the test in mirror pf nuclei (Bentley et al., 1998, 1999, 2000; is passed. At most the examiner will complain about Brandolini et al., 2002b; Lenzi et al., 2001; O’Leary et al., slightly too high, high spin states. 1997, 2002) for A = 47, 49, 50 and 51. The naive view − that the Coulomb energy should account for the MDE 8 25/2 − (mirror energy differences) turns out to be untenable. 27/2 27/2− The four pairs were analyzed in (Zuker et al., 2002) where 7 23/2− it was shown that three effects should be taken into ac- 23/2− 6 21/2− count. A typical result is proposed in Fig. 25, where VCM 21/2− and VBM stand for Coulomb an nuclear isospin break- 5 ing multipole contributions, while V is a monopole − Cm 19/2 19/2− Coulomb term generated by small differences of radii be- 4 − 17/2− 17/2 E(MeV) E(MeV) − tween the members of the yrast band. The way these 15/2− 15/2 13/2− disparate contributions add to reproduce the observed 3 13/2− pattern is striking. − − 1/2 1/2 − 2 3/2−,5/2−,7/2− 3/2 − − 11/2 11/2 − − 9/2 0.1 1 9/2 Exp − 0.08 VCm+VCM+VBM 7/2− 7/2 − − VCm 0 5/2 5/2 0.06 V CM Exp. KB3G VBM 0.04 FIG. 26 Yrast band of 51Mn; experiment vs shell model cal- 0.02 culation in the full pf-shell space. 0 MED (MeV) -0.02 Finally the transitions: The transitions in Table IV -0.04 A=49 are equally satisfactory. Note the abrupt change in both B(M1) and B(E2) for J = (17/2)−, beautifully repro- -0.06 5 7 9 11 13 15 17 19 21 23 25 27 duced by the calculation. The origin of this isomerism is 2J in the sudden alignment of two particles in the 1f7/2 or- bit, which provides an intuitive physical explanation for FIG. 25 Experimental (O’Leary et al., 1997) and calcu- the abrupt change in the MED (Bentley et al., 2000). lated (Zuker et al., 2002) MED for the pair 49Cr-49Mn. At the end one can add some extras: The electromag- netic moments of the ground state are also known: Their 2 Another calculation of the isospin non conserving ef- values µexp=3.568(2)µN and Qexp=42(7) efm (Fire- fects, and their influence in the location of the proton drip stone, 1996) compare quite well with the calculated 2 line, is due to (Ormand, 1997). He has also analyzed the µth=3.397µN and Qth=35 efm . 31

TABLE IV Transitions in 51Mn. TABLE V Transitions in 52Cr. (a) from (Brown et al., 1974).

Exp. Th. 2 2 Exp. Th. B(M1) (µN ) (µN ) 7 − 5 − 2 2 B(M1) (µN ) (µN ) 2 → 2 0.207(34) 0.177 9 − 7 − 9+ → 8+ 0.057(38) 0.040 2 → 2 0.16(5) 0.116 11 − 9 − 2 4 2 4 2 → 2 0.662(215) 0.421 B(E2) (e fm ) (e fm ) 17 − 15 − → + + 2 2 0.00012(4) 0.00003 2 → 0 131(6) 132 19 − 17 − → >0.572 0.797 + + +7 2 2 3 → 2 7−5 5 2 4 2 4 + + a B(E2) (e fm ) (e fm ) 41 → 2 83(17) 107 7 − 5 − + + 42 → 2 69(19) 26 2 → 2 528(146) 305 9 − 5 − 6+ → 4+ 59(2) 68 2 → 2 169(67) 84 9 − 7 − 8+ → 6+ 75(24) 84 2 → 2 303(112) 204 11 − 7 − 9+ → 8+ 0.5(20) 0.6 2 → 2 236(67) 154 11 − 9 − 2 → 2 232(75) 190 17 − 13 − 2 → 2 1.236(337) 2.215

+ 12 16 + 16 11

We complete this section with an even-even and an 10 15+ 15+ odd-odd nucleus. In figure 27 we show the yrast states of + + 52 9 14 14 Cr, up to the band termination, calculated in the full + + 8 13 13 + pf-shell space. The agreement of the KB3G results with 12+ 12 the experiment is again excellent. 7 6

16 E(MeV) E(MeV) 16+ 5 15 + + (10 ) 10 + 14 4 (11+) 11

+ + 13 14 3 (9+) 9 + 12 (8+) 8 2 11 + 5 + + + 5 + + 1 7+ 7 10 3 + 12 4 + + 3 + + 4 12 1 + 2 1 + + 2 9 + 0 6 6 8 + Exp. KB3G 10+ 10 E(MeV) E(MeV) 7 6 FIG. 28 Yrast band of 52Mn; experiment vs shell model cal- 8+ 5 + 8 culation in the full pf-shell space. 4 6+ 6+ 3 + + 4 4

2 + 2+ 2 1 tice the perfect correspondence between theory and ex- + 0 0+ 0 periment for the states belonging to the multiplet below Exp. KB3G 1 MeV. The experimental data for the spins beyond the band termination (11+ to 16+) come from a recent ex- FIG. 27 Yrast band of 52Cr; experiment vs shell model cal- periment (Axiotis, 2000). The agreement between theory culation in the full pf-shell space. and experiment is spectacular. There are only three ex- perimental states with spin assignment not drawn in the The electromagnetic moments of the first J =2+ state figure: a second J = (5+) at 1.42 MeV, a third J = (5+) + are known: µexp=2.41(13) µN (Speidel et al., 2000) and at 1.68MeV and a second J = (6 ) at 1.96 MeV. The cal- 2 + Qexp = 8.2(16) e fm (Firestone, 1996) and they are culation places the second J =5 at 1.37 MeV, the third in very good− agreement with the theoretical predictions: J =5+ at 1.97MeV and the second J =6+ at 1.91 MeV. 2 µth=2.496 µN and Qth=–9.4 e fm In table V the experi- The experimental magnetic and quadrupole moments for + mental values for the electromagnetic transitions between the 6 ground state are also known; µexp=3.062(2) µN 2 the states of the yrast sequence are given. The calculated and Qexp=+50(7) e fm (Firestone, 1996). The calcu- 2 values are in very good correspondence with the experi- lated values, µth=2.952 µN and Qth=+50 e fm repro- ment. duce them nicely. Some electromagnetic transitions along The yrast states of the odd-odd nucleus 52Mn, calcu- the yrast sequence have been measured. The calculations lated in the full pf-shell, are displayed in fig. 28. No- are in very good agreement with them (see table VI). 32

B(F ) and B(GT ) are defined as TABLE VI Transitions in 52Mn. 2 f tk i B(F )= h k k ± k i (80a) Exp. Th. √2J +1  P i  2 2 2 B(M1) (µN ) (µN ) σktk gA f k ± i 7+ → 6+ 0.501(251) 0.667 B(GT )= h k k i . (80b) gV √2Ji +1 8+ → 7+ >0.015 0.405   P  9+ → 8+ 1.074+3.043 0.759 Matrix elements are reduced (A3) with respect to spin −0.537 ± only, refers to β decay, σ = 2S and (gA/gV ) = 2 4 2 4 B(E2) (e fm ) (e fm ) 1.2720(18)± (Hagiwara et al., 2002) is the ratio of the + + +484 − 7 → 6 92−81 126 weak interaction axial-vector and vector coupling con- 8+ → 6+ >1.15 33 stants. 8+ → 7+ >4.15 126 For states of good isospin the value of B(F ) is fixed. It + + +300 can be altered only by a small isospin-symmetry-breaking 9 → 7 104−46 66 11+ → 9+ 54(6) 53 δC correction (Towner and Hardy, 2002). Shell model estimates of this quantity can be also found in (Ormand and Brown, 1995). B(F )= T (T + 1) T T δ (1 δ ), (81) − zi zf if − C where δif allows only transitions between isobaric analog C. Gamow Teller and magnetic dipole strength. states. Superallowed decays may shed light on the depar- tures from unitarity of the Cabibbo Kobayashi Maskawa matrix. Out of the approximately 2500 known nuclei that are bound with respect to nucleon emission, only 253 are The total strengths S± are related by the sum rules stable. The large majority of the rest decay by β emis- S−(F ) S+(F )= N Z, (82a) sion or capture, mediated by the weak interac- − − S−(GT ) S (GT )=3(N Z), (82b) tion. When protons and neutrons occupy the same or- − + − bits, as in our case, the dominant processes are allowed where N and Z refer to the initial state and S±(GT ) Fermi and Gamow-Teller transitions. The information does not contain the gA/gV factor. The comparison of obtained from the weak decays has been complemented the (p,n) and (n,p) data with the Gamow Teller sum rule by the (p,n) and (n,p) reactions in forward kinematics, (Ikeda et al., 1963) led to the long standing “quenching” that make it possible to obtain total GT strengths and problem; only approximately one half of the sum rule strength functions that cannot be accessed by the decay value was found in the experiments. data because of the limitations due to the Qβ windows. The GT strength is not protected by a conservation From a theoretical point of view, the comparison of calcu- principle and depends critically on the wavefunctions lated and observed strength functions provide invaluable used. Full 0~ω calculations already show a large quench- insight into the meaning of the valence space and the na- ing with respect to the independent particle limit as seen ture of the deep correlations detected by the “quenching” in Table VII where the result of a full pf calculation is effect. compared with that obtained with an uncorrelated Slater The half-life for a transition between two nuclear states determinant having the same occupancies: is given by (Behrens and B¨uhring, 1982; Schopper, 1966): npnh S = i k i σ k 2 (83) (2ji + 1)(2jk + 1)h || || i Xi,k ǫ 6144.4 1.6 (fA + f )t = ± . (79) the sum on i runs over the proton (neutron) orbits in the (fV /fA)B(F )+ B(GT ) valence space and k over the proton (neutron) orbits for p h S+ (S−). n and n denote the number of particles and holes, respectively. The determinantal state demands a The value 6144.4 1.6 is obtained from the nine best- quenching factor that is almost twice as large as the stan- ± known superallowed beta decays (Towner and Hardy, dard quenching factor Q2 = (0.74)2) that brings the full 2002) (see Wilkinson, 2002a,b, for an alternative study). calculation in line with experiment. fV and fA are the Fermi and Gamow-Teller phase-space factors, respectively (Chou et al., 1993; Wilkinson and Macefield, 1974). f ǫ is the phase space for electron cap- 1. The meaning of the valence space ture (Bambynek et al., 1977), that is only present in β+ decays, If t1/2 is the total lifetime, the partial lifetime of Before moving to the explanation of the quenching of a level with branching ratio br is t = t1/2/br. the GT strength, it is convenient to recall the meaning 33

54 55 35 TABLE VII Comparison of GT+ strengths. For Fe, Mn, Calculated 58 59 Experimental Ni and Co the calculations are truncated to t = 8, t = 4, 30 t = 6 and t = 4 respectively. The data are from (Alford et al., 1993; El-Kateb et al., 1994; Vetterli et al., 1990; Williams 25 et al., 1995). 20

Nucleus Uncorrelated Correlated Expt. 15 Unquenched Q = 0.74 10 51V 5.15 2.42 1.33 1.2 ± 0.1 5 54Fe 10.19 5.98 3.27 3.3 ± 0.5 Gamow Teller Strength B(GT) 55Mn 7.96 3.64 1.99 1.7 ± 0.2 0 5 10 15 20 25 30 56 Fe 9.44 4.38 2.40 2.8 ± 0.3 E (MeV) 58Ni 11.9 7.24 3.97 3.8 ± 0.4 48 48 59Co 8.52 3.98 2.18 1.9 ± 0.1 FIG. 29 Gamow Teller strength in Ca(p, n) Sc from (An- derson et al., 1990)—after elimination of the Fermi peak at 62Ni 7.83 3.65 2.00 2.5 ± 0.1 around 6 MeV—compared with the calculated peaks (KB3 in- teraction) after 700 Lanczos iterations (Caurier et al., 1995b). The peaks have been smoothed by gaussians having the in- of the valence space as discussed in Section II.A.1. The strumental width of the first measured level. 48Ca(p,n)48Sc reaction (Anderson et al., 1990) provides an excellent example. In Fig. 29, adapted from Caurier for about 30 states below 8 MeV? et al. (1995b), the experimental data are compared with the strength function produced by a calculation in the full pf-shell using the interaction KB3. The peaks have 2. Quenching all J = 1, T = 3. In the pf shell there are 8590 of them and the calculation has been pushed to 700 iterations in To understand the quenching problem it is best to start the Lanczos strength function to ensure fully converged by the (tentative) solution. The dressed states in Eq. (4) eigenstates below 11 MeV. Of these eigenstates 30 are are normalized to unity in the model space. This trick is below 8 MeV: They are at the right energy and have the essential in the formulation of linked cluster or exp S the- right strength profile. At higher energies the peaks are ories (hence Eq. (63)). It makes possible the calculation much too narrow compared with experiment. It means of the energy—and some transitions, such as the E2— that they may well be eigenstates of the effective Hamil- without knowledge of the norm of the exact wavefunc- tonian in the pf shell, but not eigenstates of the full tion. In general though, we need an expectation value system. Therefore, they should be viewed as doorways, between exact, normalized states: fˆ ˆi 2. If we subject to further mixing with the background of intrud- write h kT k i ers which dominates the level density after 8 MeV, as corroborated by the experimental tail that contains only ˆi = α 0~ω + β n~ω , (84) | i | i n| i intruders and can be made to start naturally at that en- n=06 X ergy. and a similar expression in α′, β′ for fˆ , we find The KB3 effective interaction is doing a very good job, h | but it is certainly not decoupling 8590 pf states from 2 the rest of the space. If the fact is not explicitly rec- fˆ ˆi 2 = αα′ T + β β′ T , (85) ognized we end up with the often raised (Hjorth-Jensen h kT k i  0 n n n 6 et al., 1995) “intruder problem”: Decoupling cannot be nX=0   enforced perturbatively when intruders are energetically since the GT operator does not couple states with differ- close to model states. Figure 29 indicates that although ent number of ~ω excitations. If we make two assump- few eigenstates are well decoupled, it is possible to make tions: (a) neglect the n = 0 contributions; (b) α α′; sense of many others if one interprets them as doorways. it follows that if the projection6 of the physical wavefunc-≈ The satisfactory description of the lowest states indicates tion in the 0~ω space is Q α2, its contribution to the that the model space makes sense. It means that it en- transition will be quenched≈ by Q2. sures good decoupling at the S2 level, or simply in sec- Exactly the same arguments apply to transfer ond order perturbation theory, which guarantee a state- reactions—for which = a (or a†)—but are simpler T s s independent interaction. because Tn=06 = 0. The transition strength is given by Energetically we are in good shape, and we concentrate the spectroscopic factor, which can be identified with Q on the renormalization of the GT operator: the calcu- when one particle is removed and 1 Q when it is added. lated strength has been quenched by a factor (0.74)2. The assumption that the model amplitudes− in the ex- Why? Why this factor ensures the right detailed≈ strength act wavefunctions are approximately constant is borne 34 out by systematic calculations of Q in the p shell (Chou strength, constrained by the Ikeda sum rule that relates et al., 1993, Q = 0.820(15)), the sd shell (Wildenthal the direct and inverse processes. The careful analysis et al., 1983, Q =0.77(2)) and the pf shell (illustrated in of Anderson et al. (1985) suggests, but does not prove, Fig. 30, Mart´ınez-Pinedo et al., 1996b, Q =0.744(15)). that the experimental tail in Fig. 29 contains enough strength to satisfy approximately the sum rule. A similar 1.0 result is obtained for 54Fe(p,n) (Anderson et al., 1990). More recent experiments by the Tokyo group establish 0.8 0.77 that the strength located at accessible energies exhausts 0.744 90(5)% of the sum rule in the 90Zr(n,p) and 84(5)% in 27Al(p,n) (Wakasa et al., 1998, 1997). 0.6 The theoretical problem is to calculate Q. It has been compounded by a sociological one: the full GT operator 0.4 is (gA/gV )στ, where gA/gV 1.27 is the ratio of weak T(GT) Exp. axial and vector coupling constants.≈− The hotly debated 0.2 question was whether Q was due to non nucleonic renor- malization of gA or nuclear renormalization of στ (Arima, 0.0 2001; Osterfeld, 1992). We have sketched above the nu- 0.0 0.2 0.4 0.6 0.8 1.0 clear case, along the lines proposed by Caurier, Poves, T(GT) Th. and Zuker (1995b), but under a new guise that makes it easier to understand. The calculations of Bertsch and FIG. 30 Comparison of the experimental and theoretical val- Hamamoto (1982), Dro˙zd˙z et al. (1986) and Dang et al. ues of the quantity T (GT ) in the pf shell (Mart´ınez-Pinedo (1997) manage to place significant amounts of strength et al., 1996b). The x and y coordinates correspond to theo- beyond the resonance region, but they are based on 2p-2h retical and experimental values respectively. The dashed line doorways which fall somewhat short of giving a satisfac- shows the “best-fit” for Q = 0.744. The solid line shows the tory view of the strength functions. No-core calculations result obtained in the sd-shell nuclei (Wildenthal et al., 1983) are under way, that should be capable of clarifying the issue. These numbers square well with the existing informa- The purely nuclear origin of quenching is borne out by tion on spectroscopic factors from (d, p) (Vold et al., 1978, (p,p′), (γ,γ′) and (e,e′) experiments that determine the Q 0.7) and (e,e′p) (Cavedon et al., 1982, Q =0.7) data spin and convection currents in M1 transitions, in which ≈ (see also Pandharipande et al., 1997). This consistency is gA/gV play no role (see Richter, 1995, for a complete re- significant in that it backs assumption (a) above, which view). An analysis of the data available for the N = 28 is trivially satisfied for spectroscopic factors. It opens isotones in terms of full pf-shell calculations concluded the perspective of accepting the GT data as a measure that agreement with experiment was achieved by quench- of a very fundamental quantity that does not depend ing the στ operator by a factor 0.75(2), fully consistent on particular processes. The proposed “solution” to the with the value that explains the Gamow-Teller data (von quenching problem amounts to reading data. Neumann-Cosel et al., 1998) (see Fig. 31). These re- sults rule out the hypothesis of a renormalization of the axial-vector constant gA: it is the στ operator that is quenched.

VI. SPHERICAL SHELL MODEL DESCRIPTION OF NUCLEAR ROTATIONS

Theoretical studies in the pf shell came in layers. The first, by McCullen, Bayman, and Zamick (1964), re- stricted to the f7/2 space, was a success, but had some drawbacks: the spectra were not always symmetric by in- terchange of particles and holes, and the quadrupole mo- ments had systematically the wrong sign. The first diag- onalizations in the full shell [(Pasquini, 1976), (Pasquini FIG. 31 Effective spin g-factor of the M1 operator deduced and Zuker, 1978)] solved these problems to a large extent, from the comparison of shell model calculations and data for but the very severe truncations necessary at the time the total B(M1) strengths in the stable even mass N=28 iso- made it impossible to treat on the same footing the pair- tones (from von Neumann-Cosel et al. (1998)) ing and quadrupole forces. The situation improved very much by dressing perturbatively the pure two-body f7/2 Experimentally, the challenge is to locate all the part of the Hamiltonian H2, with a three body term HR1, 35

+ mostly due to the quadrupole force (Poves and Zuker, 16 16268 + 1981b). The paper ended with the phrase: (16 ) 15727 “It may well happen, that in some cases, not in the pf shell but elsewhere, HR1 will overwhelm H2. Then, and we are only speculating, we shall speak, perhaps, of the + 16 13885 rotational coupling scheme.” + Indeed, some nuclei were indicating a willingness to 16 13306 5674 5450 become rotational but could not quite make it, simply because the perturbative treatment was not enough for 3291 them. The authors missed prophecy by one extra condi- 3029 tional: “not in the pf shell should have been “not neces- + sarily...” 14 10594 14+ 10277 At the time it was thought impossible to describe ro- tational motion in a spherical shell-model context. The 2135 glorious exception, discovered by Elliott (1958a,b), was 1871 + + (apparently) associated to strict SU(3) symmetry, ap- 12 8406 12 8459 proximately realized only near 20Ne and 24Mg. 1343 1481 + 10 7063 10+ 6978 A. Rotors in the pf shell 1874 1850 + + 8 5189 8 5128 TABLE VIII 48Cr; quadrupole properties of the yrast band 1744 1730 J B(E2)exp B(E2)th Q0(t) Q0(s) Q0(t)[f7/2,p3/2] + 6 3445 6+ 3398 2 321(41) 228 107 103 104 4 330(100) 312 105 108 104 1587 1575 + + 6 300(80) 311 100 99 103 4 1858 4 1823 8 220(60) 285 93 93 102 1106 1017 + + 10 185(40) 201 77 52 98 2 752 2 806 12 170(25) 146 65 12 80 + 752 806 0 0 0+ 0 14 100(16) 115 55 13 50 exp.48 theor. 16 37(6) 60 40 15 40 Cr

The fourth layer was started when the ANTOINE code FIG. 32 48Cr level scheme; experiment vs. theory (Section III.D) came into operation: 48Cr was definitely a well deformed rotor (Caurier et al., 1994), as can be surmised by comparing with the experimental spectrum from Lenzi et al. (1996) in Fig. 32 (see also Cameron When treated in the cranked Hartree Fock Bogoliubov et al., 1993) and the transition properties in Table VIII (CHFB) approximation, the results are not so good. The from Brandolini et al. (1998) where we have used discrepancy is more apparent than real: The predictions for the observables are very much the same in both cases (J + 1)(2J + 3) (see Caurier et al., 1995a, for the details). The reason Q (s)= Q (J),K = 1 (86) 0 3K2 J(J + 1) spec 6 is given in Fig 34, where exact KB3 diagonalizations are − B(E2,J J 2) = done, subtracting either of the two pairing contributions, → − JT = 01 and 10 (Poves and Mart´ınez-Pinedo, 1998). It 5 2 2 2 e JK20 J 2,K Q0(t) K =1/2, 1; (87) is apparent that the JT = 01-subtracted pattern is quite 16π |h | − i| 6 close to the CHFB one in Fig. 33, especially in the ro- to establish the connection with the intrinsic frame de- tational regime before the backbend. The JT = 10 sub- scriptions: A good rotor must have a nearly constant Q0, traction goes in the same direction. The interpretation is which is the case up to J = 10, then 48Cr backbends. transparent: CHFB does not “see” proton-neutron pair- This is perhaps the most striking result to come out of ing at all, and it is not very efficient in the low pair- our pf calculations, because such a behaviour had been ing regime. As it does everything else very well, the in- thought to occur only in much heavier nuclei. evitable conclusion is that pairing can be treated in first Fig. 33 compares the experimental patterns with those order perturbation theory, i.e., the energies are very sen- obtained with KB3, and with the Gogny force. The lat- sitive to it, but not the wavefunctions. Floods of ink have ter, when diagonalized gives surprisingly good results. gone into “neutron-proton” pairing, which is a problem 36

16 20 18 14 EXP 16 KB3 12 14

10 12 J J 10 8 8 Exp 6 6 SM-KB3 4 4 CHFB SM-GOGNY 2 2 0 0.0 1.0 2.0 3.0 4.0 0 1000 2000 3000 4000 5000 6000 Eγ (MeV) E_gamma (in keV)

FIG. 33 The yrast band of 48Cr; experiment vs. the shell FIG. 35 The yrast band of 50Cr; experiment vs. the shell model calculations with KB3 and the Gogny force and the model calculations with the KB3 interaction CHFB results also with the Gogny force.

