RFCasten NuclearStructureFromASimplePerspective PDF
RFCasten NuclearStructureFromASimplePerspective PDF
RFCasten NuclearStructureFromASimplePerspective PDF
A Simple Perspective
R. F. Casten
R. F. Casten
Brookhaven National Laboratory
Institutfur Kernphysik, University ofKoln
246897531
Printed in the United States of America
on acid-free paper
PREFACE
This is a very personal book, reflecting the way I see and understand nuclear
physics and my belief that if something cannot be explained simply it is not
really understood, nor will it be a fertile inspiration for new ideas. It is quite
different from existing texts. It is in no way intended to replace them, but aims
instead to complement the standard approaches to nuclear structure.
The idea for the book and much of its approach arose gradually out of both
formal and informal courses in nuclear structure I have given at the Institute
Laue-Langevin in Grenoble, France, the Institut fur Kernphysik in Koln, Ger-
many, and at Drexel University, Philadelphia, as well as from a series of very
informal tutorial-like sessions with several of my graduate students. The book
represents an attempt to cut through the often heavy mathematical formalism
of nuclear structure and to present the underlying physics of some pivotal
models in a simple way that frequently emphasizes semi-classical pictures of
nuclear and nucleonic motion and repeatedly exploits a few fundamental
ideas. Such an approach has worked for me. I can only hope others will find
it useful.
The emphasis in this book always centers on seeking the essence of the
physics: rigor is therefore often sacrificed. Of course, rigor is absolutely
essential to a proper development of nuclear models and to precise calcula-
tions. Yet it can also be terrifying with page upon page of complex formalism,
Racah algebra, tensor expansions, and the like. Necessary as this is, especially
to those who will become practitioners of particular models, many readers can
become either discouraged or buried in the formalism. Unfortunately, in
either case, the beauty, elegance, and conceptual economy of nuclear structure
theory is often missed. The complexity necessary in formal treatments at times
obscures rather than illuminates the simple physics at work. Moreover, rigor
can be found elsewhere, in many excellent texts: there is no need for another
book to repeat it. (Several of the best of these texts, such as those by de Shalit
and Talmi, Bohr and Mottelson, Brussaard and Glaudemans, deShalit and
Feshbach, or the recent work by Heyde, are cited in the reference list at the
end. They are indispensable.) What cannot be found so easily, though most in-
dividual parts of it probably exist scattered throughout the literature, is a
systematic attempt to convey a more physically intuitive way of thinking about
nuclear structure and of extracting the essential physics behind the derivations
and models.
vi Preface
I honestly do not know whether this attempt will work. I feel it is worth a try
since it is, in fact, a way of thinking used every day by practicing scientists but
that is seldom presented in formal texts. If successful, it can deepen real
understanding. As T. D. Lee once said (BNL Colloquium, May 1983): "That a
thing is elementary does not mean it cannot be deep." Moreover, since physics
research is fundamentally a creative act depending on insight and imagination
(backed, of course, by hard science—experimental or theoretical) an intuitive
sense of nuclear structure can foster new inspirations and remove much of the
mystery surrounding formal or calculational complexity. Finally, an approach
of this type can be of considerable practical use. My hope in fact is that the
reader will come to appreciate how far one can go in obtaining many results of
detailed calculations almost instantly, essentially "by inspection." It can help,
for example, in anticipating the potential usefulness of a given model, in
spotting errors in calculations, or in estimating the effects of particular pa-
rameter changes. One of the best examples is the famous Nilsson diagram:
nearly all its features, and even the semiquantitative structure of Nilsson wave
functions, can be deduced without calculation. The same applies to much of
the study of residual interactions in the shell model, to collective models, to the
structure of RPA vibrational wave functions, Coriolis coupling, or the IB A. To
facilitate the development of this "sense" of the physics, and to provide contact
between models and real data, concrete examples are almost always discussed
in some detail and compared to exact calculations.
The book is in no way intended as a complete treatise on nuclear structure,
either in overall coverage or within each topic. Other texts are more compre-
hensive. For example, many active areas of modern nuclear physics (e.g.,
relativistic heavy ion physics and quark-gluon plasmas, mesonic and quark
degrees of freedom, or baryon excitations such as delta resonances) are totally
ignored. Nevertheless, the book relates to all of them in that it discusses the
basic models of nuclear structure, which successfully describe virtually all low
energy nuclear phenomena, and which subnucleonic approaches must eventu-
ally reproduce. Just as it is difficult to discuss relativistic effects in nuclei
without first knowing how far a non-relativistic theory can go, it is necessary to
understand how far traditional nuclear structure theory can go if one is to
isolate the effects of quarks, mesons, nucleonic excitations, and the like.