16 for by the shell model calculations (Poves and Zuker, 14 1981b)(sic), (Ur et al., 1998). For spectroscopic com- parisons with the odd-odd nuclei, see Brandolini et al. 12 (2001); Lenzi et al. (1999) for 46V and Svensson et al. 50 10 (1998) for Mn. Recently, a highly deformed excited J band has been discovered in 56Ni (Rudolph et al., 1999). 8 12 It is dominated by the configuration (1f7/2) (2p3/2, 4 6 1f5/2, 2p1/2) . The calculations reproduce the band, that KB3 starts at about 5 MeV excitation energy and has a defor- KB3-P10 4 KB3-P01 mation close to β=0.4. 2 0.0 1.0 2.0 3.0 4.0 Eγ (MeV)

FIG. 34 Gamma-ray energies along the yrast band of 48Cr (in B. Quasi-SU(3). MeV), full interaction (KB3), isoscalar pairing retired (KB3– P01) and isovector pairing retired (KB3–P01). To account for the appearance of backbending rotors, a theoretical framework was developed by Zuker et al. (1995), and made more precise by Mart´ınez-Pinedo et al. for mean field theories but neither for the Gogny force (1997). Here we give a compact overview of the scheme nor for the shell model. Furthermore, the results show that will be shown to apply even to the classic examples that ordinary pairing is also a mean field problem when of rotors in the rare-earth region. nuclei are not superfluids. 48Cr has become a standard benchmark for mod- Let us start by considering the quadrupole force alone, els (Cranked Hartree Fock Bogoliuvov (Caurier et al., taken to act in the p-th oscillator shell. It will tend to 1995a), Cranked Nilsson Strutinsky (Juodagalvis and maximize the quadrupole moment, which means filling ˚ the lowest orbits obtained by diagonalizing the operator Aberg, 1998), Projected Shell Model (Hara et al., 1999), 2 2 2 Cluster Model (Descouvemont, 2002), etc.) Nuclei in the Q0 = 2q20 = 2z x y . Using the cartesian repre- sentation, 2q =2−n −n n , we find eigenvalues 2p, vicinity also have strong rotational features. The mirror 20 z − x − y 47 47 49 49 2p 3,. . . , etc., as shown in the left panel of Fig. 36, where pairs V- Cr and Cr- Mn closely follow the semiclas- − sical picture of a particle or hole hole strongly coupled to spin has been included. By filling the orbits orderly we a rotor (Mart´ınez-Pinedo et al., 1997) in full agreement obtain the intrinsic states for the lowest SU(3) represen- with the experiments (Bentley et al., 1998; Cameron tations (Elliott, 1958a,b): (λ, 0) if all states are occupied et al., 1991, 1994; O’Leary et al., 1997; Tonev et al., up to a given level and (λ, µ) otherwise. For instance: 2002). 50Cr was predicted to have a second backbending putting two neutrons and two protons in the K = 1/2 by Mart´ınez-Pinedo et al. (1996a), confirmed experimen- level leads to the (4p,0) representation. For four neu- tally (Lenzi et al., 1997) (see figure 35). When more par- trons and four protons, the filling is not complete and we have the (triaxial) (8(p 1),4) representation for which ticles or holes are added, the collective behaviour fades, − though even for 52Fe, a rotor like band appears at low we expect a low lying γ band. spin with an yrast trap at J=12+, both are accounted In jj coupling the angular part of the quadrupole op- 37

where Q + k SU3 0 quasi.SU3 δ 1 2q20 2q20 = h i = h i . (91) 3 4 r2 (p +3/2)4 h i 12 9/2 9/2 In the right panel of Fig. 37 the results are given in the usual form. 9 7/2 7/2

10.0 5/2[303] 3/2[301] 6 1/2[301] 5/2[303] 5/2 5/2 5.0 1/2[301] 3/2[301] 1/2[310] 3/2[312] 3/2[312] 1/2[310] 3 0.0 7/2[303] 3/2 3/2 1/2[321] 5/2[312] E (MeV) 7/2[303] 1/2[321] ∆ −5.0 5/2[312] 3/2[321] 0 3/2[321] 1/2[330] 1/2 1/2 1/2[330] −10.0 0.0 0.5 1.0 1.5 2.0 −0.30 −0.10 0.10 0.30 FIG. 36 Nilsson orbits for SU(3) (k = −2p) and quasi-SU(3) ε δ (k = −2p + 1/2)). FIG. 37 Nilsson diagrams in the pf shell. Energy vs. sin- gle particle splitting ε (left panel), energy vs. deformation δ erator q20 = r2C20 has matrix elements (right panel). 3[(j +3/2)2 m2] 2 In the left panel we have turned the representation jm C j +2 m −2 , (88) h | | i≈ 2(2j + 3) around: since we are interested in rotors, we start from 3m[(j + 1)2 m2]1/2 perfect ones (SU(3)) and let ε increase. At a value of jm C2 j +1 m = − (89) h | | i − 2j(2j + 2)(2j + 4) 0.8 the four lowest orbits are in the same sequence as the≈ right side of fig. 36 (Remember here that the real The ∆j = 2 numbers in Eq. (88) are—within the ap- situation corresponds to ε 1.0). The agreement even proximation made—identical to those in LS scheme, ob- extends to the next group,≈ although now there is an in- tained by replacing j by l. The ∆j = 1 matrix elements truder (1/2[310] orbit). The suggestion is confirmed by in Eq. (88) are small, both for large and small m, corre- an analysis of the wavefunctions: For the lowest two or- sponding to the lowest oblate and prolate deformed orbits bits, the overlaps between the pure quasi-SU(3) wave- respectively. If the spherical j-orbits are degenerate, the functions calculated in the restricted ∆j = 2 space (fp ∆j = 1 couplings, though small, will mix strongly the from now on) and the ones in the full pf shell exceeds two ∆j = 2 sequences (e.g., (f7/2p3/2) and (f5/2p1/2)). 0.95 throughout the interval 0.5 <ε< 1. More interest- The spin-orbit splittings will break the degeneracies and ing still: the contributions to the quadrupole moments favour the decoupling of the two sequences. Hence the from these two orbits vary very little, and remain close idea (Zuker et al., 1995) of neglecting the ∆j = 1 matrix to the values obtained at ε =0 i.e., from fig. 36). elements and exploit the correspondence We have learned that –for the rotational features– cal- culations in the restricted (fp)n spaces account remark- l j = l +1/2 m m +1/2 sign(m). n −→ −→ × ably well for the results in the full major shell (pf) (see which is one-to-one except for m = 0. The resulting last column of Table VIII). Let us move now to larger “quasi SU(3)” quadrupole operator respects SU(3) rela- spaces. In Fig. 38 we have yrast transition energies for tionships, except for m = 0, where the correspondence different configurations of 8 particles in ∆j = 2 spaces. The force is KLS, ~ω = 9 MeV, the single particle split- breaks down. The resulting spectrum for quasi-2q20 is shown in the right panel of Fig 36. The result is not tings uniform at ε = 1 MeV, and gds, say, is the lower exact for the K = 1/2 orbits but a very good approxi- sequence in the sdg, p = 4, shell. Rotational behav- mation. ior is fair to excellent at low J. As expected from the To check the validity of the decoupling, a Hartree cal- normalization property of the realistic quadrupole force culation was done for H = εH + H , where H is [Eq. (35)] the moments of inertia in the rotational region sp q sp 2 ′ 2 i.e. the observed single particle spectrum in 41Ca (essentially go as (p +3/2) (p +3/2) , if we multiply all the Eγ values by this factor the lines become parallel. The equidistant orbits with 2MeV spacings) and Hq is the quadrupole force in Eq. (33) with a properly renormal- intrinsic quadrupole moment Q0 [Eq. (86)] remains con- ized coupling. The result is exactly a Nilsson (1955) cal- stant to within 5% up to a critical J value at which the culation (Mart´ınez-Pinedo et al., 1997), bands backbend. Why and how do the bands backbend? We have no simple answer, but Fig. 39 from (Vel´azquez and Zuker, δ 2002) shows the behavior of the (gds)8 space under the H = ~ω εH 2q , (90) Nilsson sp − 3 20 influence of a symmetric random interaction, gradually   38

0.8 E(J+2)-E(J) Positive fp8 12 0.7 fp4.gds4 MeV Negative gds8 gds4.hfp4 0.6 s 9 d p 0.5 g f i h 0.4 6 r4r 5 0.3

0.2 3 50 82 0.1 J 0 g[[]]h 0 0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 132 FIG. 40 Schematic single particle spectrum above Sn. rp FIG. 38 Yrast transition energies Eγ = E(J + 2) − E(J) for is the set of orbits in shell p excluding the largest. For the different configurations, KLS interaction. upper shells the label l is used for j = l + 1/2

1.4 KLS a=-0.1 1.2 a=-0.2 Raju et al. (1973), see also Vargas et al. (2002b) for a=-0.3 more recent applications) will be favored (in which case a=-0.4 1 we have to use the left panel of Fig. 36, with pseudo- p = p 1). Other variants of SU(3) may be possible and 0.8 are well− worth exploring. In cases of truly large deforma- tion SU(3) itself may be valid in some blocks. 0.6

E(J+2)-E(J) To see how this works, consider Fig. 40 giving a 0.4 schematic view of the single particle energies in the space of two contiguous major shells—in protons (π) and neu- 0.2 trons (ν)—adequate for a SM description of the rare earth region. 0 0 5 10 15 20 25 We want to estimate the quadrupole moments for nu- J clei at the onset of deformation. We shall assume quasi- FIG. 39 Backbending patterns in (gds)8 T = 0, with the KLS SU(3) operates in the upper shells, and pseudo-SU(3) in interaction and with a random interaction plus a constant the lower ones. The number of particles in each shell JT Wrstu = −a. for which the energy will be lowest will depend on a bal- ance of monopole and quadrupole effects, but Nilsson diagrams suggest that when nuclei acquire stable defor- made more attractive by an amount a. There is no need mation, two orbits K=1/2 and 3/2—originating in the to stress the similarity with Fig. 38. The appearance of upper shells of Fig. 36—become occupied, i.e., the up- backbending rotors seems to be a general result of the per blocks are precisely the 8-particle configurations we competition between deformation and alignment, char- have studied at length. Their contribution to the electric acteristic of nuclear processes. quadrupole moment is then The group theoretical aspects of quasi-SU3 have been recently discussed and applied to the description of the Q0 = 8[eπ(pπ 1)+ eν (pν 1)], (92) sd-shell nuclei by Vargas, Hirsch, and Draayer (2001, − − 2002a). with pπ = 5, pν = 6; eπ and eν are the effective charges. Consider even-even nuclei with Z=60-66 and N=92- 98, corresponding to 6 to 10 protons with pseudo-p = 3, C. Heavier nuclei: quasi+pseudo SU(3) and 6 to 10 neutrons with pseudo-p = 4 in the lower shells. From the left part of Fig. 36 we obtain easily We have seen that quasi-SU(3) is a variant of SU(3) their contribution to Q0, which added to that of Eq. (92) that obtains for moderate spin-orbit splittings. For yields a total other forms of single particle spacings, the pseudo-SU(3) scheme (Arima et al. (1969); Hecht and Adler (1969); Q0 = 56eπ +(76+4n)eν , (93) 39

20 2 2 TABLE IX B(E2) ↑ in e b compared with experiment (Ra- 18 Expt. man et al., 1989) 16 4p−4h SDPF N Nd Sm Gd Dy 14 92 4.47 4.51 4.55 4.58 12 2.6(7) 4.36(5) 4.64(5) 4.66(5) J 10 94 4.68 4.72 4.76 4.80 8 5.02(5) 5.06(4) 6 96 4.90 4.95 4.99 5.03 4 5.25(6) 5.28(15) 2 0 98 5.13 5.18 5.22 5.26 0 1000 2000 3000 4000 5000 5.60(5) Eγ (keV)

FIG. 41 The superdeformed band of 36Ar, experiment vs. SM for 152+2nNd, 154+2nSm, 156+2nGd and 158+2nDy re- spectively. At fixed n, the value is constant in the By applying the quasi+pseudo-SU(3) recipes of Sec- four cases because the orbits of the triplet K=1/2, tion VI.C we find that the maximum deformations at- 3/2, 5/2 in Fig. 36 have zero contribution for p=3. tainable are of 8p-8h character in 40Ca and 4p-4h in 36Ar Q (given in dimensionless oscillator coordinates, i.e., 0 with Q = 180 efm2 and Q = 136 efm2 respectively. r r/b with b2 1.01A−1/3fm2), is related to the 0 0 The natural generalization of the ZBM space consists E2→ transition probability≈ from the ground state by in the d s fp orbits (remember: fp f p ), B(E2) = 10−5A2/3Q2. The results, using effective 3/2 1/2 7/2 3/2 0 which is relatively free of center of mass spuriousness.≡ charges↑ of e = 1.4, e = 0.6 calculated in (Dufour and π ν And indeed, it describes well the rotational regime of the Zuker, 1996) are compared in table IX with the available observed bands. However, to track them beyond back- experimental values . The agreement is quite remarkable bend, it is convenient to increase the space to d s pf. and no free parameters are involved. Note in particu- 3/2 1/2 The adopted interaction, sdfp.sm is the one originally lar the quality of the prediction of constancy (or rather constructed by Retamosa et al. (1997), and used in (Cau- A2/3 dependence) at fixed n, which does not depend on rier et al., 1998) (USD, KB3 for intra shell and KLS for the choice of effective charges. The discrepancy in 152Nd cross shell matrix elements), with minor monopole ad- is likely to be of experimental origin, since systematics justments dictated by new data on the single particle indicate, with no exception, much larger rates for a 2+ structure of 35Si (Nummela et al., 2001a). state at such low energy (72.6 keV). It is seen that by careful analysis of exact results one may come to very The calculations are conducted in spaces of fixed num- ber of particles and holes. In figure 41 the calculated simple computational strategies. In the last example on 36 B(E2) rates, the simplicity is such that the computation energy levels in Ar are compared to the data. The reduces to a couple of sums. agreement is excellent, except at J = 12 where the data show a clear backbending while the calculation produces a much smoother upbending pattern. In table X the calculated spectroscopic quadrupole 36 40 D. The Ar and Ca super-deformed bands moments (Qs) and the B(E2)’s are used to compute the intrinsic Q as in Eqs. (86,87), using standard ef- The arguments sketched above apply to the region fective charges δqπ=δqν =0.5. As expected both Q0(s) 16 around O, where a famous 4-particles 4-holes (4p-4h) and Q0(t) are nearly equal and constant—and close to band starting at 6.05 MeV was identified by Carter the quasi+pseudo SU(3) estimate—up to the backbend. et al. (1964), followed by the 8p-8h band starting at The calculated B(E2)’s agree well with the experimental 16.75 MeV (Chevallier et al., 1967). Shell model cal- ones (Svensson et al., 2001). The value of Q0 corresponds culations in a very small space, p1/2d5/2s1/2, could ac- to a deformation β 0.5. count for the spectroscopy in 16O, including the 4p-4h Now we examine≈ the 8p-8h band in 40Ca (Ideguchi band (Zuker et al., 1968, ZBM), but the 8p-8h one needs et al., 2001). The valence space adopted for 36Ar is trun- at least three major shells and was tackled by an α-cluster cated by limiting the maximum number of particles in model (Abgrall et al., 1967a,b). It is probably the first the 1f5/2 and 2p1/2 orbits to two. superdeformed band detected and explained. The experimental (Ideguchi et al., 2001) and calcu- In 40Ca, the first excited 0+ state is the 4p-4h band- lated yrast gamma-ray energies are compared in fig. 42. head. It is only recently that another low-lying highly The patterns agree reasonably well but the change of deformed band has been found (Ideguchi et al., 2001), fol- slope at J = 10 —where the backbend in 48Cr starts lowing the discovery of a similar structure in 36Ar (Svens- (Fig. 33)—is missed by the calculation, which only back- son et al., 2000). bends at J = 20, the band termination for the configura- 40

TABLE X Quadrupole properties of the 4p-4h configuration’s TABLE XI Quadrupole properties of the 8p-8h configura- yrast-band in 36Ar (in e2fm4 and efm2) tion’s yrast-band in 40Ca (in e2fm4 and efm2), calculated in the sdpf valence space B(E2)(J→J-2) J B(E2)(J → J-2) Qspec Q0(t) Q0(s) J EXP TH Qspec Q0(s) Q0(t) 2 589 -49.3 172 172 2 315 -36.0 126 126 4 819 -62.4 170 172 4 372(59) 435 -45.9 126 124 6 869 -68.2 167 171 6 454(67) 453 -50.7 127 120 8 860 -70.9 162 168 8 440(70) 429 -52.8 125 114 10 823 -71.6 157 164 10 316(72) 366 -52.7 121 104 12 760 -71.3 160 160 12 275(72) 315 -53.0 119 96 14 677 -71.1 149 157 14 232(53) 235 -54.3 120 82 16 572 -72.2 128 158 16 >84 131 -56.0 122 61 18 432 -75.0 111 162 20 72 -85.1 26 24 Expt. 22 8 -79.1 22 8p−8h, SDPF 24 7 -81.5 20 18 16 14 field methods using the Skyrme interaction have been re- J 12 cently applied to both the 36Ar and 40Ca superdeformed 10 bands by Bender, Flocard, and Heenen (2003a). 8 6 4 2 E. Rotational bands of unnatural parity 0 0 1000 2000 3000 4000 5000 The occurrence of low-lying bands of opposite parity to Eγ (keV) the ground state band is very frequent in pf shell nuclei. FIG. 42 The superdeformed band in 40Ca; exp. vs. 8p-8h Their simplest characterization is as particle-hole bands calculation with the hole in the 1d3/2 orbit. In a nucleus whose ground state is described by the configurations (pf)n the −1 n+1 opposite parity intruders will be (1d3/2) (pf) . The 8 −8 tion f7/2(d3/2s1/2) . The extra collectivity induced by promotion of a particle from the sd to the pf shell costs the presence of sd particles in pseudo-SU(3) orbitals is an energy equivalent to the local value of the gap, that in responsible or the delay in the alignment, but apparently this region is about 7 MeV. On the other side, the pres- too strong to allow for the change in slope at J = 10. ence of one extra particle in the pf-shell may produce +39 an important increase of the correlation energy, that can The experimental Q0(t)=180−29 obtained from the fractional Doppler shifts corresponds to a deformation compensate the energy lost by the particle hole jump. β 0.6 (Ideguchi et al., 2001). It is extracted from an For instance, the very low-lying positive parity band of ≈ 47 −1 overall fit that assumes constancy for all measured values, V, can be interpreted as (1d3/2) (a proton hole) cou- and corresponds exactly to the quasi+pseudo SU(3) es- pled to (pf 8) T=0. Indeed, the correlation energy of this timate. It also squares well with the calculated 172 efm2 pseudo-48Cr is larger than the correlation energy of the in Table XI where the steady decrease in collectivity re- ground state of 47V and even larger than the correlation mains consistent with experiment within the quoted un- energy of the real 48Cr, explaining why the band starts certainties. A reanalysis of the experimental lifetimes in at only 260 keV of excitation energy (Poves and S´anchez the superdeformed band of 40Ca (Chiara et al., 2003) sug- Solano, 1998). The most extreme case is 45Sc, where the −1 gest that for low spins the deformation would be smaller intruder band based in the configuration (1d3/2) cou- due to the mixing with less deformed states of lower np- pled to 46Ti, barely miss (just by 12 keV) to become the nh rank. ground state. Many bands of this type have been ex- This calculation demonstrates that a detailed descrip- perimentally studied in recent years, mainly at the Gasp tion of very deformed bands is within reach of the shell and Euroball detectors (Brandolini et al., 1999), and ex- model. The next step consists in remembering that states plained by shell model calculations. at fixed number of particles are doorways that will frag- We present now some recent results in 48V (Brandolini ment in an exact calculation that remains to be done. et al., 2002a). The positive parity levels of this odd- The band in 36Ar has been also described by the pro- odd nucleus are very well reproduced by the shell model jected shell model in (Long and Sun, 2001). Beyond mean calculation using the interactions KB3 (or KB3G) as can 41

FIG. 43 The ground state (K=4+) and the side band (K=1+) of 48V, experiment vs. theory.