Even within traditional nuclear structure, many areas are bypassed. One
reason for this selectivity is that it allows deeper and more detailed discussions
of the subjects that are treated, discussions that would normally be found only
in specialized monographs. A more practical and honest reason is that the
book reflects that small part of the subject with which I am somewhat familiar
and where I felt I could attempt something that was more than just parroting
existing texts.
Finally, in addition to selectivity in topics there is a selectivity in treatment
of each subject. For many areas, there exist several equivalent approaches.
For example, there are works on residual interactions in the shell model that
barely mention the concept of seniority and others that stress it throughout.
Likewise, high spin phenomena in deformed nuclei can be discussed in terms
Preface vii
of the Nilsson model with Coriolis coupling or with the formalism of the
cranked shell model. In each case, I have used the approach (in these
examples, seniority and the Nilsson-Coriolis concepts, respectively) that makes
the essential physics clearest to me (and, I hope, to the reader) and with which
I feel most comfortable.
This book, while my responsibility, owes much to many people, both to those
who molded my own education or who were important in the development of
the approach I use here, and to those who played specific roles in the book
itself. I cannot mention all but would particularly like to single out, for special
thanks and gratitude, the following:
PARTI INTRODUCTION
1. INTRODUCTION, 3
1.1 Introduction, 3
1.2 The Nuclear Force, 6
1.3 Pauli Principle and Antisymmetrization, 16
1.4 Two-State Mixing, 17
1.5 Multistate Mixing, 22
1.6 Two-State Mixing and Transition Rates, 25
REFERENCES, 365
INDEX, 371
PART I
INTRODUCTION
This page intentionally left blank
1
INTRODUCTION
1.1 Introduction
The atomic nucleus is not a single object but a collection of species ranging
from hydrogen to the actinides, and displaying an unbelievably rich and
fascinating variety of phenomena. The nucleus is extremely small, namely
about 10 ~12 to 10~13 cm in diameter, and can contain up to a couple of hundred
individual protons and neutrons that orbit relative to one another and interact
primarily via the nuclear and Coulomb forces. This system may seem so
complex that little could ever be learned of its detailed structure. Indeed,
many of us involved in research into nuclear structure proclaim loudly and
strenuously that we have barely scratched the surface (both literally and
figuratively, as we shall see) in our understanding of nuclear structure. From
another perspective, however, we have an immense number of facts about
nuclei and we understand an enormous amount, often in great detail, concern-
ing what the individual nucleons do in atomic nuclei, how this leads to the
observed nuclear phenomena, how and why these phenomena change from
nucleus to nucleus, and how certain nucleons interact with each other in the
nuclear medium. We have basic models—the shell model and collective
models—both geometric and algebraic, that provide a framework for our
understanding and that are extremely simple, and yet subtle and refined. It is
only after these models and framework are appreciated that the limitations in
our knowledge become focused and identifiable; the identification of the
problems that persist is a prerequisite to further advancement. In this book we
emphasize the known and understood as a background, map, and guide to the
unknown.
We hope the reader perusing this book will come to appreciate two prin-
ciple facts: namely, the beautiful richness and variety of nuclear physics and the
extent to which we can understand nuclear data and models by invoking a few
extremely basic ideas and drawing upon arguments that are physically trans-
parent and intuitive. We will see that it is possible, in many if not most cases,
to understand the detailed results of complex calculations with an absolute
minimum of formalism and often by inspection. As an example, even such
seemingly complicated results of nuclear models such as the famous Nilsson
diagram and the detailed structure of Nilsson wave functions, or of the
microscopic RPA wave functions of collective vibrations, can be derived,
nearly quantitatively, without any calculation whatsoever.
3
4 Introduction
ing can provide in simplifying the systematics of nuclear data. Following this,
a brief treatment of algebraic models, principally the interacting boson ap-
proximation (IB A) model is given.
In Chapters 7 and 8, the discussion turns to odd mass nuclei, with primary
emphasis on deformed nuclei, and an extensive discussion of the Nilsson
model and its consequences, extensions, and testing via single nucleon transfer
reactions.