FIG. 44 The negative parity K=1− and K=4− bands of 48V, be seen in Fig. 43, another example of good quality “pure experiment vs. theory. spectroscopy”. The levels are grouped in two bands, one connected to the 4+ ground state, that would correspond to K=4+ in a Nilsson context (the aligned coupling of the last unpaired proton K=3/2− and the last neutron K=5/2−) and a second one linked to the 1+ member of the ground state multiplet that results of the anti-aligned coupling. The negative parity states can be interpreted as the 49 result of the coupling of a hole in the 1d3/2 orbit to Cr. As we have mentioned earlier, the ground state band of 49Cr can be understood in a particle plus rotor picture, − and assigned K=5/2 . Now we have to couple the 1d3/2 hole to our basic rotor 48Cr. Therefore the structure of the negative parity states of 48V is one particle and one hole coupled to the rotor core of 48Cr. Hence we expect that the lowest lying bands would have K=1− and K=4− (aligned and anti-aligned coupling of the quasiparticle and the quasihole). These two bands have been found FIG. 45 Theoretical and experimental B(M1) and B(E2) experimentally at about 500 keV of excitation energy. transition probabilities in the ground state band of 48V Their structure is in very good agreement with the SM predictions (see Fig. 44) except for a small shift of the K=4− and K=8− bandheads relative to the K=1− state. Fig. 45. Notice that the calculation reproduces even the A more stringent test of the theoretical predictions is tiniest details of the experimental results, as, for instance provided by the electromagnetic properties. In this par- the odd-even staggering of the B(M1)’s. The B(E2)’s cor- ticular nucleus the experimental information is very rich respond to a deformation β=0.2 that is clearly smaller and a detailed comparison can be made for both the pos- than that of 48Cr. itive and negative bands and for E2 and M1 transitions. The transition probabilities in the negative parity The results for the ground state band are collected in bands are presented in Fig. 46. Now the experimen- 42

this monopole drift provides direct evidence for the need of three-body mechanisms. For the p shell, the situation is similar: The imposing 16 11.5 MeV p1/2-d5/2 gap in O is down to some 3 MeV 12 in C. The monopole drift of the s1/2-d5/2 gap brings it from about 8 MeV in 28Si to nearly -1 MeV in 12C. According to Otsuka et al. (2001a) the drift may well 8 continue: in He, the s1/2-p1/2 bare gap is estimated at 0.8 MeV. All the numbers above (except the last) are experi- mental. The bare monopole values are smaller because correlations increase substantially the value of the gaps, but they do not change qualitatively the strong monopole drifts. Their main consequence is that “normal” states, i.e., those described by 0~ω p or sd calculations often “coexist” with “intruders” that involve promotion to the next oscillator shell. Let us examine how this happens.

FIG. 46 Theoretical and experimental B(M1) and B(E2) 48 transition probabilities in the negative parity bands of V A. N=8; 11Li: halos

The 1/2+ ground state in 11Be provided one of the first ~ tal results have larger uncertainties. Despite that, the examples of intrusion. The expected 0 ω normal state agreement in absolute values and trends is very good for lies 300 keV higher. The explanation of this behavior the lowest band (K=1−). The quantitative agreement is has varied with time (Aumann et al., 2001; Sagawa et al., worse for the other two bands, nevertheless, the main fea- 1993; Suzuki and Otsuka, 1994; Talmi and Unna, 1960), see also (Brown, 2001) for a recent review and (Suzuki tures of the data are well reproduced. The deformation ~ of the K=1− band extracted from the B(E2) properties et al., 2003) for a new multi- ω calculation, but the idea turns out to be larger than the deformation of the ground has remained unchanged. The normal state corresponds state band and close to that of 48Cr. to a hole on the N = 8 closure. The monopole loss of promoting a particle to the sd shell is compensated by a −2 pairing gain for the p1/2 holes. A quadrupole gain due to VII. DESCRIPTION OF VERY NEUTRON RICH NUCLEI the interaction of the sd neutron with the p2 protons is plausible, but becomes questionable when we note that The study of nuclei lying far from the valley of stabil- the same phenomenon occurs in 9He which has no p2 ity is one of the most active fields in today’s experimen- particles, suggesting the need for a further reduction of tal nuclear physics. What it specific about these nuclei, the monopole loss (Otsuka et al., 2001a). from a shell model point of view? As everywhere else: The interest in 11Be—as a “halo” nucleus—was revived the model space and the monopole behavior of the in- by the discovery of the remarkable properties of 11Li, teraction. They go together because the effective single which sits at the drip line ( 200 keV two-neutron sepa- particle energies (ESPE, see Section V) depend on occu- ration energy) and has a very≈ large spatial extension, due pancies, which in turn depend on the model space. to a neutron halo (Hansen and Jonson, 1987; Tanihata In broad terms, the specificity of light and medium et al., 1985). neutron rich nuclei is that the EI closures (correspond- The shell model has no particular problem with halo ing to the filling of the p +1/2 orbit of each oscillator nuclei, whose large size is readily attributed to the large shell) take over as boundaries of the model spaces. As size of the s1/2 orbit. As shown by (Kahana et al., 1969a) discussed in Section II.B.2, the oscillator closures (HO) the use of Woods Saxon wavefunctions affect the matrix may be quite solid for double magic nuclei, but they be- elements involving this orbit, but the uncertainties in- come vulnerable in the semi-magic cases. volved are easily absorbed by the monopole field. As a For instance the d -f neutron gap in 40Ca of consequence, the whole issue hinges on the s contribu- 3/2 7/2 ≈ 1/2 7 MeV goes down to 2.5 MeV around 28O. Since the tion to the wavefunctions, which is sensitively detected ≈ − 11 d5/2 orbit is well below its sd partners, now quite close by the β decay to the first excited 1/2 state in Be. In to f7/2, the natural model space is no longer the sd shell, Borge et al. (1997) it was shown that calculations produc- but the EI space bounded by the N = 14 and 28 closures, ing a 50% split between the neutron closed shell and the 2 −2 supplemented by the p3/2 subshell whenever the p3/2-f7/2 s1/2 p1/2 configuration lead to the right lifetime. This gap becomes small: from 6 MeV in 56Ni, it drops to result was confirmed by Simon et al. (1999). Further 4.5 MeV in 48Ca, then 2≈ MeV in 40Ca, and finally ≈0 confirmation comes from Navin et al. (2000), with solid MeV in 28Si. As explained≈ at the end of Section II.B.2≈ indications that the supposedly semi-magic 12Be ground 43

2 −2 10 state is dominated by the same s1/2 p1/2 configuration. 2p3/2 5 1f7/2 1d3/2 32 0 2s1/2 B. N=20; Mg, deformed intruders 1d5/2 -5 In the mid 1970’s it was the sd shell that attracted the 11 most attention. Nobody seemed to remember Be, and -10 everybody (including the authors of this review active at the time) were enormously surprised when a classic -15 experiment Thibault et al. (1975) established that the mass and β-decay properties of the 31Na ground state— -20 expected to be semimagic at N = 20—could not possi- -25 28 30 32 34 36 38 40 bly be that of a normal state (Wildenthal and Chung, A 1979). The next example of a frustrated semi-magic was 10 32 et al. Mg (Detraz , 1979). Early mean field calcula- 2p3/2 5 tions had interpreted the discrepancies as due to defor- 1f7/2 mation (Campi et al., 1975) but the experimental confir- 2d3/2 0 2s1/2 mation took some time (Guillemaud-Mueller et al., 1984; 2d5/2 Klotz et al., 1993; Motobayashi et al., 1995). -5 Exploratory shell model calculations by Storm et al. -10 (1983), including the 1f7/2 orbit in the valence space, was able to improve the mass predictions, however, de- formation was still absent. To obtain deformed solutions -15 demanded the inclusion of the 2p3/2 orbit as demon- -20 strated by Poves and Retamosa (1987). These calcu- lations were followed by many others (Fukunishi et al., -25 28 30 32 34 36 38 40 1992; Heyde and Woods, 1991; Otsuka and Fukunishi, A 1996; Poves and Retamosa, 1994; Siiskonen et al., 1999; Warburton et al., 1990), that mapped an “island of inver- FIG. 47 Effective single particle energies (in MeV) at N=20 with the sdpf.sm interaction (upper panel) and the Tokyo sion”, i.e. the region where the intruder configurations group interaction (bottom panel) are dominant in the ground states. The detailed con- tour of this “island” depends strongly on the behavior of the effective single particle energies (ESPE), in turn 6 O dictated by the monopole hamiltonian. Let’s recall, that F according to what we have learned in Section VI.B, the Ne 2−4 2−4 4 Na configurations (d5/2s1/2)p (f7/2p3/2)n ,corresponding to N = 20,Z = 10-12, have a “quasi-SU(3)” quadrupole Mg coherence close to that of SU(3) (i.e., maximal). 2 The ESPE in Fig. 47 represent Hm for the sdfp.sm and Tokyo group interactions (Utsuno et al., 1999). Within details they are quite close, and such as efficiently to 0 favor deformation. Both interactions lead to an “island of inversion” for Z=10, 11 and 12; N=19, 20 and 21. -2 Consider now some detailed information obtained with sdpf.sm by Caurier et al. (2001c). The S2N values of 16 18 20 22 24 26 28 30 Fig. 48 locate the neutron drip line, consistent with what N is known for oxygen and fluorine, where the last bound isotopes are 24O and 31F (Sakurai et al., 1999). Note FIG. 48 Two neutron separation energies (in MeV) calculated the kink due to deformed correlations in the latter. For with the sdpf.sm interaction. the other chains, the behavior is smoother and the last predicted bound isotopes are 34Ne, 37Na and 40Mg. Some results for the even Mg isotopes (N=18, N=20 prefer the intruder. A preliminary measure of the 4+ and N=22) are gathered in table XII. In 30Mg the normal excitation energy reported by Azaiez (1999) goes in the configuration is the one that agrees with the existing ex- same direction. Data and calculations suggest prolate perimental data (Pritychenko et al., 1999). In 32Mg the deformation with β 0.5. In 34Mg the normal config- situation is the opposite as the experimental data (the 2+ uration that contains≈ two pf neutrons, is already quite excitation energy (Guillemaud-Mueller et al., 1984) and collective. It can be seen in the table that it resembles the 0+ 2+ B(E2) (Motobayashi et al., 1995)) clearly the 32Mg ground state. The 4p-2h intruder is even more → 44

TABLE XII Properties of the even magnesium isotopes. N stands for normal and I for intruder. Energies in MeV, B(E2)’s in e2fm4 and Q’s in efm2

30Mg 32Mg 34Mg N I EXP N I EXP N I EXP + ∆E(0I ) +3.1 -1.4 +1.1 0+ 0.0 0.0 0.0 0.0 0.0 0.0 2+ 1.69 0.88 1.48 1.69 0.93 0.89 1.09 0.66 0.67 4+ 4.01 2.27 2.93 2.33 (2.29) 2.41 1.86 2.13 6+ 6.82 3.75 9.98 3.81 3.52 3.50 B(E2) 2+ → 0+ 53 112 59(5) 36 98 90(16) 75 131 126(25) 4+ → 2+ 35 144 17 123 88 175 6+ → 4+ 23 140 2 115 76 176 + Qspec(2 ) -12.4 -19.9 -11.4 -18.1 -15.4 -22.7 deformed (β 0.6) and a better rotor. Results from the tent. This can be seen in Fig. 49, where we have plotted Riken experiments≈ of Yoneda et al. (2000) and Iwasaki the strength functions of the 2p-2h 0+ states in the full et al. (2001) seem to favor the intruder option. In the space. In 40Mg the 2p-2h state represents the 60% of the QMCD calculations of Utsuno et al. (1999), the ground ground state while in 52Cr it is the dominant component state band is dominantly 4p-2h; the 2+ comes at the right (70%) of the first excited 0+. This is a very interesting place, but the 4+ is too high. Clearly, 34Mg is at the edge illustration of the mechanism of intrusion; the intruder of the “island of inversion”. Another manifestation of the state is present in both nuclei, but it is only in the very intruder presence in the region has been found at Isolde neutron rich one that it becomes the ground state. (Nummela et al., 2001b): The decay of 33Na indicates that the ground state of 33Mg has Jπ=3/2+ instead of the expected Jπ=3/2− or Jπ=7/2−. This inversion is nicely reproduced by the sdpf.sm calculation.

C. N=28: Vulnerability

Let us return for a while on Fig. 47. The scale does not do justice to a fundamental feature: the drift of the p3/2 f7/2 gap, which decreases as protons are removed. As explained− at the end of Section II.B.2, this behavior is contrary to a very general trend in heavier nuclei, and demands a three-body mechanism to resolve the contra- diction. Next we remember that the oscillator closures are quite vulnerable even at the strict monopole level. As we have seen, quadrupole coherence takes full advantage of this vulnerability. On the contrary, the EI closures are very robust. However, because of the drift of the p3/2 f7/2 gap even the N = 28 closure becomes vul- nerable.− The sdpf.sm interaction leads to a remarkable result summarized in Table XIII: the decrease of the gap combined with the gain of correlation energy of the 2p-2h lead to a breakdown of the N = 28 closure for 44S and 40Mg. The double magic 42Si resists. Towards N=Z, 52Cr and 54Fe exhibit large correlation energies associated to the prolate deformed character of their 2p-2h neutron configurations (β 0.3). FIG. 49 LSF of the 2p-2h 0+ bandheads of 40Mg (upper ∼ 52 When the 2p-2h bandheads of 40Mg and 52Cr are al- panel) and Cr (bottom panel) in the full 0~ω space lowed to mix in the full 0~ω space, using them as pivots in the LSF procedure, they keep their identity to a large ex- Table XIV gives an idea of the properties of the iso- 45

TABLE XIII N=28 isotones: quasiparticle neutron gaps, dif- TABLE XIV N=28 isotones: Spectra, quadrupole properties ference in correlation energies between the 2p-2h and the 0p- and occupancies 0h configurations and their relative position 40Mg 42Si 44S 46Ar 40Mg 42Si 44S 46Ar 48Ca 50Ti 52Cr 54Fe 56Ni E∗(2+)(MeV) 0.81 1.49 1.22 1.51 gap 3.35 3.50 3.23 3.84 4.73 5.33 5.92 6.40 7.12 E∗(4+) 2.17 2.68 2.25 3.46 ∆ECorr 8.45 6.0 6.66 5.98 4.08 7.59 10.34 10.41 6.19 ∗ + E (02 ) 1.83 1.57 1.26 2.93 ∗ + 2 E2p−2h -1.75 1.0 -0.2 1.7 5.38 3.07 1.50 2.39 8.05 Q(2 )(e fm ) -21 16 -17 20 B(E2)(e2 fm4) 108 71 93 93

hn7/2i 5.54 6.16 6.16 6.91 tones in which configuration mixing is appreciable, but no 8 (f7/2) % 3 28 24 45 clearcut deviation from N=28 magicity can be detected. In Fig. 50 we have plotted the low energy spectra of the heaviest known sulphur isotopes. The agreement with TABLE XV 2+ energies and g intruder occupation in the the accumulated experimental results (see Glasmacher, 1 9/2 Nickel isotopic chain, from (Sorlin et al., 2002) 1998, for a recent review) is excellent and extends to the new data of Sohler et al. (2002). Analyzing their proton 62Ni 64Ni 66Ni 68Ni 70Ni 72Ni 74Ni occupancies we conclude that the rise in collectivity along + the chain is correlated with the equal filling of the d3/2 E(2 )calc. 1.11 1.24 1.49 1.73 1.50 1.42 1.33 and s orbitals, (the d orbital remains always nearly 1/2 5/2 E(2+) 1.173 1.346 1.425 2.033 1.259 closed). The maximum proton collectivity is achieved exp. when both orbitals are degenerate, which corresponds to B(E2↑)calc. 775 755 520 265 410 505 690 the Pseudo-SU(3) limit. For the neutrons, the maximum hn9/2i 0.24 0.43 0.67 1.07 0.84 0.55 0.45 collectivity occurs at N = 24, the f7/2 mid-shell. Ac- cording to the calculation, 42S is a prolate rotor, with an incipient γ band. In 44S the spherical and deformed configurations mix equally. space collapses under the invasion of deformed intruders. The N=27 isotones also reflect the regular transition Naturally, we expect something similar in the neighbor- from sphericity to deformation in their low-lying spec- hood of N = 40 isotones. Recent experiments involving trum. The excitation energy of the 3/2− state should the Coulomb excitation of 66−68Ni (Sorlin et al., 2002), be sensitive to the correlations and to the neutron gap. the β decay of 60−63V (Sorlin et al., 2003), and the de- While in 47Ca, it lies quite high (at around 2 MeV) cay and the spectroscopy of the 67mFe (Sawicka et al., 45 due to the strong f7/2 closure, in Ar it appears at 2003), support this hypothesis. In addition, the β decays about 0.4 MeV. Concerning 43S, the information comes of the neutron-rich isotopes 64Mn and 66Mn, investigated from two recent experiments: the mass measurements at at Isolde (Hannawald et al., 1999), that show a sudden Ganil by Sarazin et al. (2000) that observed a low-lying drop of the 2+ energies in the daughter nuclei 64Fe and isomer around 400 keV excitation energy and the MSU 66Fe, point also in the same direction. Coulex experiment of Ibbotson et al. (1999) that detected In defining the new model spaces, the full pf results are a strong E2 transition from the ground state to an excited quite useful: 56Ni turns out to be an extraordinarily good state around 940 keV. According to our calculations the pivot (refer to Fig. 15), and as soon as a few particles ground state corresponds to the deformed configuration are added it becomes a good core. Therefore, for the − − and has spin 3/2 . The spherical single hole state 7/2 first time we are faced with a genuine EI space r3 g9/2 would be the first excited state and his lifetime is con- (remember r3 is the “rest” of the p = 3 shell). Its validity sistent with that of the experimental isomer. The third is restricted to nearly spherical states. Deformation will known state is short lived and it should correspond to demand the addition of the d5/2 subshell. − the 7/2 member of the ground state band. The effective interaction is based on three blocks: i) the TBME from KB3G effective interaction (Poves et al., 2001), ii) the G-matrix of ref. (Hjorth-Jensen et al., 1995) D. N=40; from “magic” 68Ni to deformed 64Cr with the modifications of ref. (Nowacki, 1996) and iii) the Kahana, Lee and Scott G-matrix (Kahana et al., 1969b) Systematics is a most useful guide, but it has to be for the remaining matrix elements. We have computed used with care. The p and sd shells can be said to be all the nickel isotopes and followed their behavior at and “full” 0~ω spaces, in the sense that the region bound- beyond N=40. Let’s first focus on 68Ni. The calculated aries are well defined by the N = 4, 8 and 20 closures, low energy spectrum, an excited 0+ at 1.85 MeV, 2+ at with the exception for the very neutron-rich halo nuclei 2.15 MeV, 5− at 2.72 MeV and 4+ at 3.10 MeV, fit re- and the island of inversion discussed previously. In the markably well with the experimental results. The ground pf-shell it is already at N = Z 36 that the 0~ω model state wave function is 50% closed shell. This number ≈ 46

FIG. 50 Predicted level schemes of the heavy sulphur isotopes, compared with the experiment. Energies in keV, B(E2)’s in e2 fm4 is very sensitive to modifications of the pf-g gap, and 1200 smaller values of the closed shell probability can be ob- tained without altering drastically the rest of the prop- 1000 erties. The SM results have been compared with those of ] other methods in (Langanke et al., 2003b). In table XV 4 800 fm 2

TABLE XVI Spectroscopic properties of 60−64Cr in the fpgd 600 valence space 400 60Cr 62Cr 64Cr B(E2) [e 200 E∗(2+) (MeV) 0.67 0.65 0.51 2 Qs(e.fm ) -23 -27 -31 0 B(E2)↓(e2.fm4) 288 302 318 26 28 30 32 34 36 38 40 42 44 2 Neutron number Qi(e.fm ) from Qs 82 76 109 2 Qi(e.fm ) from B(E2) 101 103 106 FIG. 51 B(E2) (0+ → 2+) in e2 fm4 for the Nickel isotopes, E∗(4+) (MeV) 1.43 1.35 1.15 from (Sorlin et al., 2002) 2 Qs(e.fm ) -37 -30 -43 B(E2)↓(e2.fm4) 426 428 471 2 2000, 2003). If N=40 were a , one would Qi(e.fm ) from Qs 102 84 119 + 2 expect higher 2 energies as the shell closure approaches, Qi(e.fm ) from B(E2) 117 117 123 particularly in 64Cr. As seen in Fig. 52, exactly the oppo- site seems to happen experimentally. The three possible + we compare the excitation energy of the 2 states in the model spaces tell an interesting story: pf r3 is accept- ≡ Nickel chain with the available experimental data (in- able at N = 32, 34. The addition of g9/2 does more harm cluding the very recent 70Ni value (Sorlin et al., 2002)). than good because the gaps have been arbitrarily reduced The agreement is quite good, although the experiment to reproduce the N = 36 experimental point by allowing gives a larger peak at N=40. Similarly the B(E2)’s in 2p-2h jumps including d5/2. But then, in N = 38 the Fig. 51 follow the experimental trends including the drop 2+ is too high, and the experimental trend promises no at 68Ni recently measured (Sorlin et al., 2002). Note that improvement at N = 40. + for a strict closure the transition would vanish. Properties of the 21 energies given by the pfgd calcu- The β decays of 60V and 62V indicate very low en- lation are collected in table XVI. They point to prolate ergies for the 2+ states in 60Cr and 62Cr (Sorlin et al., structures with deformation β 0.3. 1 ≈ 47

2 argument Wuosmaa et al. (1998) estimate the partial β− exp. half-life to be (6.0 1.3[stat] 1.1[theor]) 105. The in- fp ± ± × fpg fluence of this half-life on the age of galactic cosmic-rays fpgd 1.5 has been discussed recently by Yanasak et al. (2001). An- other possible cosmic ray chronometer, 56Ni, could mea- sure the time between production of group nuclei in 1 supernovae and the accelaration of part of this matterial to form cosmic rays (Fisker et al., 1999). Before acceler- E(2+) (Mev) ation, the decay of 56Ni proceeds by to 0.5 the 1+ state in 56Co with a half-live of 6.075(20) days. After acceleration 56Ni is stripped of its , the transition to the 1+ state is no longer energetically al- 0 30 32 34 36 38 40 42 lowed and the decay proceeds to the 3+ state at 158 keV Neutron number via a second forbidden unique transition. Currently only 56 + a lower limit for the half-life of totally ionized Ni could FIG. 52 Experimental and calculated 2 energies in the pf, be established (2.9 104 y) (Sur et al., 1990). A recent pfg and pfgd spaces. shell-model calculation× by (Fisker et al., 1999) predict a half-live of 4 104 y that is too short for 56Ni to serve as a cosmic ray× chronometer. VIII. OTHER REGIONS AND THEMES Nuclear beta-decay and electron capture are important A. Astrophysical applications during the late stages of stellar evolution (see Langanke and Mart´ınez-Pinedo, 2003, for a recent review). At the Astrophysical environments involve conditions of tem- relevant conditions in the star electron capture and β de- perature and densities that are normally not accessible in cay are dominated by Gamow-Teller (and Fermi) transi- laboratory experiments. The description of the different tions. Earlier determinations of the appropriate weak in- nuclear processes requires then theoretical estimates. As teraction rates were based in the phenomenological work discussed in previous sections, shell-model calculations of Fuller, Fowler, and Newman (1980, 1982a,b, 1985). are able to satisfactorily reproduce many experimental The shell model makes it possible to refine these esti- results so that it should be possible to obtain reliable mates. For the sd shell nuclei, important in stellar oxygen predictions for nuclei and/or conditions not yet accessi- and silicon burning, we refer to Oda et al. (1994). More ble experimentally. recently, it has been possible to extend these studies to pf As a first example, consider the decay properties of shell nuclei relevant for the pre-supernova evolution and nuclei totally stripped of the atomic electrons at typi- collapse (Caurier et al., 1999a; Langanke and Mart´ınez- cal cosmic rays energies of 300 MeV/A. A nucleus such Pinedo, 2000; Langanke and Mart´ınez-Pinedo, 2001). as 53Mn, unstable in normal conditions, becomes stable. The astrophysical impact of the shell-model based weak Another isotope, 54Mn, could be used as a chronome- interaction rates have been recently studied by Heger ter to study the propagation of iron group nuclei on et al. (2001a,b) cosmic-rays (Duvernois, 1997) provided its half-life un- The basic ingredient in the calculation of the differ- der cosmic-rays conditions is known. The necessary cal- ent weak interaction rates is the Gamow-Teller strength culation involves decay by second-forbidden unique tran- distribution. The GT+ sector directly determines the sitions to the ground states of 54Cr and 54Fe. Due to electron capture rate and also contributes to the beta- phase space arguments the β− decay to 54Fe is expected decay rate through the thermal population of excited to dominate. In two difficult and elegant experiments states (Fuller et al., 1982a). The GT− strength con- the very small branching ratio for the β+ decay to the tributes to the determination of the β-decay rate. To ground state of 54Cr has been measured: (1.8 0.8) 10−9 be applicable to calculating stellar weak interaction rates by Zaerpoor et al. (1997) and (1.20 0.26) 10± −9 by× Wu- the shell-model calculations should reproduce the avail- ± × osmaa et al. (1998). Taking the weighted mean of this able GT+ (measured by (n,p)-type reactions) and GT− values, (1.26 0.25) 10−9, and knowing that of 54Mn, (measured in (p,n)-type reactions). Figure 53 com- ± × + 8 312.3(4) d, a partial β half-life of (6.8 1.3) 10 yr pares the shell-model GT+ distributions with the pi- is obtained compared with 5.6 108 yr± from× the shell oneering measurements performed at TRIUMF. These model calculation of Mart´ınez-Pinedo× and Vogel (1998), measurements had a typical energy resolution of which yields 5.0 105 yr for the dominant β− branch. 1 MeV. Recently developed techniques, involving (d, 2He)≈ Both theoretical× results are sensitive to uncertainties in charge-exchange reactions at intermediate energies (Rak- the renormalization of the unique second-forbidden oper- ers et al., 2002), have improved the energy resolution ators that should be removed taking the ratio of the β− by an order of magnitude or more. Figure 54 compares + and β half-lives. Multiplying the theoretical ratio by the shell-model GT+ distribution computed using the the experimental value of the β+ half-life yields a value KB3G interaction (Poves et al., 2001) with a recent ex- of (6.0 1.2) 105 yr for the β− branch. Using a similar perimental measurement of the 51V(d, 2He) performed at ± × 48

1.0 0.5 )

+ 0.4 54Fe Exp 51V 0.8 0.4 51 2 51 SM V(d, He) Ti B(GT 0.3 0.6 0.3

0.4 0.2 0.2

0.2 0.1 0.1 0.0 0.0 56Fe 55Mn 0.6 0.4 0.4 0.1 0.2 0.2 GT Strength 0.2

0.0 0.0 ) + 0.3 58Ni 59Co 0.6 SM 1.0 B(GT 0.4 0.4 0 1 2 3 4 5 6 7 0.5 Ex [MeV] 0.2 FIG. 54 Comparison of the shell-model GT+ distribution (lower panel) for 51V with the high resolution (d, 2He) data 0.0 0.0 (from B¨aumer et al., 2003). The shell-model distribution in- −2 0 2 4 6 8 10 12 −2 0 2 4 6 8 10 12 cludes a quenching factor of (0.74)2. E (MeV) E (MeV)

FIG. 53 Comparison of shell model GT+ distributions with experimental data (Alford et al., 1993; El-Kateb et al., 1994; ization calculations due to the enormous dimensions of R¨onnqvist et al., 1993) for selected nuclei. The shell model the valence space. However, this dimensionality prob- results (discrete lines) have been folded with the experimen- lem does not apply to Shell-Model Monte-Carlo meth- tal resolution (histograms) and include a quenching factor of 2 ods (SMMC, see section IV.B.3). Moreover, the high (0.74) (adapted from Caurier et al., 1999a). temperatures present in the astrophysical environment makes necessary a finite temperature treatment of the nucleus, this makes SMMC methods the natural choice KVI (B¨aumer et al., 2003). for this type of calculations. Initial studies by Langanke Figure 55 compares the shell-model GT− distributions et al. (2001) showed that the combined effect of nuclear with the data obtained in (p,n) charge-exchange reaction correlations and finite temperature was rather efficient measurements for 54,56Fe and 58,60Ni (Anderson et al., in unblocking Gamow-Teller transitions on neutron rich 1990; Rapaport et al., 1983). The GT− operator act- germanium isotopes. More recently this calculations have ing on a nucleus with neutron excess and ground state been extended to cover all the relevant nuclei in the range isospin T can lead to states in the daughter nucleus with A = 65–112 by Langanke et al. (2003a). The resulting three different isospin values (T 1,T,T +1). As a conse- electron-capture rates have a very strong influence in the − quence, the GT− strength distributions have significantly collapse (Langanke et al., 2003a) and post-bounce (Hix more structure and extend over a larger excitation energy et al., 2003) interval than the GT+ distributions, making their theo- The astrophysical r-process is responsible for the syn- retical reproduction more challenging. Nevertheless, the thesis of at least half of the elements heavier than A agreement with the experimental data is quite satisfac- 60 (Wallerstein et al., 1997). Simulations of the r-process≈ tory. The shell-model results for 58Ni have been recently require the knowledge of nuclear properties far from the compared with high-resolution data (50 keV) obtained (Kratz et al., 1998; Pfeiffer et al., 2001). using the (3He,t) reaction (Fujita et al., 2002). As the relevant nuclei are not experimentally accessible, Shell-model diagonalization techniques have been used theoretical predictions for the relevant quantities (i.e., to determine astrophysically relevant weak interaction neutron separation energies and half-lives) are needed. rates for nuclei with A 65. Nuclei with higher masses The calculation of β decay half-lives usually requires two are relevant to study the≤ collapse phase of core-collapse ingredients: the Gamow-Teller strength distribution in supernovae (Langanke and Mart´ınez-Pinedo, 2003). The the daughter nucleus and the relative energy scale be- calculation of the relevant electron-capture rates is cur- tween parent and daughter (i.e. the Qβ value). Due rently beyond the possibilities of shell-model diagonal- to the huge number of nuclei relevant for the r pro- 49

10.0 8.0 54 IAS 56 measurements (K¨appeler et al., 1998). Most of the re- 8.0 Fe Fe IAS 6.0 action rate tables are therefore based on global model 6.0 4.0 predictions. The most frequently used model is the sta- 4.0 tistical Hauser Feshbach approach (Rauscher and Thiele-

(mb/(sr MeV)) 2.0 σ 2.0 mann, 2000). For nuclei near the drip-lines or near closed 0.0 0.0 shell configurations, the density of levels is not high enough for the Hauser Feshbach approach to be appli- 3.0 1.0 cable. For these cases alternative theoretical approaches 2.0

GT such as the nuclear shell model need to be applied. Shell- 0.5 1.0 model calculations were used for the determination of the relevant reaction rates for sd-shell nu- 0.0 0.0 0 5 10 15 0 5 10 15 clei necessary for rp process studies (Herndl et al., 1995). E (MeV) E (MeV) This calculations have been recently extended to include 10.0 6.0 pf-shell nuclei (Fisker et al., 2001). 58Ni IAS 60Ni 8.0 Knowledge of neutrino nucleus reactions is necessary 4.0 6.0 IAS for many applications, e.g. the neutrino oscillation stud- 4.0 ies, detection of supernova neutrinos, description of the 2.0 (mb/(sr MeV))

σ 2.0 neutrino transport in supernovae and

0.0 0.0 studies. Most of the relevant neutrino reactions have

2.0 not been studied experimentally so far and their cross 1.5 sections are typically based on nuclear theory (see Kolbe 1.0 et al., 2003, for a recent review). The model of choice GT 1.0 for the theoretical description of neutrino reactions de- 0.5 pends of the energy of the neutrinos that participate in 0.0 0.0 0 5 10 15 0 5 10 15 the reaction. E (MeV) E (MeV) For low neutrino energies, comparable to the nuclear excitation energy, neutrino-nucleus reactions are very FIG. 55 Comparison of (p, n) L = 0 forward angle cross sec- sensitive to the appropriate description of the nuclear tion data (Rapaport et al., 1983) (upper panels) with the cal- 54 response that is very sensitive to correlations among nu- culated GT− strength distributions (lower panels). For Fe the solid curve in the lower panel shows the experimental cleons. The model of choice is then the nuclear shell- model. 0~ω calculations have been used for the calcula- GT− data from (Anderson et al., 1990), while the shell model results are given by the dashed line. The Fermi transition to tion of neutrino absorption cross sections (Sampaio et al., the isobaric analog state (IAS) is included in the data (upper 2001) and scattering cross sections (Sampaio et al., 2002) panels) but not in the calculations. A quenching factor of for selected pf shell nuclei relevant for supernovae evo- (0.74)2 is included in the calculations. (from Caurier et al., lution. For lighter nuclei complete diagonalizations can 1999a). be performed in larger model spaces, e.g. 4~ω calcula- tions for 16O (Haxton, 1987; Haxton and Johnson, 1990) and 6~ω calculations for 12C (Hayes and Towner, 2000; cess, the estimates of the half-lives are so far based Volpe et al., 2000). Other examples of shell-model cal- on a combination of global mass models and the quasi culations of neutrino cross sections are the neutrino ab- particle random-phase approximation (see Langanke and sorption cross sections on 40Ar of Ormand et al. (1995) Mart´ınez-Pinedo, 2003, for a description of the different for solar neutrinos (see Bhattacharya et al., 1998, for models). However, recently shell-model calculations have an experimental evaluation of the same cross section), become available for some key nuclei with a magic neu- this cross section have been recently evaluated by (Kolbe tron number N = 50 (Langanke and Mart´ınez-Pinedo, et al., 2003) for supernova neutrinos. And the evaluation 2003), N = 82 (Brown et al., 2003; Mart´ınez-Pinedo and by Haxton (1998) of the solar neutrino absorption cross Langanke, 1999), and N = 126 (Mart´ınez-Pinedo, 2001). section on 71Ga relevant for the GALLEX and SAGE All this calculations suffer from the lack of spectroscopic solar neutrino experiments. information on the regions of interest that is necessary For higher neutrino energies the standard method of to fine tune the effective interactions. This situation is choice is the random phase approximation as the neutrino improving at least for N = 82 thanks to the recent spec- reactions are sensitive mainly to the total strength and troscopic data on 130Cd (Dillmann et al., 2003). energy centroids of the different multipoles contributing rates are the key input data for sim- to the cross section. In some selected cases, the Fermi and ulations of stellar burning processes. Experiment-based Gamow-Teller contribution to the cross section could be reaction rates for the simulation of explosive processes determined from shell-model calculation that is supple- such as novae, supernovae, x-ray bursts, x-ray pulsars, mented by RPA calculations for higher multipoles. This and merging neutron stars are scarce because of the ex- type mix calculation has been carried out for several iron perimental difficulties associated with radioactive beam isotopes (Kolbe et al., 1999; Toivanen et al., 2001) and 50 for 20Ne (Heger et al., 2003). the ground state and occasionally a few excited states of the grand daughter, both even-even. Besides, it is nec- essary to have a good description of the Gamow- Teller B. ββ-decays strength functions of both of them, which implies a de- tailed description of the odd-odd intermediate nucleus. The is the rarest nuclear weak pro- This is why this calculation is a challenge for the nuclear cess. It takes place between two even-even isobars, when models, and why agreement with the experiment in this the decay to the intermediate nucleus is energetically for- channel is taken as a quality factor to be applied to the bidden or hindered by the large spin difference between predictions of the models for the neutrinoless mode. the parent ground state and the available states in the It is also a show-case of the use of the Lanczos strength intermediate nuclei. It comes in three forms: The two- function (LSF) method. It works as follows: Once the neutrino decay ββ2ν : relevant wave functions of parent and grand daughter A A − − are obtained, it is straightforward to build the doorway Z XN Z+2 XN−2 + e1 + e2 +¯ν1 +¯ν2 + + −→ states ~σt− 0initial and ~σt+ Jfinal . In a second step, one of them| is fragmentedi using| LSF,i producing, at it- is just a second order process mediated by the Standard + model weak interaction It conserves the lepton number eration N, N 1 states in the intermediate nucleus, with and has been already observed in a few nuclei. excitation energies Em. Overlapping these vectors with The second mode, the neutrinoless decay ββ : the other doorway, entering the appropriate energy de- 0ν nominators and adding up the N contributions gives an A A − − approximation to the exact value of M 2ν (N=1 is just Z XN Z+2 XN−2 + e + e GT −→ 1 2 the closure approximation). Finally, the number of iter- needs an extension of the standard model of electroweak ations is increased until full convergence is reached. The interactions as it violates lepton number. A third mode, method is very efficient, for instance, in the A=48 case, ββ0ν,χ is also possible 20 iterations suffice largely. The contributions of the dif- ferent intermediate states to the final matrix element are A A − − X X − + e + e + χ plotted in Fig. 56. Z N −→Z+2 N 2 1 2 in some extensions of the standard model and proceeds via emission of a light neutral boson, a Majoron χ. The last two modes, not yet experimentally observed, require massive neutrinos –an issue already settled by the recent measures by Super-Kamiokande (Fukuda et al., 1998), SNO (Ahmad et al., 2002) and KamLAND (Eguchi et al., 2003). Interestingly, the double beta decay without emis- sion of neutrinos would be the only way to sign the Majo- rana character of the neutrino and to distinguish between the different scenarios for the neutrino mass differences. Experimentally, the three modes show different electron energy spectra ((see figure 2 in Zdesenko, 2002)), the ββ2ν and ββ0ν,χ are characterized by a continuous spec- trum ending at the maximum available energy Qββ, while the ββ0ν spectrum consists in a sharp peak at the end of the Qββ spectrum. This should, in principle, make it easier the signature of this mode. In what follows we shall concentrate in the ββ2ν and ββ0ν modes. The theoretical expression of the half-life of the 2ν mode can be written as: FIG. 56 LSF for 48Ca → 48Ti 2ν decay. Each bar corresponds to a contribution to the matrix element. Notice the interfering [T 2ν ]−1 = G M 2ν 2, (94) positive and negative contributions. 1/2 2ν | GT | with In the pf-shell there is only one double beta emit- + σ + + σ + 48 2ν J ~ t− 1m 1m ~ t− 0 ter, namely Ca. For many years, the experimental MGT (J)= h || || ih || || i (95) Em + E0(J) information was limited to a lower limit on the 2ν-ββ m 2ν 19 X half-life T1/2> 3.6 10 yr (Bardin et al., 1970). The ×48 (there is an implicit sum over all the nucleons). G2ν calculation of the Ca half-life was one of the first contains the phase space factors and the axial coupling published results stemming from the full pf-shell cal- 2ν constant gA. The calculation of MGT requires the precise culations using the code ANTOINE. The resulting ma- 2ν + 2ν + knowledge of the ground state of the parent nuclei and trix elements are: MGT (0 )=0.083 and MGT (2 )=0.051. 51

For the ground state to ground state decay 48Ca 48Ti, This was believed to make the results less dependent of −17 −1 → G2ν =1.1 10 yr (Tsuboi et al., 1984). The the nuclear model employed to obtain the wave func- phase space× factor hinders the transition to the 2+ tion. An assumption that has not survived to the actual that represents only about 3% of the total probabil- calculations. Full details of the calculations as well as ity. Putting everything together, and using the Gamow- predictions for other 0ν and 2ν decays in heavier double Teller quenching factor, already discussed, the resulting beta emitters can be found in Retamosa et al. (1995) and half-life is T2ν = 3.7 1019 yr, (Caurier et al., 1990) Caurier et al. (1996) (see also Suhonen and Civitarese, 1/2 × (see also erratum in Caurier et al., 1994). The pre- 1998, for a recent and very comprehensive review of the diction was a success, because a later measure gave nuclear aspects of the double beta decay). The upper 2ν +2.4 19 bounds on the neutrino mass resulting of our SM calcu- T1/2= 4.3−1.1[stat] 1.4[syst] 10 yr (Balysh et al., ± × lations, assuming a reference half-life T0ν 1025 y. are 1996). 1/2 ≥ Among the other ββ emitters in nature (around 30), collected in table XVIII. only a few are potentially interesting for experiment since they have a Q value, sufficiently large ( 2.5MeV ) for ββ ≥ TABLE XVIII 0ν matrix elements and upper bounds on the the 0ν signal not to be drowned in the surrounding natu- 0ν 25 neutrino mass for T ≥ 10 y. hmν i in eV. ral radioactivity. With the exception of 150Nd, all of them 1/2 can be described within a shell model approach. The re- Parent 48Ca 76Ge 82Se 48 76 sults for the 2ν mode of the lightest emitters Ca, Ge 0ν MGT 0.63 1.58 1.97 and 82Se, for which full space calculations are doable, M 0ν -0.09 0.19 -0.22 are gathered in table XVII (Caurier et al., 1996). The F hm i 0.94 1.33 0.49 SMMC method has also been applied to the calculation ν of 2ν double beta decays in (Radha et al., 1996).

For the heavier emitters, some truncation scheme has TABLE XVII Calculated T2ν half-lives for several nuclei and 1/2 to be employed and the seniority truncation seem to be 0+ → 0+ transitions the best, since the dimensions are strongly reduced in Parent 48Ca 76Ge 82Se particular for 0+ states. An interesting feature in the T 2ν th.(y) 3.7 × 1019 2.6 × 1021 3.7 × 1019 calculation of the double beta decay matrix elements is 1/2 76 2ν 19 21 19 shown in Fig. 57 for the Ge case : the convergence of T1/2 exp.(y) 4.3 × 10 1.8 × 10 8.0 × 10 0ν the MGT matrix element is displayed as a function of the truncation of the valence space, either with the seniority v or with the configurations t (t is the maximum number The expression for the neutrinoless beta decay half-life, of particles in the g 9 orbital). The t = 16 and v = 16 in the 0+ 0+ case, can be brought to the following form 2 → values correspond to the full space calculation and are (Doi et al., 1985; Takasugi, 1981): consequently equal. The two truncation schemes show very distinct patterns, with the seniority truncation be- (0ν) + + −1 [T (0 > 0 ] = ing more efficient. Such patterns have been also observed 1/2 − in the Tellurium and Xenon isotopes. This seems to in- dicate that the most favorable nuclei for the theoretical 2 2 2 (0ν) gV (0ν) mν calculation of the 0ν mode would be the spherical emit- G0ν M M h i ters, where seniority is an efficient truncation scheme. GT − g F m  A  !  e  where mν is the effective neutrino mass, G0ν the kine- h i (0ν) (0ν) matic space factor and MGT and MF the following C. Radii isotope shifts in the Calcium isotopes matrix elements (m and n sum over nucleons): We have already seen in section VI.E that deformed M (0ν) = 0+ h(r)(σ~ σ~ )t t 0+ (96) np-nh configurations can appear at very low excitation GT f n · m n m i n,m energy around shell closures and even to become yrast D X in the case of neutron rich nuclei. This situation simply reflects the limitation of the spherical mean field descrip- (0ν) + + tion of the nucleus, and shows that even in magic cases, MF = 0f h(r)tn tm 0i , (97) n,m the correlations produce a sizeable erosion of the Fermi D X surface. This is the case in the Calcium isotopic chain. where, due to the presence of the neutrino propagator, Experiments based on optical isotope shifts or muonic 2 the “neutrino potential” h(r) is introduced. In this case, data, reveal that the nuclear charge radii rc fol- the matrix elements are just expectation values of two low a characteristic parabolic shape with a pronouncedh i body operators, without a sum over intermediate states. odd-even staggering (Fricke et al., 1995; Palmer, 1984). 52

4 functions, can be expressed as: v truncation t truncation 1 δr2(A)= nπ (A)b2 (98) 3 c Z pf where Z=20 for the calcium chain, b is the oscillator pa- rameter, and nπ is extracted from the calculated wave 2 pf functions. The charge radii isotope shifts of the Calcium isotopes relative to 40Ca are shown in Fig. 59, together 1 with the experimental values. The global trends are very well reproduced, although the calculated shifts are a bit smaller than the experimental ones. This is probably due 0 to the limitations in the valence space, that excludes the 0 5 10 15 5 5 1 Truncation 1d 2 , the 1f 2 and the 2p 2 orbits.