Most of the collective models discussed up to this point will have been
phenomenological or macroscopic. We will not discuss detailed microscopic
approaches at length but will introduce such approaches in Chapter 9 since
they provide both the microscopic justification of macroscopic models and a
simple physical picture of collective excitations (especially vibrations) that will
allow the reader to anticipate their detailed structure without calculation.
Finally, in Chapter 10, a few selected experimental techniques for studying
atomic nuclei will be discussed, primarily to give the reader a glimpse of the
richness of experimental probes available, of their differences and of the
different types of nuclear data and structure that they elucidate, and to provide
some simple physical ideas that may be helpful in understanding the "parame-
ters" that govern the design of such experiments and the extraction of informa-
tion from them.
With this discussion of the philosophy and outline of the material to follow,
we turn now to the three "cornerstones" mentioned earlier that are of such
central importance to everything that follows. Many readers know that the
nuclear force is attractive, that the Pauli principle is important, and understand
that residual interactions can mix neighboring states. They might be tempted
to skip over these sections and of course that is their prerogative; indeed, these
pages contain nothing that is new or not to be found elsewhere. However, they
do present a somewhat different perspective and provide a ground and back-
ground for what follows.
Fig. 1.1. Binding energies per nucleon. The solid curve is the result of a typical semi-empirical
mass formula that includes corrections for surface effects, Coulomb repulsion, the Pauli principle,
and pairing effects. Many of these topics will be discussed later on. (Based on Eisberg, 1974.)
• For all practical purposes, the nuclear force can be neglected when
considering atomic and molecular phenomena.
These two facts tell us that the nuclear force must be short range. A few
further empirical observations allow us to refine this considerably.
• Nuclear binding energies, per nucleon, at first increase rapidly with A,
until about A ~ 10 to 20, where they level off at approximately 8.5 MeV
and remain roughly constant thereafter. These binding energies per
nucleon are shown in Figure 1.1.
• The masses of mirror nuclei, which are defined as pairs of nuclei with
interchanged numbers of protons and neutrons, (Z, N\ = (N, Z)2, are
nearly identical, after correcting for the different strengths of the Cou-
lomb interaction in the two nuclei.
• The sequencing, spin parity (/*) values, and excitation energies of excited
states in mirror nuclei are also nearly identical.
• Proton and neutron separation energies, denoted S(p) and S(n), are
defined as the energies required to remove the last proton or neutron to
infinity, and have a characteristic behavior with changing proton number
and neutron number. Typical examples of such separation energies are
8 Introduction
Fig. 1.2. Neutron separation energies near the N = 82 magic number (de Shalit, 1974).
shown in Figs. 1.2 and 1.3 from which it is evident that S(p) decreases
with increasing Z and increases with increasing N while S(n) decreases
with increasing N and increases with increasing Z. That is, each decreases
with an increasing number of the same type of nuclcon and increases with
an increasing number of the olherlype. Although not exactly germane to
the present discussion, we note for later use that the separation energies
also show particularly large and sudden drops at certain special numbers
of protons and neutrons, called magic numbers, namely N or Z = 2,8,20,
40, 50, 82,126. Those at 82 are evident in Figs. 1.2 and 1.3.
Measurements of electron scattering provide abundant evidence of a
nearly constant nuclear density independent of the number of nucleons
A. This, in turn, implies that the nuclear volume must increase linearly
with A. Neither of these facts may seem particularly surprising at first
but it should be recalled that such is not the case with atomic systems
whose sizes are nearly independent of Z. Note that if the nuclear volume
V °t A, then, assuming a roughly spherical nucleus, the nuclear radius
scales as Am. Innumerable studies have shown that a good approxima-
tion to the nuclear radius is R = R Aw where R -1.2 fin.
Introduction 9
Fig. 1.3. Proton separation energies near the Z = 82 magic number (de Shalit, 1974).
• There is only one bound state of the deuteron, the simplest nuclear
system, with one proton and one neutron.
• This bound state has total angular momentum 7 = 1 , that is, in the
deuteron, the intrinsic spins (1/2) of the neutron and proton are aligned
parallel to each other, not antiparallel. (This result assumes that the two
nucleons have no relative orbital angular momentum, a result that will be
justified in Chapter 3.)
• The deuteron has a nonzero quadrupole moment, that is, it has, on
average, a preference for a nonspherical shape.