FIG. 57 Variation of the hmν i limit as a function of senior- 0.3 ity truncations (circles) and configuration truncations (dia- monds) ) 2

0.2 For the description of the nuclei around N=Z=20, a (40)] (fm 2 3 1 7 3 model space comprising the orbits 1d 2 , 2s 2 , 1f 2 and 2p 2 is a judicious choice (see Caurier et al., 2001a). The inter- 0.1 (A) − r action is the same used to describe the neutron rich nuclei 2 in the sd-pf valence space, called sdpf.sm in section VII, with modified single particle energies to reproduce the 0 spectrum of 29Si. Several important features of the nuclei at the sd/pf interface are reproduced by the calculation, among oth- Isotope shift [r ers, the excitation energies of the intruder 0+ states in −0.1 2 40 42 44 46 48 + 3 A the Calcium isotopes as well as the location of the 2 states in the Scandium isotopes (see Fig. 58), the excita- tion energies of the 2+ and 3− and the B(E2)’s between FIG. 59 Isotopes shifts in Calcium chain: experimental (cir- + + the 21 and the 01 states in the Calcium isotopes. cles) versus shell model calculations (stars)

5

exp D. Shell model calculations in heavier nuclei 4 There are some regions of heavy nuclei in which phys- 0 + in Ca 3 2 ically sound valence spaces can be designed that are at sm the same time tractable. A good example is the space 2 comprising the neutron orbits between N=50 and N=82 for the isotopes (Hjorth-Jensen et al., 1995; Nowacki, sm 1996). When protons are allowed the dimensions grow Energy (MeV) 1 rapidly and the calculations have been limited until now exp to nuclei with few particle or holes on the top of the 0 3/2+ in Sc closed shells (see Covello et al., 1997, for a review of the work of the Napoli group). −1 The SMMC, that can overcome these limitations, has 40 42 44 46 48 been applied in this region to 128Te and 128Xe that are A candidates to γ soft nuclei by Alhassid et al. (1996) and also to several Dysprosium isotopes A=152-162 in FIG. 58 Comparison between calculated and experimental in- the Kumar-Baranger space by White et al. (2000). The truder excitation energies in the Calcium and Scandium iso- QMCD method has been also applied to the study of topic chains the spherical to deformed transition in the even Barium isotopes with A=138-150 by Shimizu et al. (2001). Due to the cross shell pairing interaction, protons and Hints on the location of the hypothetical islands of su- neutrons are lifted from the sd to the fp orbitals. The per heavy elements are usually inspired in mean-field cal- former produce an increase of r2 that, using HO wave culations of the single particle structure. The predictions h c i 53

+ + for nuclei far from stability can be tested also in nuclei E* 20± ± 19 19 19+ 18+ [keV] ± 20 + much closer to stability, where shell model calculations 18± 20 5000 17± 19 19± 18+ are now feasible. In particular, recent systematic mean + + 18 field calculations suggest a substantial gap for Z = 92 16 + ± 16± 218 17 ± 18 9 (16 ) 17± and N = 126 ( U) corresponding to the 1h 2 shell clo- + ± ± 16± 15+ 17± 15± 16± 15 sure. On the other hand, the single quasiparticle energies 17 16± ± 17 215 4000 15 15± extrapolated from lighter N=126 isotones up to Ac, do + ± 12+ 14± 13 + 13 14+ + 14 not support this conclusion. To shed light on these con- (14+) 12 14+ + + 10+ troversial predictions, shell model calculations (as well as 14+ + (12 ) 12 13+ 216 12+ 14+ + experimental spectroscopic studies of Th (Hauschild + 12+ 10 12 12+ 10+ et al., 2001)) were undertaken in the N = 126 isotones up 3000 + 10 218 10 3± to U. Shell model calculations were performed in the ± ± 3 ± ± 11 ± ± 11 9 7 13 3 5 1 11 (11 ) 11 ± 1h 2 , 2f 2 , 1i 2 , 3p 2 , 2f 2 , 3p 2 proton valence space, using 3 the realistic Kuo-Herling interaction (Kuo and Herling, + 8+ 1971) as modified by Brown and Warburton (Warburton + + + 6+ 8 + (8+) 8+ 8 8 (6 ) 6 4+ and Brown, 1991). The calculation reproduces nicely the 2000 + 8+ 8 + 4+ + 6 (4+) + 6 8+ ground state energies, the 2 energy systematics as well + ± 2+ 4+ 4 (3 ) 214 216 2+ as the high-spin trends. The cases of Ra, Th and + + + 2 (2 ) 218U are shown in figure 60. The only deviations between 2 theory and experiment are in the 3− energy and reflect 1000 the particular nature of theses states which are known to be very collective and corresponding to particle hole excitations of the 208Pb core. No shell gap for Z=92, corresponding to the 1h 9 clo- 2 + + + + + 0 0 0 0 0 0 sure, is predicted. On the contrary, the ground state EX SM2 EX SM2 SM2 of the N=126 isotones is characteristic of a superfluid 214 216 218 88Ra126 90Th126 92U126 regime with seniority zero components representing more than 95% of the wave functions for all nuclei. FIG. 60 Experimental vs theoretical spectrum of 214Ra, 216Th, and 218U

E. Random Hamiltonians in Fig. 39). This is very much in line with what was found The study of random Hamiltonians is a vast interdis- by Cortes et al. (1982), and the interesting point is that ciplinary subject that falls outside the scope of this re- the collective ingredient induced by the displacement of view (see Porter, 1965, for a collection of the pioneering the matrix elements is the quadrupole force. papers). Therefore here we shall only give a bibliograph- It is unlikely that we shall learn much more from purely ical guide to work that has recently attracted wide at- random Hamiltonians. However, the interplay of random tention and may have consequences in future shell model and collective interactions may deserve further study. In studies. Johnson et al. (1998, 1999) noticed that random particular: We know that a monopole plus pairing Hamil- interactions had a strong tendency to produce J = 0 tonian is a good approximation. Would it be a good idea ground states. This is an empirical fact that was hith- to replace the rest of the interaction by a random one, erto attributed to the pairing force. Bijker and Frank instead of neglecting it? (2000a,b); Bijker et al. (1999) showed that in an inter- acting boson context this also occurred, and was asso- ciated to the typical forms of collectivity found in the IX. CONCLUSION IBM. For the fermion problem no collectivity occurs for purely random interactions (Horoi et al., 2001). The ori- Nuclei are idiosyncratic, especially the lighter ones, ac- gin of J = 0 ground state-dominance was attributed to cessible to shell model treatment. Energy scales that are “geometric chaoticity” (Mulhall et al., 2000). The ge- well separated in other systems (as vibrational and ro- ometric aspects of the phenomenon were investigated in tational states in molecules) do overlap here, leading to some simple cases by Zhao and Arima (2001); Zhao et al. strong interplay of collective and single particle modes. (2001) and Chau Huu-Tai et al. (2003). Vel´azquez and Nonetheless, secular behavior—both in the masses and Zuker (2002) argued that the general cause of J = 0 the spectra—eventually emerges and the pf shell is the ground state-dominance was to be found in time-reversal boundary region were rapid variation from nucleus to nu- invariance, and showed that when the random matrix ele- cleus is replaced by smoother trends. As a consequence, ments were displaced to have a negative centroid, well de- larger calculations become associated with more trans- veloped rotational motion appeared in the valence spaces parent physics, and give hints on how to extend the shell where the realistic interactions would also produce it (as model philosophy into heavier regions where exact diag- 54 onalizations become prohibitive. In this review we have The one particle creation and anhilation operators not hesitated to advance some ideas on how this could † r+rz be achieved, by suggesting some final solution to the Arrz = a Brrz =˜arrz = ( 1) ar−rz . (A1) rrz − monopole problem, and exploiting the formal properties of the Lanczos construction. can be coupled to quadratics in A and B, The shell model has been craft and science: one in- X† (rs) = (A A )Γ , X (rs) = (B B )Γ , vented model spaces and interactions and forced them ΓΓz r s Γz ΓΓz r s Γz γ γ on the spectra. Sometimes it worked very well. Then Sγz (rt) = (ArBt)γz . (A2) one wondered why such a phenomenology succeeded, to discover that there was not so much phenomenology after † −1/2 † ZΓΓz (rs)=(1+ δrs) XΓΓz (rs) is the normalized pair all. It is to be hoped that this state of affairs will persist. operator and ZΓΓz (rs) its Hermitian conjugate. For reduced matrix elements we use Racah’s definition

Acknowledgments < αα P γ ββ >= (A3) z| γz | z α γβ This review owes much to many colleagues; in the ( 1)α−αz < α P γ β > . first place to Joaqu´ınRetamosa who has been a mem- − αz γz βz k k − ! ber of the Strasbourg-Madrid Shell Model collabora- tion for many years, then to Guy Walter, Mike Bentley, The normal and multipole representations of H are Franco Brandolini, David Dean, Jean Duflo, Marianne obtained through the basic recoupling Dufour, Luis Egido, Jos´eMar´ıaG´omez, Hubert Grawe, Mar´ıaJos´eGarc´ıaBorge, Morten Hjorth-Jensen, Karl- heinz Langanke, Silvia Lenzi, Peter Navr´atil, Peter von Γ 1/2 (X† (rs)X (tu))0 = ( 1)u+t−Γ δ S0 + Neumann-Cosel, Achim Richter, Luis Miguel Robledo, − Γ Γ − − r st ru Dirk Rudolph, Jorge S´anchez Solano, Olivier Sorlin, Carl   r s Γ Svensson, Petr Vogel to name only a few. This work [Γγ]1/2( 1)s+t−γ−Γ (Sγ Sγ )0, (A4) − rt su has been partly supported by the IN2P3-France, CICyT- γ (u t γ) Spain agreements and by a grant of the MCyT-Spain, X BFM2000-030. Many thanks are given to Benoit Speckel whose inverse is for his continuous computational support. γ 1/2 (Sγ Sγ )0 = ( 1)u−t+γ δ S0 (A5) rt su − r st ru− APPENDIX A: Basic definitions and results h i 1/2 s+t−γ−Γ r s Γ † 0 [Γγ] ( 1) (X (rs)XΓ(tu)) . − u t γ Γ p is the principal oscillator quantum number Γ ( ) • X D = (p + 1)(p + 2) is the (single fluid) degeneracy The “normal” representation of is then p V • of shell p = Γ Z† Z = Orbits are called r, s, etc. D =2j +1 V Vrstu rsΓ · tuΓ r r ≤ ≤ • r sX t u,Γ ˆ 1/2 Γ † 0 δrs = δjr js but r = s, and both have the same ξrsξtu[Γ] rstu(XrsΓXtuΓ) , (A6) • parity. 6 − V (rstuX)Γ mr is the number of particles in orbit r, Tr is used where we have used • for both the isospin and the isospin operator. In − neutron-proton (np) scheme, mrx specifies the fluid (1 + δ ) 1/2 if r s, ξ = rs (A7) x. Alternatively we use nr and zr. rs 1/2 ≤ ((1 + δrs) /2 if no restriction, Γ Γ Γ Vrstu or rstu are two body matrix elements. Wrstu • is usedV after the monopole part has been sub- so as to have complete flexibility in the sums. Accord- tracted. ing to (A4), can be transformed into the “multipole” representationV A few equations have to exhibit explicitly angular mo- 1/2 γ γ γ 0 mentum (J), and isospin (T ) conservation. We use Bruce = ξrsξtu [γ] ωrtsu(SrtSsu) French’s notations (French, 1966): Γ stands for JT . V (rstu)γ h Then ( 1)Γ = ( 1)J+T , [Γ] = (2J + 1)(2T + 1), and X − − r j +1/2 ˆ 1/2 0 0 in general F (Γ) = F (J)F (T ). Also( 1) = ( 1) r , +δstδru[s] ωrussSru , (A8) − − [r] = 2(2jr + 1). Expressions carry to neutron-proton i formalism simply by dropping the isospin factor. where (sum only over Pauli-allowed Γ) 55

It follows that the form of must be HmT

γ s+t−γ−Γ r s Γ Γ ωrtsu = ( 1) Wrstu[Γ], (A9) ˆ ˆ − u t γ mT = + (artsuSrtsu brtsuTrtsu)δrtδsu, (B3) Γ ( ) H K − X Xall where the sum is over all possible contributions. This is a r s Γ W Γ = ( 1)s+t−γ−Γ ωγ [γ]. (A10) special Hamiltonian containing only λ = 0 terms. Trans- rstu − rtsu γ ( u t γ ) forming to the normal representation through Eq. (A10) X ω00 = δ δ = J0 = J1 = [rs]−1/2 Eq.(A5) suggests an alternative to (A8) λ0 τ0 ⇒Vrstu Vrstu 01 J0 −1/2 J1 −1/2 ω = δλ0δτ1 = rstu = 3[rs] , rstu = [rs] 1/2 γ ⇒V V − = ξrsξtu[γ] ω (A11) V rstu and therefore (rstuX)γ γ 1/2 † γ γ 0 γ+r−s ˆ 0 Srtsu = ZrsΓ ZtuΓ = (B4a) (SrtSsu) ( 1) δstδruSru , · × − − r Γ  h i  X = Z† Z + Z† Z where each term is associated with a pure two body rsJ0 · tuJ0 rsJ1 · tuJ1 operator. XJ XJ

3 † 1 † APPENDIX B: Full form of H Trtsu = Z ZtuJ0 Z ZtuJ1 (B4b) m − 4 rsJ0 · − 4 rsJ1 · X X If we were interested only in extracting the diagonal and inverting parts of the monopole Hamiltonian in jt scheme, mT , H 1 the solution would consist in calculating traces of , and Z† Z = (S 4T ) (B5a) H rsJ0 · tuJ0 4 rtsu − rtsu showing that they can be written solely in terms of num- J ber and isospin operators. Before we describe the tech- X nique for dealing with the non diagonal parts of mT H 1 i.e. for generalizing to them the notion of centroid, some Z† Z = (3S +4T ). (B5b) remarks on the interest of such an operation may be in rsJ1 · tuJ1 4 rtsu rtsu J order. X The full m contains all that is required for Hartree When j = j = j = j and r = s and t = u, both H r s t u 6 6 Fock (HF) variation, but it goes beyond. Minimizing the Srust, Trust and Srtsu, Trtsu are present. They can be energy with respect to a determinantal state will invari- calculated from (B5a) and (B5b) by exchanging t and u, ably lead to an isospin violation because neutron and proton radii tend to equalize (Duflo and Zuker, 2002), † J † ZrsJ0 ZutJ0 = ( 1) ZrsJ0 ZtuJ0 = which demands different orbits of neutrons and protons. · − − · X 1 X Therefore, to assess accurately the amount of isospin vio- = (Srust 4Trust) (B6a) lation in the presence of isospin-breaking forces we must 4 − ensure its conservation in their absence. More generally, a full diagonalization of m is of great intrinsic interest. † J † H Z ZutJ1 = ( 1) Z ZtuJ1 = Now, the technical details. rsJ1 · − rsJ1 · X X1 Define the generalized number and isospin operators = (3S +4T ) (B6b) 4 rust rust 1 S = δˆ [r]1/2S00 ,T = δˆ [r]1/2S01 (B1) rs rs rs rs 2 rs rs and by combining (B5a), (B5b), (B6a) and (B6b)

J which, for δrs = 1, reduce to Srr = nr,Trr = Tr. By † (1 ( 1) ) definition contains the two body quadratic forms in ZrsJ0 ZtuJ0 ± − = (B7a) Hm · 2 Srs and Trs: XJ 1 S = ζ ζ (S S δ S ), (B2a) ((Srtsu 4Trtsu) (Srust 4Trust)) rtsu rs tu rt su − st ru 8 − ∓ −

3 J † (1 ( 1) ) Trtsu = ζrsζtu(Trt Tsu δstSru), (B2b) Z Z ± − = (B7b) · − 4 rsJ1 · tuJ1 2 XJ which in turn become for δrt = δsu = 1, mrs and Trs in 1 ((3S +4T ) (3S +4T )) Eqs. (10a) and (10b) 8 rtsu rtsu ± rust rust 56

To write we introduce the notations then HmT

Φ(P )=1 (1 δ )(1 δ ), (B8) mT = + (1 Φ(e))(aαSα + bαTα) − − rs − tu H K − ˆ ˆ α Φ(e) (δrs δrs)(δtu δtu), (B9) d Xd e e ≡ − − +Φ(e)(aαSα + bαTα + aαSα¯ + bαTα¯), with (B14a) 1 So for r = s or t = u, Φ(P ) =1 and for jr = js = jt = ju 1 0 aα = (3¯ + ¯ ) and Φ(P )=0, Φ(e) = 1. Then 4 Vα Vα 1 ¯1 ¯0 bα = ( α α) (B14b) mT = + 4 V − V H K 1 T T ρT ρT d ¯+1 ¯−1 ¯+0 ¯−0 δˆ δˆ (1 Φ(e)) Ω + Φ(e) Ω aα = (3 α +3 α + α + α ) rt su − Vrstu rstu Vrstu rstu 8 V V V V r≤s h i 1 Xt≤u ae = (3¯+1 3¯−1 ¯+0 + ¯−0) T,ρ=± α 8 Vα − Vα − Vα Vα † J T d 1 +1 −1 +0 −0 ρ = sign ( 1) , Ωrstu = ZrsJT ZtuJT b = (3¯ +3¯ ¯ ¯ ) − · α 2 Vα Vα − Vα − Vα XJ J e 1 ¯+1 ¯−1 ¯+0 ¯−0 ±T † (1 ( 1) ) bα = (3 α α + α α ) (B14c) Ωrstu = Z ZtuJT ± − . (B10) 2 V − V V − V rsJT · 2 XJ T The values of the generalized centroids rstu and ρT are 1. Separation of Hmnp and Hm0 determined by demanding that V = Vcontain H − HmT HM no contributions with λ = 0. In other words In j formalism mnp is mT under another guise: neu- tron and protonH shells areH differentiated and the oper- JT JT W = ators Trs and Srs are written in terms of four scalars rstu Vrstu− T ρT Srxsy ; x, y = n or p. δˆ δˆ (1 Φ(e)) + Φ(e) , (B11) − rt su − V rstu Vrstu We may also be interested in extracting only the purely h i isoscalar contribution to mT , which we call m0. The 0τ H H must be such that ωrtsu = 0, and from Eq. (A10) power of French’s product notation becomes particularly evident here, because the form of both terms is identical. T It demands some algebraic manipulation to find [J]W JT =0 ∴ = [J] JT / [J]. (B12) rstu Vrstu Vrstu X(J) X(J) X(J) or = + ¯ S (1 Φ(e))+ Hmnp Hm0 K Vα α − Applying this prescription to all the terms leads to (ob- α X viously δˆ δˆ = 1 in all cases) 1 rt su + ((¯+ + ¯−)S + (¯− ¯+)S )Φ(e) (B15) 2 Vα Vα α Vα − Vα α¯ T  = JT [J]/ [J] Vrstu Vrstu with X(J) X(J) T 1 Dr(Ds + 2Φ(P )( 1) ) = Γ [Γ]/ [Γ] [J]= − Vrstu Vrstu 4 1+Φ(P ) (Γ) (Γ) X(J) X X ±T JT J J [Γ] = Dr(Ds Φ(P ))/(1 + Φ(P )) rstu = rstu[J](1 ( 1) )/ [J](1 ( 1) ) − V V ± − ± − (Γ) J J X X X ± Γ+2r J 1 = rstu[Γ](1 ( 1) ) [J](1 ( 1) )= Dr(Dr 2) Vrstu V ± − ± − 4 ∓ Γ J X X Γ+2r 2r ±T [Γ](1 ( 1) )= Dr(Dr ( 1) ) (B16) Dr = [r], Φ(e)=1, for rstu . (B13) ± − ∓ − V Γ X

Through eqs.(B5a), (B5b), (B7a), and (B7b) we can Of course we must remember that for mnp, obtain the form of in terms of the monopole opera- 2r H HmT Dr =2jr +1, ( 1) = 1, Γ J, etc., while for m0 tors by regrouping the coefficients affecting each of them. − 2r− ≡ H Dr = 2(2jr + 1), ( 1) = +1, Γ JT ; etc. To simplify the presentation we adopt the following con- − ≡ It should be noted that m0 is not obtained by simply vention discarding the b coefficientsH in eqs.(B14), because we can extract some γ =00 contribution from the Tα operators. α rstu r s,t u, δˆ δˆ =1, BUT ≡ ≤ ≤ rt su The point will become quite clear when considering the ( Sα = Srtsu, Sα¯ = Srust,Tα = Trtsu,Tα¯ = Trust, diagonal contributions. 57

2. Diagonal forms of Hm A coordinates and momenta while only A 1 of them can be linearly independent since the solutions− cannot de- √ We are going to specialize to the diagonal terms of pend on the center of mass coordinate R = ( i ri)/ A Γ mT and mnp, which involve only rsrs matrix ele- or momentum P = ( pi)/√A. The way out is to im- H H V T i P ments whose centroids will be called simply rs and rs pose a factorization of the wavefunctions into relative and T V V P (the overline in rstu was meant to avoid confusion with CM parts: Φ(r1 r2 ...rA)=Φrel φCM . The potential en- V 1 0 possible matrix elements rstu or rstu, it can be safely ergy is naturally given in terms of relative values, and for dropped now). Then (B14)V becomesV Eq. (11) and (B15) the kinetic energy we should do the same by referring to becomes the CM momentum,

P 2 2 2 1 2 d d (pi ) = pi P = (pi pj) , (C1) mnp or m0 = + rs nr(ns δrs)/(1+δrs) (B17) − √A − A − H H K V − i i ij r≤s X X X X and change accordingly in Eq. (2). As we are only We rewrite the relevant centroids incorporating explic- Kij itly the Pauli restrictions interested in wavefunctions in which the center of mass is at rest (or in its lowest possible state), we can add a Γ Γ 2 2 Γ rsrs[Γ](1 ( 1) δrs) CM operator to the Hamiltonian = + λ(R + P ), rs = V − − (B18a) H ⇒ H V D (D δ ) and calling rij = ri rj we have P r s rs − JT − J+T 4 [J](1 ( 1) δrs) T J rsrs 2 2 2 2 1 2 2 rs = V − − T (B18b) R + P = (r + p ) (r + p ), (C2) V Dr(Ds +2δrs( 1) ) i i − A ij ij P − i ij X X

1 1 0 3 δrs so that upon diagonalization the eigenvalues will be of ars = (3 rs + rs)= rs + brs 4 V V V 4 D 1 the form E = Erel + λ(NCM + 3/2), and it only re- r − 1 0 mains to select the states with NCM =0. We are tak- brs = (B18c) Vrs −Vrs ing for granted separation of CM and relative coordi- The relationship between a and makes it possible nates. Unfortunately, this happens only for spaces that rs Vrs to combine eqs. (11) and (B17) in a single form are closed under the CM operator (C2): for NCM to be a good quantum number the space must include all possi- d 1 ble states with NCM oscillator quanta. The problem was m = + rs nr(ns δrs)+ H K (1 + δrs) V − raised by Elliott and Skyrme (1955) who initiated the X  3nrn¯r study of “particle-hole” (ph) excitations on closed shells. +brs Tr Ts δrs (B19) They noted that acting with R iP on the IPM ground · − 4(Dr 1) −  −  state of 16O ( 0 ) leads to | i in which now the brs term can be dropped to obtain d or d . 1 m0 mnp (¯p s √2¯p s √5¯p d +¯p d +3¯p d ) 0 , (C3) H In theH np scheme each orbit r goes into two r and r 1 1 3 1 1 3 3 3 3 5 n p r18 − − | i and the centroids can be obtained through (x, y = n or p, x = y) wherep ¯2j removes a particle and s1 or d2j add a particle 6 on 0 . As R iP has tensorial rank J πT =1−0, Eq. (C3) | i − − 1 2δrs 2δrs is telling us that out of five possible 1 0 excitations, one = 1 1 + 0 1+ Vrxsy 2 Vrs − D Vrs D is “spurious” and has to be discarded.   r   r  = 1 (B20) Assume now that we are interested in 2p2p excita- Vrxsx Vrs tions. They involve jumps of two oscillator quanta (2~ω ) Note that the diagonal terms depend on the represen- and the CM eigenstates (R iP )2 0 involve the op- d d erator in Eq. (C3), but also− jumps| toi other shells, of tation: mnp = mT in general. H 6 H the typesp ¯ p¯(sd) s¯(sd) or (sd¯ ) (pf)¯p (sd) p¯(pf). Therefore, as anticipated,≡ relative-CM factorization≡ can APPENDIX C: The center of mass problem be achieved only by including all states involving a given number of oscillator quanta. The clean way to pro- What do you do about center of mass? is probably the ceed is through complete N~ω spaces, discussed in Sec- standard question most SM practitioners prefer to ignore tions II.A.1 and III.F. The 0~ω and EI spaces are also or dismiss. Even if we may be tempted to do so, there is free of problems (the latter because no 1~ω 1−0 states no excuse for ignoring what the problem is, and here we exist). It remains to analyze the EEI valence spaces, shall explain it in sufficient detail, so as to dispel some where CM spuriousness is always present but strongly common misconceptions. suppressed because the main contributors to R iP –of The center of mass (CM) problem arises because in a the type pj = p +1j 1, with the largest j–− are al- ⇒ ± many body treatment it is most convenient to work with ways excluded. Consider EEI(1)=p1/2, d5/2, s1/2. The 58

− only possible 1~ω 1 0 state isp ¯1s1, which according to P. Vogel, 1996, Phys. Rev. Lett. 77, 5186. Eq. (C3) accounts for (1/18)% = 5.6% of the spurious Bambynek, W., H. Behrens, M. H. Chen, B. Crasemann, state. This apparently minor problem was unduly trans- M. L. Fitzpatrick, K. W. D. Ledingham, H. Genz, M. Mut- formed into a serious one through a proposal by Gloeck- terer, and R. L. Intermann, 1977, Rev. Mod. Phys. 49, 77; 49, 961(E). ner and Lawson (1974). It amounts to project CM spuri- 11 ousness through the = +λ(R2+P 2) prescription in Bansal, R. K., and J. B. French, 1964, Phys. Lett. , 145. Baranger, M., and K. Kumar, 1968, Nucl. Phys. A 110, 490. the EEI(1) space byH identifying⇒ H R iP p¯ s . The pro- − ≡ 1 1 Bardin, R. K., P. J. Gollon, J. D. Ullman, and C. S. Wu, cedure is manifestly incorrect, as was repeatedly pointed 1970, Nucl. Phys. A 158, 337. out (see for instance Whitehead et al., 1977) but the mis- B¨aumer, C., A. M. van den Berg, B. Davids, D. Frekers, D. De conception persists. An interesting and viable alternative Frenne, E.-W. Grewe, P. Haefner, M. N. H. F. Hofmann, was put forward by Dean et al. (1999) in calculations with M. Hunyadi, E. Jacobs, B. J. A. Korff, K. Langanke, et al., two contiguous major shells. Further work on the subject 2003, Phys. Rev. C 68, 031303. would be welcome. Behrens, H., and W. B¨uhring, 1982, Electron Radial Wave Functions and Nuclear Beta-decay (Clarendon, Oxford). Bender, M., H. Flocard, and P.-H. Heenen, 2003a, arXiv:nucl- th/0305021. References Bender, M., P.-H. Heenen, and P. G. Reinhard, 2003b, Rev. Mod. Phys. 75, 121. Abgrall, Y., G. Baron, E. Caurier, and G. Monsonego, 1967a, Bennaceur, K., F. Nowacki, J. Okolowicz, and M. Ploszajczak, Phys. Lett. 26B, 53. 1999, Nucl. Phys. A 651, 289. Abgrall, Y., E. Caurier, and G. Monsonego, 1967b, Phys. Bennaceur, K., F. Nowacki, J. Okolowicz, and M. Ploszajczak, Lett. 24B, 609. 2000, Nucl. Phys. A 671, 203. Abzouzi, A., E. Caurier, and A. P. Zuker, 1991, Phys. Rev. Bentley, M. A., C. D. O’Leary, A. Poves, G. Mart´ınez-Pinedo, Lett. 66, 1134. D. E. Appelbe, R. A. Bark, D. M. Cullen, S. Ert¨urk, and Ahmad, Q. R., R. C. Allen, T. C. Andersen, J. D. Anglin, A. Maj, 1998, Phys. Lett. B 437, 243. J. C. Barton, E. W. Beier, M. Bercovitch, J. Bigu, S. Biller, Bentley, M. A., C. D. O’Leary, A. Poves, G. Mart´ınez-Pinedo, R. A. Black, I. Blevis, R. J. Boardman, et al. (SNO), 2002, D. E. Appelbe, R. A. Bark, D. M. Cullen, S. Ert¨urk, and Phys. Rev. Lett. 89, 011301. A. Maj, 1999, Phys. Lett. B 451, 445. Alford, W. P., B. A. Brown, S. Burzynski, A. Celler, D. Frek- Bentley, M. A., S. J. Williams, D. T. Joss, C. D. O’Leary, ers, R. Helmer, R. Henderson, K. P. Jackson, K. Lee, A. Ra- A. M. Bruce, J. A. Cameron, P. Fallon, L. Frankland, hav, A. Trudel, and M. C. Vetterli, 1993, Phys. Rev. C 48, W. Gelletly, C. J. Lister, G. Mart´ınez-Pinedo, A. Poves, 2818. et al., 2000, Phys. Rev. C 62, 051303R. Alhassid, Y., G. F. Bertch, D. J. Dean, and S. E. Koonin, Bertsch, G. F., and I. Hamamoto, 1982, Phys. Rev. C 26, 1996, Phys. Rev. Lett. 77, 1444. 1323. Alhassid, Y., D. J. Dean, S. E. Koonin, G. Lang, and W. E. Bes, D. R., and R. A. Sorensen, 1969, in Adv. Nuc. Phys., Ormand, 1994, Phys. Rev. Lett. 72, 613. edited by M. Baranger and E. Vogt (Plenum Press, New Alhassid, Y., S. Liu, and H. Nakada, 1999, Phys. Rev. Lett. York). 83, 4265. Bethe, H. A., 1936, Phys. Rev. 50, 332. Altshuler, B. L., Y. Gefenand, A. Kamenevand, and L. S. Bhattacharya, M., A. Garc´ıa, N. I. Kaloskamis, E. G. Adel- Levitov, 1997, Phys. Rev. Lett. 78, 2803. berger, H. E. Swanson, R. Anne, M. Lewitowicz, M. G. Anderson, B. D., T. Chittrakarn, A. R. Baldwin, C. Lebo, Saint-Laurent, W. Trinder, C. Donzaud, D. Guillemaud- R. Madey, P. C. Tandy, J. W. Watson, B. A. Brown, and Mueller, S. Leenhardt, et al., 1998, Phys. Rev. C , 3677. C. C. Foster, 1985, Phys. Rev. C 31, 1161. Bijker, R., and A. Frank, 2000a, Phys. Rev. Lett. 84, 420. Anderson, B. D., C. Lebo, A. R. Baldwin, T. Chittrakarn, Bijker, R., and A. Frank, 2000b, Phys. Rev. C 62, 014303. R. Madey, and J. W. Watson, 1990, Phys. Rev. C 41, 1474. Bijker, R., A. Frank, and S. Pittel, 1999, Phys. Rev. C 60, Andreozzi, F., N. Lo Iudice, and A. Porrino, 2003, to be pub- 021302. lished in Phys. Rev. C, eprint nucl-th/0307046. Bloom, S. D. B., 1984, Prog. Part. Nucl. Phys. 11, 505. Arima, A., 2001, Prog. Part. Nucl. Phys. 46, 119. Bohr, A., 1952, Kgl. Dan. Vid. Sel. Mat. Fys. Med. 26, 14. Arima, A., M. Harvey, and K. Shimuzi, 1969, Phys. Lett. B Bohr, A., and B. R. Mottelson, 1953, Kgl. Dan. Vid. Sel. Mat. 30, 517. Fys. Med. 27, 16. Arima, A., and F. Iachello, 1975, Phys. Rev. Lett. 35, 1069. Bohr, A., and B. R. Mottelson, 1969, Nuclear Structure, vol. Aumann, T., A. Navin, D. P. Balamuth, D. Bazin, B. Blank, I (Benjamin, New York). B. A. Brown, J. E. Bush, J. A. Caggiano, B. Davids, Bohr, A., B. R. Mottelson, and D. Pines, 1958, Phys. Rev. T. Glasmacher, V. Guimara˜es, P. G. Hansen, et al., 2001, 110, 936. Phys. Rev. Lett. 84, 35. Borcea, R., J. Ayst¨o,¨ E. Caurier, P. Dendooven, J. D¨oringa, Axel, O., J. H. D. Jensen, and H. E. Suess, 1949, Phys. Rev. M. Gierlik, M.G´orska, H.Grawe, M. Hellstr¨om, Z. Janas, 75, 1766. A. Jokinen, M. Karny, et al., 2001, Nucl. Phys. A 695, 69. Axiotis, M., 2000, private communication. Borge, M. J. G., H. Fymbo, D. Guillemaud-Mueller, P. Horn- Azaiez, F., 1999, in Experimental Nuclear Physics in Europe, shoj, F. Humbert, B. Jonson, T. E. Leth, G. Mart´ınez- edited by B. Rubio, M. Lozano, and W. Gelletly, volume Pinedo, T. Nilsson, G. Nyman, A. Poves, I. Ramos-Lerate, 495 of AIP Conf. Proc., p. 171. et al., 1997, Phys. Rev. C 55, R8. Balysh, A., A. De Silva, V. I. Lebedev, K. Lou, M. K. Moe, Brandolini, F., S. M. Lenzi, D. R. Napoli, R. V. Rivas, H. So- M. A. Nelson, A. Piepke, A. Pronskiy, M. A. Vient, and macal, C. A. Ur, D. Bazzacco, A. C. J, G. de Angelis, 59

M. de Poli, C. Fahlander, A. Gadea, et al., 1998, Nucl. 693, 374. Phys. A 642, 387. Caurier, E., F. Nowacki, A. Poves, and J. Retamosa, 1996, Brandolini, F., N. Marginean, S. Hankonen, N. H. Medina, Phys. Rev. Lett. 77, 1954. R. V. Rivas, J. S´anchez Solano, S. M. Lenzi, S. Lunardi, Caurier, E., F. Nowacki, A. Poves, and J. Retamosa, 1998, D. R. Napoli, A. Poves, C. A. Ur, D. Bazzacco, et al., 2002a, Phys. Rev. C 58, 2033. Phys. Rev. C 66, 024304. Caurier, E., A. Poves, and A. P. Zuker, 1990, Phys. Lett. B Brandolini, F., N. H. Medina, S. M. Lenzi, D. R. Napoli, 252, 13. A. Poves, R. V. Rivas, J. S´anchez Solano, C. A. Ur, D. Bu- Caurier, E., A. Poves, and A. P. Zuker, 1995b, Phys. Rev. curescu, M. De Poli, R. Menegazzo, D. Bazzacco, et al., Lett. 74, 1517. 1999, Phys. Rev. C 60, 041305. Caurier, E., M. Rejmund, and H. Grawe, 2003, Phys. Rev. C Brandolini, F., N. H. Medina, R. V. Rivas, S. M. Lenzi, , 67, 054310. A. Gadea, C. A. Ur, D. Bazzacco, R. Menegazzo, P. Pavan, Caurier, E., A. P. Zuker, A. Poves, and G. Mart´ınez-Pinedo, C. Rossi-Alvarez, A. Algora-Pineda, et al., 2001, Phys. Rev. 1994, Phys. Rev. C 50, 225. C 64, 044307. Cavedon, J. M., J. B. Frois, D. Goutte, , M. Huet, P. Leconte, Brandolini, F., J. S´anchez Solano, S. M. Lenzi, N. H. Medina, C. Papanicolas, X. H. Pham, S. Platchkov, S. Williamson, A. Poves, C. A. Ur, D. Bazzacco, G. de Angelis, M. De W. Boeglin, and I. Sick, 1982, Phys. Rev. Lett. 49, 978. Poli, E. Farnea, A. Gadea, D. R. Napoli, et al., 2002b, Chau Huu-Tai, P., A. Frank, N. A. Smirnova, and P. Van Phys. Rev. C 66, 021302. Isacker, 2003, arXiv:nucl-th/0301061. Brown, B. A., 2001, Prog. Part. Nucl. Phys. 47, 517. Chevallier, P., F. Scheibling, G. Goldring, I. Plesser, and Brown, B. A., R. Clement, H. Schatz, J. Giansiracusa, W. A. M. W. Sachs, 1967, Phys. Rev. 160, 827. Richter, M. Hjorth-Jensen, K. L. Kratz, B. Pfeiffer, and Chiara, C. J., E. Ideguchi, M. Devlin, D. R. LaFosee, W. B. Walters, 2003, Nucl. Phys. A 719, 177c. F. Lerma, W. Reviol, S. K. Yu, D. G. Sarantites, C. Back- Brown, B. A., A. Etchegoyen, W. D. M. Rae, N. S. Godwin, tash, A. Galindo-Uribarri, M. P. Carpenter, R. V. F. W. A. Richter, C. H. Zimmerman, W. E. Ormand, and Janssens, et al., 2003, Phys. Rev. C 67, 041303. J. S. Winfield, 1985, OXBASH code, Technical Report 524, Chou, W.-T., E. K. Warburton, and B. A. Brown, 1993, Phys. MSU-NSCL. Rev. C 47, 163. Brown, B. A., D. B. Fossan, J. M. McDonald, and K. A. Coester, F., and H. K¨ummel, 1960, Nucl. Phys. 17, 477. Snover, 1974, Phys. Rev. C 9, 1033. Cohen, S., and D. Kurath, 1965, Nucl. Phys. 73, 1. Brown, B. A., and B. H. Wildenthal, 1988, Annu. Rev. Nucl. Cortes, A., R. U. Haq, and A. P. Zuker, 1982, Phys. Lett. B Part. Sci. 38, 29. 115, 1. Brueckner, K., 1954, Phys. Rev. 96, 508. Covello, A., F. Andreozzi, L. Coraggio, A. Gargano, T. T. S. Cameron, J. A., M. A. Bentley, A. M. Bruce, R. A. Cunning- Kuo, and A. Porrino, 1997, Prog. Part. Nucl. Phys. 18, ham, W. Gelletly, H. G. Price, J. Simpson, D. D. Warner, 165. and A. N. James, 1991, Phys. Rev. C 44, 1882. Dang, N. D., A. Arima, T. Suzuki, and S. Yamaji, 1997, Nucl. Cameron, J. A., M. A. Bentley, A. M. Bruce, R. A. Cunning- Phys. A 621, 719. ham, W. Gelletly, H. G. Price, J. Simpson, D. D. Warner, Davidson, E. R., 1975, J. Comp. Phys. 17, 87. and A. N. James, 1994, Phys. Rev. C 49, 1347. Dean, D. J., S. E. Koonin, K. Langanke, P. B. Radha, and Cameron, J. A., M. A. Bentley, A. M. Bruce, R. A. Cunning- Y. Alhassid, 1995, Phys. Rev. Lett. 74, 2009. ham, H. G. Price, J. Simpson, D. D. Warner, A. N. James, Dean, D. J., M. T. Russell, M. Hjorth-Jensen, S. E. Koonin, W. Gelletly, and P. V. Isacker, 1993, Phys. Lett. B 319, K. Langanke, and A. P. Zuker, 1999, Phys. Rev. C 59, 58. 2474. Campi, X., H. Flocard, A. K. Kerman, and S. Koonin, 1975, Decharg´e, J., and D. Gogny, 1980, Phys. Rev. C 21, 1568. Nucl. Phys. A 251, 193. Descouvemont, P., 2002, Nucl. Phys. A 709, 275. Carter, E. B., G. E. Mitchell, and R. H. Davis, 1964, Phys. Detraz, C., D. Guillemaud, G. Huber, R. Klapisch, M. de Rev. 133, 1421. Saint-Simon, C. Thibault, and F. Touchard, 1979, Phys. Caurier, E., J. L. Egido, G. Mart´ınez-Pinedo, A. Poves, J. Re- Rev. C 19, 164. tamosa, L. M. Robledo, and A. P. Zuker, 1995a, Phys. Rev. Dillmann, I., K.-L. Kratz, A. W¨ohr, O. Arndt, B. A. Brown, Lett. 75, 2466. P. Hoff, M. Hjorth-Jensen, U. K¨oster, A. N. Ostrowski, Caurier, E., K. Langanke, G. Mart´ınez-Pinedo, and B. Pfeiffer, D. Seweryniak, J. Shergur, et al., 2003, Phys. F. Nowacki, 1999a, Nucl. Phys. A 653, 439. Rev. Lett. 91, 162503. Caurier, E., K. Langanke, G. Mart´ınez-Pinedo, F. Nowacki, Doi, M., T. Kotani, and E. Takasugi, 1985, Prog. Theor. Phys. and P. Vogel, 2001a, Phys. Lett. B 522, 240. Suppl. 83, 1. Caurier, E., G. Mart´ınez-Pinedo, F. Nowacki, A. Poves, J. Re- Dro˙zd˙z, S., V. Klemt, J. Speth, and J. Wambach, 1986, Phys. tamosa, and A. P. Zuker, 1999b, Phys. Rev. C 59, 2033. Lett. B 166, 18. Caurier, E., G. Mart´ınez-Pinedo, F. Nowacki, A. Poves, J. Re- Duflo, J., and A. P. Zuker, 1995, Phys. Rev. C 52, R23. tamosa, and A. P. Zuker, 1999c, Nucl. Phys. A 164, 747c, Duflo, J., and A. P. Zuker, 1999, Phys. Rev. C 59, R2347. proceedings INPC/98, B. Frois editor. Duflo, J., and A. P. Zuker, 2002, Phys. Rev. C 66, 051303(R). Caurier, E., P. Navr´atil, W. E. Ormand, and J. P. Vary, Dufour, M., and A. P. Zuker, 1996, Phys. Rev. C 54, 1641. 2001b, Phys. Rev. C 64, 051301(R). Dukelsky, J., and S. Pittel, 2001, Phys. Rev. C 63, 061303. Caurier, E., P. Navr´atil, W. E. Ormand, and J. P. Vary, 2002, Dukelsky, J., S. Pittel, S. S. Dimitrova, and M. V. Stoistov, Phys. Rev. C 66, 024314. 2002, Phys. Rev. C 65, 054319. Caurier, E., and F. Nowacki, 1999, Acta Physica Polonica 30, Duvernois, M. A., 1997, Astrophys. J. 481, 241. 705. Eguchi, K., S. Enomoto, K. Furuno, J. Goldman, H. Hanada, Caurier, E., F. Nowacki, and A. Poves, 2001c, Nucl. Phys. A H. Ikeda, K. Ikeda, K. Inoue, K. Ishihara, W. Itoh, 60