Let us consider each of these facts in turn and see what we can learn about
the nuclear force. Note that, except for the listing of the magic numbers, the
10 Introduction
Fig. 1.5. Level schemes and binding energies (inset box at right with binding energies in MeV
relative to 27A1) of the mirror nuclei27A1 and 27Si, as well as the isobar "Mg.
systems. However, the interaction in the T = 0 p-n system need not be the
same. (As we will see, it is considerably stronger.) Thus there is no a priori
reason to expect that 27 Mg, with fewer p-n interactions, should have the same
sequence of energy levels as 27A1 or 27Si and, indeed, it does not. Furthermore,
27
Mg is less bound than 27A1, even though it has fewer protons and might,
therefore, be expected to be more tightly bound. The reason is that it also has
fewer T = 0 p-n interactions (see Table 1.1). This is already one piece of
evidence suggesting the T - 0 interaction is stronger than the T = I . The
concept of the different and stronger, p-n interaction in the T = 0 channel will
be of enormous importance later. For example, it determines why the excita-
tion spectra of odd-odd nuclei differ so substantially from those of even-even
nuclei. Moreover, its effects are intimately connected with those of the Pauli
principle since T= 1 corresponds to a symmetric alignment of the two isospins
in the p-n system, while T = 0 corresponds to an antisymmetric alignment.
Nuclear separation energies provide crucial information about the outer-
most nucleons and therefore about certain subtle aspects of the nuclear force
in the "valence" region. As we will see, the most important nuclear model, the
shell model, treats nuclei in terms of individual nucleons that orbit as inde-
pendent particles in a central potential. Each orbit carries certain quantum
numbers and a specific wave function. This is an excellent approximation of
the actual motion except that there are important "residual interactions"
beyond those encompassed by the central potential that must be considered
when dealing with nuclei containing several particles outside closed or magic
configurations. This will be a major topic of discussion in Chapters 4 and 5. We
showed examples of separation energy data in Figs. 1.2 and 1.3 earlier and
summarized the trends, which are valid for all mass regions: that S(p) de-
creases with increasing Z, that S(n) decreases with increasing TV, and that each
increases with increasing number of particles of the other type. Superimposed
on this general behavior is a fine structure in that S(p) and S(n) display
odd-even oscillation in Z and N such that nuclei with even numbers of either
protons or neutrons have larger separation energies (i.e., are more bound).
Though these separation energy data are widely familiar, it is seldom appreci-
ated how much they tell us about the nuclear force.
The separation energies refer to the ground states of their respective nuclei:
in nuclei with even numbers of protons and neutrons the ground state always
has spin and parity JK = 0+. Invariably, this state is much lower in energy than
any intrinsic excitation. The fact that S(p) and S(n) are larger when Z and N
14 Introduction
are even thus implies that there is a special attractive interaction in pairs of
protons or neutrons coupled to JK = 0+. Later, we shall see that this is a property
of short-range interactions resulting from the Pauli principle. The separation
energy data also shows that the p-n interaction is strong and attractive since
S(p) increases with increasing N and S(n) increases with increasing Z. In
contrast, the decrease of each separation energy with increasing numbers of
nucleons of the same type gives the fundamentally critical result that, aside
from the pairing interaction, the residual interaction between like nucleons is
repulsive. This fact, pointed out in the early 1960s by Talmi, is seldom
recognized or remembered; however, its consequences, are profound. For
example, anticipating some concepts and jargon from upcoming chapters, it is
one reason why singly magic nuclei do not become deformed and why the
accumulation of proton neutron interaction strength is essential for the onset
of collectivity and deformation in nuclei.
The properties of the simplest bound nuclear system, the deuteron, tell us
still more about the nuclear force. The essential features, summarized earlier,
are that there is only one bound state, that it has J* = 1+, and that the deuteron
has a finite quadrupole moment. The fact that there is only one bound state
and, moreover, that it is only weakly bound (the deuteron binding energy is
2.23 MeV) serves to emphasize the essential weakness of the so-called strong
nuclear force. By weak here we mean weak in comparison to the kinetic
energy of relative motion of the two nucleons. For example, this implies that
the relative kinetic energy of two nucleons cannot be changed substantially by
the strong interaction. This will be important in Chapter 3, when we discuss
the reason why essentially independent particle motion is possible in a densely
packed, strongly interacting nuclear medium.