T. Iwamoto, T. Kawaguchi, et al. (KamLAND), 2003, Hagiwara, K., K. Hikasa, K. Nakamura, M. Tanabashi, Phys. Rev. Lett. 90, 021802. M. Aguilar-Benitez, C. Amsler, R. Barnett, P. Burchat, El-Kateb, S., K. P. Jackson, W. P. Alford, R. Abegg, R. E. C. Carone, C. Caso, G. Conforto, O. Dahl, et al. (Par- Azuma, B. A. Brown, A. Celler, D. Frekers, O. Haeusser, ticle Data Group), 2002, Phys. Rev. D 66, 010001, URL R. Helmer, R. S. Henderson, K. H. Hicks, et al., 1994, Phys. http://pdg.lbl.gov. Rev. C 49, 3128. Halbert, E. C., J. B. McGrory, B. H. Wildenthal, and Elliott, J. P., 1958a, Proc. Roy. Soc. London , 128. S. Pandya, 1971, in Adv. Nuc. Phys., edited by M. Baranger Elliott, J. P., 1958b, Proc. Roy. Soc. London , 562. and E. Vogt (Plenum Press, New York). Elliott, J. P., A. D. Jackson, H. A. Mavromantis, E. A. Hannawald, M., T. Kautsch, A. W¨ohr, W. Walters, K.-L. Sanderson, and B. Sing, 1968, Nucl. Phys. A 121, 241. Kratz, V. N. Fedoseyev, V. I. Mishin, W. B¨ohmer, B. Pfeif- Elliott, J. P., and A. M. Lane, 1957, in Handbuch der Physik fer, V. Sebastian, Y. Jading, U. K¨oster, et al., 1999, Phys. XXXIX (Springer Verlag, Berlin). Rev. Lett. 82, 1391. Elliott, J. P., and T. H. R. Skyrme, 1955, Proc. Roy. Soc. Hansen, P. G., and B. Jonson, 1987, Europhysics Lett. 4, 409. London A 232, 561. Hara, K., and Y. Sun, 1995, Int. J. Mod. Phys. E 4, 637. Endt, P. M., and C. van der Leun, 1990, Nucl. Phys. A 521, Hara, K., Y. Sun, and T. Mizusaki, 1999, Phys. Rev. Lett. 1. 83, 1922. Entem, D. R., and R. Machleidt, 2002, Phys. Lett. B 524, 93. Hauschild, K., M. Rejmund, and H. Grawe, 2001, Phys. Rev. Entem, D. R., R. Machleidt, and H. Witala, 2002, Phys. Rev. Lett. 87, 072501. C 65, 064005. Haxton, W. C., 1987, Phys. Rev. D 36, 2283. Fiase, J., A. Hamoudi, J. M. Irvine, and F. Yazici, 1988, J. Haxton, W. C., 1998, Phys. Lett. B 431, 110. Phys G 14, 27. Haxton, W. C., and C. Johnson, 1990, Phys. Rev. Lett. 65, Firestone, R. B., 1996, Table of Isotopes (Wiley, New York). 1325. Fisker, J. L., V. Barnard, J. G¨orres, K. Langanke, Hayes, A. C., and I. S. Towner, 2000, Phys. Rev. C 61, G. Mart´ınez-Pinedo, and M. C. Wiescher, 2001, At. Data. 044603. Nucl. Data Tables 79, 241. Hecht, K. T., and A. Adler, 1969, Nucl. Phys. A A137, 129. Fisker, J. L., G. Mart´ınez-Pinedo, and K. Langanke, 1999, Heger, A., E. Kolbe, W. Haxton, K. Langanke, G. Mart´ınez- Eur. Phys. J. A 5, 229. Pinedo, and S. E. Woosley, 2003, submitted to Phys. Rev. French, J. B., 1966, in Lectures XXXVI, edited Lett., eprint astro-ph/0307546. by C. Bloch (Academic Press, New York). Heger, A., K. Langanke, G. Mart´ınez-Pinedo, and S. E. French, J. B., 1969, in Isospin in Nuclear Physics, edited by Woosley, 2001a, Phys. Rev. Lett. 86, 1678. D. H. Wilkinson (North Holland, Amsterdam). Heger, A., S. E. Woosley, G. Mart´ınez-Pinedo, and K. Lan- French, J. B., E. C. Halbert, J. B. McGrory, and S. S. M. ganke, 2001b, Astrophys. J. 560, 307. Wong, 1969, in Adv. Nuc. Phys., edited by M. Baranger Heisenberg, J. H., and B. M. Mihaila, 2000, Phys. Rev. C 59, and E. Vogt (Plenum Press, New York), p. 193. 1440. Fricke, G., C. Bernhardt, K. Heilig, L. A. Schaller, L. Schel- Herndl, H., J. G¨orres, M. Wiescher, B. A. Brown, and L. van lenberg, E. B. Shera, and C. W. D. Jager, 1995, At. Data Wormer, 1995, Phys. Rev. C 52, 1078. Nucl. Data Tables 60, 177. Heyde, K., and J. L. Woods, 1991, J. Phys. G 17, 135. Fujita, Y., H. Fujita, T. Adachi, G. P. A. Berg, E. Caurier, Hix, W. R., O. E. B. Messer, A. Mezzacappa, H. Fujimura, K. Hara, K. Hatanaka, Z. Janas, J. Kamiya, M. Liebend¨orfer, J. Sampaio, K. Langanke, D. J. T. Kawabata, K. Langanke, et al., 2002, Eur. Phys. J. A Dean, and G. Mart´ınez-Pinedo, 2003, Phys. Rev. Lett. 91, 13, 411. 201102. Fukuda, Y., T. Hayakawa, E. Ichihara, K. Inoue, K. Ishi- Hjorth-Jensen, M., 1996, private communication to APZ. hara, H. Ishino, Y. Itow, T. Kajita, J. Kameda, S. Kasuga, Hjorth-Jensen, M., T. T. S. Kuo, and E. Osnes, 1995, Phys. K. Kobayashi, Y. Kobayashi, et al. (Super-Kamiokande), Repts. 261, 126. 1998, Phys. Rev. Lett. 81, 1562. Honma, M., T. Mizusaki, and T. Otsuka, 1996, Phys. Rev. Fukunishi, N., T. Otsuka, and T. Sebe, 1992, Phys. Lett. B Lett. 77, 3315. 296, 279. Honma, M., T. Otsuka, B. A. Brown, and T. Mizusaki, 2002, Fuller, G. M., W. A. Fowler, and M. J. Newman, 1980, As- Phys. Rev. C 65, 061301. trophys. J. Suppl. 42, 447. Horoi, M., B. A. Brown, and V. Zelevinsky, 2001, Phys. Rev. Fuller, G. M., W. A. Fowler, and M. J. Newman, 1982a, As- Lett. 87, 062501. trophys. J. 252, 715. Horoi, M., B. A. Brown, and V. Zelevinsky, 2002, Phys. Rev. Fuller, G. M., W. A. Fowler, and M. J. Newman, 1982b, As- C 65, 027303. trophys. J. Suppl. 48, 279. Horoi, M., B. A. Brown, and V. Zelevinsky, 2003, Phys. Rev. Fuller, G. M., W. A. Fowler, and M. J. Newman, 1985, As- C 67, 034303. trophys. J. 293, 1. Horoi, M., A. Volya, and V. Zelevinsky, 1999, Phys. Rev. Lett. Garrett, P. E., W. E. Ormand, D. Appelbe, R. W. Bauer, J. A. 82, 2064. Becker, L. A. Bernstein, J. A. Cameron, M. P. Carpenter, Iachello, F., and A. Arima, 1987, The interacting boson model R. V. F. Janssens, C. J. Lister, D. Seweryniak, E. Tavukcu, (Cambridge University Press, Cambridge). et al., 2001, Phys. Rev. Lett. 87, 132502. Ibbotson, R. W., T. Glasmacher, P. F. Mantica, and Glasmacher, T., 1998, Annu. Rev. Nucl. Part. Sci. 48, 1. H. Scheit, 1999, Phys. Rev. C 59, 642. Gloeckner, D., and R. Lawson, 1974, Phys. Lett. 53B, 313. Ichimura, M., and A. Arima, 1966, Prog. Theor. Phys. 36, Guillemaud-Mueller, D., C. Detraz, M. Langevin, F. Naulin, 296. M. de Saint-Simon, C. Thibault, F. Touchard, and Id Betan, R., R. J. Liotta, N. Sandulescu, and T. Vertse, 2002, M. Epherre, 1984, Nucl. Phys. A 426, 37. Phys. Rev. Lett. 89, 042501. 61

Ideguchi, E., D. G. Sarantites, W. Reviol, A. V. Afanasiev, Langanke, K., G. Mart´ınez-Pinedo, O. E. B. Messer, J. M. M. Devlin, C. Baktash, R. V. F. Janssens, D. Rudolph, Sampaio, D. J. Dean, W. R. Hix, A. Mezzacappa, A. Axelsson, M. P. Carpenter, A. Galindo-Uribarri, D. R. M. Liebend¨orfer, H.-T. Janka, and M. Rampp, 2003a, Phys. LaFosse, et al., 2001, Phys. Rev. Lett. 87, 222501. Rev. Lett. 90, 241102. Ikeda, K. I., S. Fujii, and J. I. Fujita, 1963, Phys. Lett. 3, Langanke, K., J. Terasaki, F. Nowacki, D. J. Dean, and 271. W. Nazarewicz, 2003b, Phys. Rev. C 67, 044314. Iljinov, A. S., M. V. Mebel, N. Bianchi, E. de Sanctis, Lawson, R. D., and M. H. Macfarlane, 1965, Nuc. Phys. A C. Guaraldo, V. Lucherini, V. Muccifora, E. Polli, A. R. 66, 80. Reolon, and P. Rossi, 1992, Nucl. Phys. A A543, 517. Lee, H. C., 1969, code made available to APZ. Iwasaki, H., T. Motobayashi, H. Sakurai, K. Yoneda, Lenzi, S. M., N. Marginean, D. R. Napoli, C. A. Ur, A. P. T. Gomi, N. Aoi, N. Fukuda, , Z. F¨ul¨op, U. Futakami, Zuker, G. de Angelis, A. Algora, M. Axiotis, D. Bazzacco, Z. Gacsi, Y. Higurashi, et al., 2001, Phys. Lett. B 522, N. Belcari, M. A. Bentley, P. G. Bizzeti, et al., 2001, Phys. 227. Rev. Lett. 87, 122501. Janssens, R. V. F., B. Fornal, P. F. Mantica, B. A. Brown, Lenzi, S. M., D. R. Napoli, A. Gadea, M. A. Cardona, D. Ho- R. Broda, P. Bhattacharyya, M. P. Carpenter, M. Cin- jman, M. A. Najarajan, C. Rossi-Alvarez, N. H. Medina, ausero, P. J. Daly, A. D. Davies, T. Glasmacher, Z. W. G. de Angelis, D. Bazzacco, M. E. Debray, M. De Poli, Grabowski, et al., 2002, Phys. Lett. B 546, 55. et al., 1996, Z. Phys. A 354, 117. Jiang, M. F., R. Machleidt, D. B. Stout, and T. T. S. Kuo, Lenzi, S. M., D. R. Napoli, C. A. Ur, D. Bazzacco, F. Bran- 1989, Phys. Rev. C 40, 1857. dolini, J. A. Cameron, E. Caurier, G. de Angelis, M. De Johnson, C. W., G. F. Bertsch, and D. J. Dean, 1998, Phys. Poli, E. Farnea, A. Gadea, S. Hankonen, et al., 1999, Phys. Rev. Lett. 80, 2749. Rev. C 60, 021303. Johnson, C. W., G. F. Bertsch, D. J. Dean, and I. Talmi, Lenzi, S. M., C. A. Ur, D. R. Napoli, M. A. Najarajan, D. Baz- 1999, Phys. Rev. C 61, 014311. zacco, D. M. Brink, M. A. Cardona, G. de Angelis, M. De Juodagalvis, A., and S. Aberg,˚ 1998, Phys. Lett. B 428, 227. Poli, A. Gadea, D. Hojman, S. Lunardi, et al., 1997, Phys. Kahana, S., H. Lee, and C. Scott, 1969a, Phys. Rev. 180, Rev. C 56, 1313. 956. Long, G.-L., and Y. Sun, 2001, Phys. Rev. C 63, 021305. Kahana, S., H. Lee, and C. Scott, 1969b, Phys. Rev. 185, Machleidt, R., 2001, Phys. Rev. C 63, 024001. 1378. Machleidt, R., F. Sammaruca, and Y. Song, 1996, Phys. Rev. K¨appeler, F., F.-K. Thielemann, and M. Wiescher, 1998, C 53, R1483. Annu. Rev. Nucl. Part. Sci. 48, 175. Mantica, P. F., A. C. Morton, B. A. Brown, A. D. Davis, Kawarada, H., and A. Arima, 1964, J. Phys. Soc. Japan 19, T. Glasmacher, D. E. Groh, S. N. Liddick, D. J. Morrisey, 1768. W. F. Muelle, H. Schatz, A. Stolz, S. L. Tabor, et al., 2003, Klotz, G., P. Baumann, M. Bounajma, A. Huck, A. Knipper, Phys. Rev. C 67, 014311. G. Walter, G. Marguier, C. Richard-Serre, A. Poves, and Marsden, D. C. J., P. Navr´atil, S. A. Coon, and B. R. Barrett, J. Retamosa, 1993, Phys. Rev. C 47, 2502. 2002, Phys. Rev. C 66, 044007. Kolbe, E., K. Langanke, and G. Mart´ınez-Pinedo, 1999, Phys. Mart´ınez-Pinedo, G., 2001, Nucl. Phys. A 688, 357c. Rev. C 60, 052801. Mart´ınez-Pinedo, G., and K. Langanke, 1999, Phys. Rev. Kolbe, E., K. Langanke, G. Mart´ınez-Pinedo, and P. Vogel, Lett. 83, 4502. 2003, J. Phys. G: Nucl. Part. Phys. 29, 2569. Mart´ınez-Pinedo, G., A. Poves, E. Caurier, and A. P. Zuker, Koonin, S. E., D. J. Dean, and K. Langanke, 1997a, Phys. 1996a, Phys. Rev. C 53, R2602. Repts. 278, 2. Mart´ınez-Pinedo, G., A. Poves, L. M. Robledo, E. Caurier, Koonin, S. E., D. J. Dean, and K. Langanke, 1997b, Ann. F. Nowacki, J. Retamosa, and A. P. Zuker, 1996b, Phys. Rev. Nucl. Part. Sci. 47, 463. Rev. C 54, R2150. Kratz, K.-L., B. Pfeiffer, and F.-K. Thielemann, 1998, Nucl. Mart´ınez-Pinedo, G., and P. Vogel, 1998, Phys. Rev. Lett. Phys. A 630, 352. 81, 281. Kumar, K., and M. Baranger, 1968, Nucl. Phys. A 110, 529. Mart´ınez-Pinedo, G., A. P. Zuker, A. Poves, and E. Caurier, K¨ummel, H. K., K. H. L¨uhrmann, and J. G. Zabolitzky, 1978, 1997, Phys. Rev. C 55, 187. Phys. Rep. 36, 1. Mayer, M., 1949, Phys. Rev. 75, 1969. Kuo, T. T. S., and G. Brown, 1968, Nucl. Phys. A 114, 241. Mazzocchi, C., Z. Janas, J. D¨oring, M. Axiotis, L. Batist, Kuo, T. T. S., and G. E. Brown, 1966, Nucl. Phys. A 85, 40. R. Borcea, D. Cano-Ott, E. Caurier, G. de Angelis, Kuo, T. T. S., and G. H. Herling, 1971, NRL Memorandum E. Farnea, A. Faßbender, A. Gadea, et al., 2001, Eur. Phys. Report 2258. J. A 12, 269. Lanczos, C., 1950, J. Res. Nat. Bur. Stand. 45, 252. McCullen, J. D., B. F. Bayman, and L. Zamick, 1964, Phys. Langanke, K., 1998, Phys. Lett. B 438, 235. Rev. 134, 515. Langanke, K., D. J. Dean, P. B. Radha, Y. Alhasid, and S. E. Michel, N., W. Nazarevicz, M. Ploszajczak, and K.Bennaceur, Koonin, 1995, Phys. Rev. C 52, 718. 2002, Phys. Rev. Lett. 89, 042502. Langanke, K., E. Kolbe, and D. J. Dean, 2001, Phys. Rev. C Michel, N., W. Nazarevicz, M. Ploszajczak, and J. Okolowicz, 63, 032801. 2003, Phys. Rev. C 67, 054311. Langanke, K., and G. Mart´ınez-Pinedo, 2000, Nucl. Phys. A Mihaila, B. M., and J. H. Heisenberg, 2000, Phys. Rev. C 61, 673, 481. 054309. Langanke, K., and G. Mart´ınez-Pinedo, 2001, At. Data. Nucl. Mizusaki, T., 2000, RIKEN Accel. Prog. Rep. 33, 14. Data Tables 79, 1. Mizusaki, T., and M. Imada, 2002, Phys. Rev. C 65, 064319. Langanke, K., and G. Mart´ınez-Pinedo, 2003, Rev. Mod. Mizusaki, T., and M. Imada, 2003, Phys. Rev. C 67, 041301. Phys. 75, 819. Mizusaki, T., T. Otsuka, M. Honma, and B. A. Brown, 2001, 62

Phys. Rev. C 63, 044306. Osterfeld, F., 1992, Rev. Mod. Phys. 64, 491. Mizusaki, T., T. Otsuka, M. Honma, and B. A. Brown, 2002, Otsuka, T., R. Fujimoto, Y. Utsuno, B. A. Brown, M. Honma, Nucl. Phys. A 704, 190c. and T. Mizusaki, 2001a, Phys. Rev. Lett. 87, 082502. Mizusaki, T., T. Otsuka, Y. Utsuno, M. Honma, and T. Sebe, Otsuka, T., and N. Fukunishi, 1996, Phys. Rep. 264, 297. 1999, Phys. Rev. C 59, 1846. Otsuka, T., M. Honma, and T. Mizusaki, 1998, Phys. Rev. Mon, K. K., and J. B. French, 1975, Ann. Phys. (NY) 95, 90. Lett. 81, 1588. Motobayashi, T., Y. Ikeda, Y. Ando, K. Ieki, M. Inoue, Otsuka, T., M. Honma, T. Mizusaki, N. Shimizu, and Y. Ut- N. Iwasa, T. Kikuchi, M. Kurokawa, S. Moriya, S. Ogawa, suno, 2001b, Prog. Part. Nucl. Phys. 47, 319. H. Murakami, S. Shimoura, et al., 1995, Phys. Lett. B 346, Palmer, C. W. P., 1984, J. Phys. B 25, 2197. 9. Pandharipande, V. R., I. Sick, and P. K. A. deWitt Huberts, Mulhall, D., A. Volya, and V. Zelevinsky, 2000, Phys. Rev. 1997, Rev. Mod. Phys. 69, 981. Lett. 85, 4916. Papenbrock, T., and D. J. Dean, 2003, Phys. Rev. C 67, Nakada, H., and Y. Alhassid, 1997, Phys. Rev. Lett. 79, 2939. 051303. Navin, A., D. W. Anthony, T. Aumann, T. Baumann, B. D, Pasquini, E., 1976, Ph.D. thesis, Universit´eLouis Pasteur, Y. Blumenfeld, B. A. Brown, T. Glasmacher, P. G. Hansen, Strasbourg, report CRN/PT 76-14. R. W. Ibbotson, P. A. Lofy, V. Maddalena, et al., 2000, Pasquini, E., and A. P. Zuker, 1978, in Physics of Medium Phys. Rev. Lett. 85, 266. light nuclei, Florence 1977, edited by P. Blasi (Editrice Navr´atil, P., and B. R. Barrett, 1996, Phys. Rev. C 54, 2986. Compositrice, Bologna). Navr´atil, P., and B. R. Barrett, 1998, Phys. Rev. C 57, 562. Peierls, R. E., and J. Yoccoz, 1957, Proc. Phys. Soc. (London) Navr´atil, P., and B. R. Barrett, 1999, Phys. Rev. C 59, 1906. A70, 381. Navr´atil, P., G. P. Kamuntaviˇcious, and B. R. Barrett, 2000, P´eru, S., M. Girod, and J. F. Berger, 2000, Eur. Phys. J. A Phys. Rev. C 61, 044001. 9, 35. Navr´atil, P., and W. E. Ormand, 2002, Phys. Rev. Lett. 88, Petrovici, A., K. W. Schmid, A. Faessler, J. H. Hamilton, and 152502. A. V. Ramayy, 1999, Prog. Part. Nucl. Phys. 43, 485. Navr´atil, P., W. E. Ormand, J. P. Vary, and B. R. Barrett, Pfeiffer, B., K.-L. Kratz, F.-K. Thielemann, and W. B. Wal- 2001, Phys. Rev. Lett. 87, 172502. ters, 2001, Nucl. Phys. A 693, 282. Navr´atil, P., J. P. Vary, and B. R. Barrett, 2000a, Phys. Rev. Pieper, S. C., V. R. Pandharipande, R. B. Wiringa, and Lett. 84, 5728. J. Carlson, 2001, Phys. Rev. C 64, 014001. Navr´atil, P., J. P. Vary, and B. R. Barrett, 2000b, Phys. Rev. Pieper, S. C., K. Varga, and R. B. Wiringa, 2002, Phys. Rev. C 62, 054311. C 66, 044310. Nilsson, S. G., 1955, Kgl. Dan. Vid. Sel. Mat. Fys. Med. 29, Pieper, S. C., and R. B. Wiringa, 2001, Ann. Rev. Nucl. Part. 16. Sci. 51, 53. Novoselsky, A., and M. Valli`eres, 1997, Drexel university shell Porter, C. E. (ed.), 1965, Statistical theories of spectra: Fluc- model code. tuations (Academic Press, New York). Novoselsky, A., M. Valli`eres, and O.La’adan, 1997, Phys. Rev. Poves, A., and G. Mart´ınez-Pinedo, 1998, Phys. Lett. B 430, Lett. 79, 4341. 203. Novoselsky, A., M. Valli`eres, and O.La’adan, 1998a, Phys. Poves, A., and J. Retamosa, 1987, Phys. Lett. B 184, 311. Rev. C 57, 19. Poves, A., and J. Retamosa, 1994, Nucl. Phys. A 571, 221. Novoselsky, A., M. Valli`eres, and O.La’adan, 1998b, Phys. Poves, A., and J. S´anchez Solano, 1998, Phys. Rev. C 58, 179. Rev. Lett. 81, 5955. Poves, A., J. S´anchez-Solano, E. Caurier, and F. Nowacki, Nowacki, F., 1996, Ph.D. thesis, Ires, Strasbourg. 2001, Nucl. Phys. A 694, 157. Nummela, S., P. Baumann, E. Caurier, , P. Dessagne, A. Joki- Poves, A., and A. P. Zuker, 1981a, Phys. Rep. 71, 141. nen, A. Knipper, G. Le Scornet, C. Mieh´e, F. Nowacki, Poves, A., and A. P. Zuker, 1981b, Phys. Rep. 70, 235. M. Oinonen, Z. Radivojevic, et al., 2001a, Phys. Rev. C Pritychenko, B. V., T. Glasmacher, P. D. Cottle, M. Fauer- 63, 044316. bach, R. W. Ibbotson, K. W. Kemper, V. Maddalena, Nummela, S., F. Nowacki, P. Baumann, E. Caurier, J. Ced- A. Navin, R. Ronningen, A. Sakharuk, H. Scheit, and V. G. erk¨all, S. Courtin, P. Dessagne, A. Jokinen, A. Knipper, Zelevinsky, 1999, Phys. Lett. B 461, 322. G. Le Scornet, L. G. Lyapin, C. Mieh´e, et al., 2001b, Phys. Pudliner, B. S., V. R. Pandharipande, J. Carlson, S. Pieper, Rev. C 64, 054313. and R. W. Wiringa, 1997, Phys. Rev. C 56, 1720. Oda, T., M. Hino, K. Muto, M. Takahara, and K. Sato, 1994, Radha, P. B., D. J. Dean, S. E. Koonin, T. T. S. Kuo, K. Lan- At. Data Nucl. Data Tables 56, 231. ganke, A. Poves, J. Retamosa, and P. Vogel, 1996, Phys. O’Leary, C. D., M. A. Bentley, D. E. Appelbe, D. M. Cullen, Rev. Lett. 76, 2642. S. Ert¨urk, R. A. Bark, A. Maj, and T. Saitoh, 1997, Phys. Raju, R. D. R., J. P. Draayer, and K. T. Hecht, 1973, Nucl. Rev. Lett. 79, 4349. Phys. A A202, 433. O’Leary, C. D., M. A. Bentley, S. M. Lenzi, G. Mart´ınez- Rakers, S., C. B¨aumer, D. Frekers, R. Schmidt, A. M. van Pinedo, D. D. Warner, A. M. Bruce, J. A. Cameron, M. P. den Berg, V. M. H. M. N. Harakeh, M. A. de Huu, H. J. Carpenter, C. N. Davids, P. Fallon, L. Frankland, W. Gel- W¨ortche, D. De Frenne, M. Hagemann, J. Heyse, E. Jacobs, letly, et al., 2002, Phys. Lett. B 525, 49. et al., 2002, Phys. Rev. C 65, 044323. Ormand, E. W., and C. W. Johnson, 2002, Redstick code. Raman, S., C. W. Nestor, S. Kahane, and K. H. Bhatt, 1989, Ormand, W. E., 1997, Phys. Rev. C 55, 2407. At. Data. Nucl. Data Tables 42, 1. Ormand, W. E., and B. A. Brown, 1995, Phys. Rev. C 52, Rapaport, J., T. Taddeucci, T. P. Welch, C. Gaarde, 2455. J. Larsen, D. J. Horen, E. Sugarbaker, P. Koncz, C. C. Ormand, W. E., P. M. Pizzochero, P. F. Bortignon, and R. A. Foster, C. D. Goodman, C. A. Goulding, and T. Master- Broglia, 1995, Phys. Lett. B 345, 343. son, 1983, Nucl. Phys. A 410, 371. 63

Rauscher, T., and F.-K. Thielemann, 2000, At. Data Nucl. Sorlin, O., S. Leenhart, C. Donzaud, J. Duprat, F. Azaiez, Data Tables 75, 1. F. Nowacki, H. Grawe, Z. Dombr´adi, F. Amorini, A. Astier, Retamosa, J., E. Caurier, and F. Nowacki, 1995, Phys. Rev. D. Bailborodin, M. Belleguic, et al., 2002, Phys. Rev. Lett. C 51, 371. 88, 092501. Retamosa, J., E. Caurier, F. Nowacki, and A. Poves, 1997, Speidel, K. H., R. Ernst, O. Kenn, J. Gerber, P. Maier- Phys. Rev. C 55, 1266. Komor, N. Benczer-Koller, G. Kumbartzki, L. Zamick, Richter, A., 1995, Prog. Part. Nucl. Phys. 34, 261. M. S. Fayache, and Y. Y. Sharon, 2000, Phys. Rev. C 62, Richter, W. A., M. J. van der Merwe, R. E. Julies, and B. A. 031301R. Brown, 1991, Nucl. Phys. A 523, 325. Stoks, V. G. J., R. A. M. Klomp, M. C. M. Rentmeester, and Rodriguez-Guzm´an, R., J. L. Egido, and L. M. Robledo, 2000, J. J. de Swart, 1993, Phys. Rev. C 48, 792. Phys. Lett. B 474, 15. Stoks, V. G. J., R. A. M. Klomp, C. P. F. Terheggen, and R¨onnqvist, T., H. Cond´e, N. Olsson, E. Ramstr¨om, R. Zorro, J. J. de Swart, 1994, Phys. Rev. C 49, 2950. J. Blomgren, A. H˚akansson, A. Ringbom, G. Tibell, O. Jon- Storm, M. H., A. Watt, and R. R. Whitehead, 1983, J. Phys. sson, L. Nilsson, P. U. Renberg, et al., 1993, Nucl. Phys. A G 9, L165. 563, 225. Suhonen, J., and O. Civitarese, 1998, Phys. Rep. 300, 123. Rudolph, D., C. Baktash, M. J. Brinkman, E. Caurier, M. De- Sur, B., E. B. Norman, K. T. Lesko, E. Browne, and R.-M. vlin, J. Dobaczewski, P. H. Heenen, H. Q. Jin, D. R. Larimer, 1990, Phys. Rev. C 42, 573. LaFosse, W. Nazarewicz, F. Nowacki, A. Poves, et al., 1999, Suzuki, K., 1982, Prog. Theor. Phys. 68, 246. Phys. Rev. Lett. 82, 3763. Suzuki, K., and S. Y. Lee, 1980, Prog. Theor. Phys. 64, 2091. Rutsgi, M. L., H. W. Kung, R. Raj, R. A. Nisley, and H. H. Suzuki, T., R. Fujimoto, and T. Otsuka, 2003, Phys. Rev. C Hull, 1971, Phys. Rev. C 4, 874. 67, 044302. Sagawa, H., B. A. Brown, and H. Esbensen, 1993, Phys. Lett. Suzuki, T., and T. Otsuka, 1994, Phys. Rev. C 50, R555. B 309, 1. Svensson, C. E., S. M. Lenzi, D. R. Napoli, A. Poves, C. A. Ur, Sakurai, H., S. M. Lukyanov, M. Notani, N. Anoi, D. Bazzacco, F. Brandolini, J. A. Cameron, G. de Angelis, D. Beaumel, N. Fukuda, M. Hirai, E. Ideguchi, N. Imai, A. Gadea, D. S. Haslip, S. Lunardi, et al., 1998, Phys. Rev. M. Ishihara, H. Iwasaki, T. Kubo, et al., 1999, Phys. Lett. C 58, 2621R. B 448, 180. Svensson, C. E., A. O. Macchiavelli, A. Juodagalvis, A. Poves, Sampaio, J. M., K. Langanke, and G. Mart´ınez-Pinedo, 2001, I. Ragnarsson, S. Aberg,˚ D. E. Appelbe, R. A. F. Austin, Phys. Lett. B 511, 11. C. Baktash, G. C. Ball, M. P. Carpenter, E. Caurier, et al., Sampaio, J. M., K. Langanke, G. Mart´ınez-Pinedo, and D. J. 2000, Phys. Rev. Lett. 85, 2693. Dean, 2002, Phys. Lett. B 529, 19. Svensson, C. E., A. O. Macchiavelli, A. Juodagalvis, A. Poves, Sarazin, F., H. Savajols, W. Mittig, F. Nowacki, N. O. Orr, I. Ragnarsson, S. Aberg,˚ D. E. Appelbe, R. A. F. Austin, Z. Ren, P. Roussel-Chomaz, G. Auger, D. Bailborodin, C. Baktash, G. C. Ball, M. P. Carpenter, E. Caurier, et al., A. V. Belozyorov, C. Borcea, E. Caurier, et al., 2000, Phys. 2001, Phys. Rev. C 63, 061301. Rev. Lett. 84, 5062. Takasugi, E., 1981, Phys. Lett. B 103, 219. Sawicka, M., J. Daugas, H. Grawe, S. Cwiok, D. Balaban- Talmi, I., and I. Unna, 1960, Annu. Rev. Nucl. Sci. 10, 353. ski, R. Beraud, C. Bingham, C. Borcea, M. L. Commara, Tanihata, I., H. Hamagaki, O. Hasimoto, N. Yoshikawa, G. de France, G. Georgiev, M. Gorska, et al., 2003, Eur. K. Sugimoto, O. Yamakawa, T. K. Kobayashi, and N. Taka- Phys. J. A 16, 51. hashi, 1985, Phys. Rev. Lett. 55, 2676. Schmid, K. W., 2001, Prog. Part. Nucl. Phys. 46, 145. Thibault, C., R. Klapisch, C. Rigaud, A. M. Poskancer, Schopper, H. W., 1966, Weak Interactions and Nuclear Beta R. Prieels, L. Lesard, and W. Reisdorf, 1975, Phys. Rev. C Decay (North-Holland Publishing Company). 12, 644. Serot, B. D., and J. D. Walecka, 1986, Adv. Nucl. Phys. 16, Toivanen, J., E. Kolbe, K. Langanke, G. Mart´ınez-Pinedo, 1. and P. Vogel, 2001, Nucl. Phys. A 694, 395. Shimizu, N., T. Otsuka, T. Mizusaki, and M. Honma, 2001, Tonev, D., P. Petkov, A. Dewald, T. Klug, P. von Brentano, Phys. Rev. Lett. 86, 1171. W. Andrejtscheff, S. M. Lenzi, D. R. Napoli, N. Marginean, Siiskonen, T., P. O. Lippas, and J. Rikowska, 1999, Phys. F. Brandolini, C. A. Ur, M. Axiotis, et al., 2002, Phys. Rev. Rev. C 60, 034312. C 65, 034314. Simon, H., D. Aleksandrov, T. Aumann, L. Axelsson, T. Bau- Towner, I. S., and J. C. Hardy, 2002, Phys. Rev. C 66, 035501. mann, M. J. G. Borge, L. V. Chulkov, R. Collatz, J. Cub, Tsuboi, T., K. Muto, and H. Horie, 1984, Phys. Lett. B 143, W. Dostal, B. E. T. W. Elze, and H. Emling, 1999, Phys. 2122. Rev. Lett. 83, 496. Ur, C. A., D. Bucurescu, S. M. Lenzi, G. Mart´ınez-Pinedo, Sohler, D., Z. Dombr´adi, J. Tim´ar, O. Sorlin, F. Azaiez, D. R. Napoli, D. Bazzaco, F. Brandolini, D. M. Brink, J. A. F. Amorini, M. Belleguic, C. Bourgeois, C. Donzaud, Cameron, E. Caurier, G. de Angelis, M. de Poli, et al., 1998, J. Duprat, D. Guillemaud-Mueller, F. Ibrahim, et al., 2002, Phys. Rev. C 58, 3163. Phys. Rev. C 66, 054302. Utsuno, Y., T. Otsuka, T. Mizuzaki, and M. Honma, 1999, Sorlin, O., C. Donzaud, L. Axelsson, M. Belleguic, R. Be- Phys. Rev. C 60, 054315. raud, C. Borcea, G. Canchel, E. Chabanat, J. M. Dau- Vargas, C. E., J. G. Hirsch, and J. P. Draayer, 2001, Nucl. gas, A. Emsallem, M. Girod, D. Guillemaud-Mueller, et al., Phys. A 690, 409. 2000, Nucl. Phys. A 669, 351. Vargas, C. E., J. G. Hirsch, and J. P. Draayer, 2002a, Nucl. Sorlin, O., C. Donzaud, F. Nowacki, J. C. Ang´elique, Phys. A 697, 655. F. Azaiez, C. Bourgeois, V. Chiste, S. Gr´evy, Vargas, C. E., J. G. Hirsch, and J. P. Draayer, 2002b, Phys. D. Guillemaud-Mueller, F. Ibrahim, K.-L. Kratz, M. Le- Rev. C 66, 064309. witowitcz, et al., 2003, Eur. Phys. J. A 16, 55. Vautherin, D., and D. M. Brink, 1972, Phys. Rev. C 5. 64

Vel´azquez, V., and A. P. Zuker, 2002, Phys. Rev. Lett. 88, 1995, Phys. Rev. C 51, 1144. 072502. Wiringa, R. B., and S. C. Pieper, 2002, Phys. Rev. Lett. 89, Vetterli, M. C., O. Haeusser, R. Abegg, W. P. Alford, 182501. A. Celler, D. Frekers, R. Helmer, R. Henderson, K. H. Wiringa, R. B., S. C. Pieper, J. Carlson, and V. R. Pandhari- Hicks, K. P. Jackson, R. G. Jeppesen, C. A. Miller, et al., pande, 2000, Phys. Rev. C 62, 014001. 1990, Phys. Rev. C 40, 559. Wiringa, R. B., V. G. J. Stoks, and R. Schiavilla, 1995, Phys. Vold, P. B., D. Cline, R. Boyd, H. Clement, W. Alford, and Rev. C 51, 38. J. Kuehner, 1978, Nucl. Phys. A A302, 12. Wuosmaa, A. H., I. Ahmad, S. M. Fischer, J. P. Greene, Volpe, C., N. Auerbach, G. Col`o, T. Suzuki, and N. van Giai, G. Hackman, V. Nanal, G. Savard, J. P. Schiffer, P. Wilt, 2000, Phys. Rev. C 62, 015501. S. M. Austin, B. A. Brown, S. J. Freedman, et al., 1998, von Neumann-Cosel, P., A. Poves, J. Retamosa, and Phys. Rev. Lett. 80, 2085. A. Richter, 1998, Phys. Lett. B 443, 1. Yanasak, N. E., M. E. Wiedenbeck, R. A. Mewaldt, A. J. Wakasa, T., H. Sakai, H. Okamura, H. Otsu, S. Fujita, Davis, A. C. Cummings, J. S. George, R. A. Leske, E. C. S. Ishida, N. Sakamoto, T. Uesaka, Y. Satou, M. Green- Stone, E. R. Christian, T. T. von Rosenvinge, W. R. Binns, field, and K. Hatanaka, 1998, Phys. Lett. B 426, 257. P. L. Hink, et al., 2001, Astrophys. J. 563, 768. Wakasa, T., H. Sakai, H. Okamura, H. Otsu, S. Fujita, Yoneda, K., H. Sakurai, T. Gomi, T. Motobayashi, N. Aoi, S. Ishida, N. Sakamoto, T. Uesaka, Y. Satou, M. B. Green- N. Fukuda, U. Futakami, Z. Gacsi, Y. Higurashi, N. Imai, field, and K. Hatanaka, 1997, Phys. Rev. C 55, 2909. N. Iwasa, H. Iwasaki, et al., 2000, Phys. Lett. B 499, 233. Wallerstein, G., I. Iben, P. Parker, A. M. Boesgaard, G. M. Zaerpoor, K., Y. D. Chan, D. E. DiGregorio, M. R. Hale, A. E. Champagne, C. A. Barnes, F. K¨appeler, V. V. Dragowsky, M. M. Hindi, M. C. P. Isaac, K. S. Krane, R. M. Smith, R. D. Hoffman, F. X. Timmes, C. Sneden, et al., Larimer, A. O. Macchiavelli, R. W. Macleod, P. Miocinovic, 1997, Rev. Mod. Phys. 69, 995. and E. B. Norman, 1997, Phys. Rev. Lett. 79, 4306. Warburton, E. K., J. A. Becker, and B. A. Brown, 1990, Phys. Zamick, L., 1965, Phys. Lett. 19, 580. Rev. C 41, 1147. Zdesenko, Y., 2002, Rev. Mod. Phys. 74, 663. Warburton, E. K., and B. A. Brown, 1991, Phys. Rev. C 43, Zhao, Y. M., and A. Arima, 2001, Phys. Rev. C 64, 041301. 602. Zhao, Y. M., A. Arima, and N. Yoshinaga, 2001, Phys. Rev. Weidenm¨uller, H. A., 1990, Nuc. Phys. A507, 5c. C 66, 034302. White, J. A., S. E. Koonin, and D. J. Dean, 2000, Phys. Rev. Zheng, D. C., B. R. Barrett, L. Jaqua, J. P. Vary, and R. J. C 61, 034303. McCarthy, 1993, Phys. Rev. C 48, 1083. Whitehead, R. R., 1980, in Moment Methods in Many Zheng, D. C., J. P. Vary, and B. R. Barrett, 1994, Phys. Rev. Fermion Systems, edited by B. J. Dalton, S. M. Grimes, C 50, 2841. J. D. Vary, and S. A. Williams (Plenum, New York), p. Zuker, A. P., 1969, Phys. Rev. Lett 23, 983. 235. Zuker, A. P., 1984, in Mathematical and computational meth- Whitehead, R. R., A. Watt, B. J. Cole, and I. Morrison, 1977, ods in Nuclear Physics, edited by J. S. Dehesa, J. M. G. in Adv. Nuc. Phys., volume 9, p. 123. G´omez, and A. Polls (Springer, Berlin), volume 209 of Lec- Wildenthal, B. H., 1984, Prog. Part. Nucl. Phys. 11, 5. ture Notes in Physics, p. 157. Wildenthal, B. H., and W. Chung, 1979, Phys. Rev. C 19, Zuker, A. P., 2001, Phys. Rev. C 64, 021303(R). 164. Zuker, A. P., 2003, Phys. Rev. Lett. 90, 042502. Wildenthal, B. H., M. S. Curtin, and B. A. Brown, 1983, Zuker, A. P., B. Buck, and J. B. McGrory, 1968, Phys. Rev. Phys. Rev. C 28, 1343. Lett. 21, 39. Wilkinson, D. H., 2002a, Nucl. Inst. & Meth. in Phys. Res. A Zuker, A. P., B. Buck, and J. B. McGrory, 1969, Shell Model 495, 65. Wavefunctions for A=15-18 Nuclei, Technical Report PD- Wilkinson, D. H., 2002b, Nucl. Inst. & Meth. in Phys. Res. A 99 BNL 14085, Brookhaven National Laboratory. 488, 654. Zuker, A. P., S. Lenzi, G. Mart´ınez-Pinedo, and A. Poves, Wilkinson, D. H., and B. E. F. Macefield, 1974, Nucl. Phys. 2002, Phys. Rev. Lett. 89, 142502. A 58(1974). Zuker, A. P., J. Retamosa, A. Poves, and E. Caurier, 1995, Wilkinson, J. H., 1965, The algebraic eigenvalue problem Phys. Rev. C 52, R1741. (Clarendon Press, Oxford). Zuker, A. P., L. Waha-Ndeuna, F. Nowacki, and E. Caurier, Williams, A. L., W. P. Alford, E. Brash, B. A. Brown, 2001, Phys. Rev. C 64, 021304(R). S. Burzynski, H. T. Fortune, O. Haeusser, R. Helmer, R. Henderson, P. P. Hui, K. P. Jackson, B. Larson, et al.,