It will also be shown in Chapter 3 that for a rather general central potential,
the lowest energy state corresponds to zero orbital angular momentum (an S
state). Thus, both the proton and the neutron in the deuteron must be in S
orbital angular momentum states and the total angular momentum in the
ground state can arise only from the proton and neutron intrinsic spins, 1/2 h
(henceforth, in this book, we shall generally omit the units ft in referring to
angular momentum and intrinsic spin). There are two possible ways of
coupling these two spins: to a total spin S = 0 or 1. The fact that the deuteron
chooses the latter highlights an essential point: even though the nuclear force
may have no explicit spin dependence, there can be large energy differences
between states of different spins in multiparticle configurations. We shall
discuss this point extensively in Chapter 4 where we shall see that the implicit
spin dependence of nuclear forces is a reflection of the Pauli principle and that
this has critical nuclear structure consequences.
Finally, the nonzero quadrupole moment of the deuteron is our first indica-
tion of the tendency of the proton-neutron interaction to lead to nonspherical
nuclear shapes. Moreover, it is an indication that the nuclear force cannot be
completely described by a spherically symmetric central potential. (In the
particular case of the deuteron the finite quadrupole moment is evidence for
tensor forces that couple a spin dependent component to a central potential,
Introduction 15
where m is the pion mass. The distance they can travel in this period is
Obviously, y^n) = 0 f°r ri = r2 and thus satisfies the Pauli principle. But, for
any rn, it also follows that
Introduction 17
So, the Pauli principle can be formulated mathematically by the statement that
a two-particle nuclear wave function *P must be antisymmetric with respect to
the interchange of the two partners. For multiparticle states, the antisymmetry
must extend to interchanges of any pair of particles.
Although the present argument was phrased in terms of spatial coordinates,
it can be extended to other spaces leading to the generalized antisymmetriza-
tion condition given earlier.
It is impossible to overemphasize the importance of the Pauli principle in
nuclear physics. It has obvious and direct consequences as well as subtle,
indirect, but no less real, effects. We shall encounter it continually.
The final wave functions are denoted by Roman numerals, ¥, and *?„ and have
energies £, and Eu. In general, the mixing depends both on the initial
separation and on the matrix element. A large spacing reduces the effect of a
given matrix element. Conversely, even a small matrix element may induce
large mixing if the unperturbed states are close in energy. In order to present
the results so that this two-parameter aspect is circumvented, yielding a single
universal mixing expression valid for any interaction and any initial spacing,
we define the ratio
18 Introduction
INTERACTION-. v
Fig. 1.6. Two-state mixing: definitions and notation.
of the unperturbed energy spacing to the strength of the matrix element. Then
the perturbed energies are
where the + sign is for En and the - sign for Er It follows that the final energy
difference is
or, in units of the unperturbed splitting AEu, the final separation is given by the
simple result
A more useful result is the amount, AEs , by which each energy is shifted by
the interaction. A/^ I is given by
or, again, in units of A£ u , one obtains a result independent of the initial spacing:
Introduction 19
The essential point of Eqs. 1.6 and 1.8 is that both the final energy difference
(in units of A£J and ft are functions only of R, the ratio of the unperturbed
energy splitting to the mixing matrix element. Equations 1.3-1.6 and 1.8 are
universal expressions completely independent of the nature of the interaction
or the initial splitting. The same ratio, R = AEJV always gives the same final
wave functions, energies, and energy shifts (in units of AEu).
These results are so important, and will be referred to, either quantitatively
or qualitatively, so frequently that it is useful to dwell on them. Equations 1.6
and 1.8 are plotted in Fig. 1.7. To illustrate the results and get a feeling for the
hFig. 1.7. Universal two-state mixing curves. The one on the left gives the smaller of the two mixing
amplitudes, ft, while the curves on the right give the energy shift of each level in units of the
unperturbed energy separation. Here the lower curve goes with the upper abscissa scale, while the
upper curve goes with the lower scale.
20 Introduction
Table 1.2. Examples of two-state mixing energy shifts and mixing amplitudes (from Eqs. 1.6
and 1.8). R = AE/V
Specific case: A
R A£/A£B ft V(keV) S ^ '
where E0 is the (common) initial energy. Thus each state is shifted by the
Introduction 21
Fig. 1.8. The two limiting cases of strong and weak mixing.
mixing matrix element. This illustrates the important result that, for any
isolated two-state system, the final separation can never be closer than twice the
mixing matrix element. As trivial as this sounds, it is often forgotten but is
extremely useful. For example, it was one of the early arguments used to
demonstrate that Coriolis matrix elements had to be attenuated: examples of
isolated pairs of 13/2+ states were found that were closer than twice the
calculated Coriolis mixing matrix elements.
In the case of degenerate orbits, it is clear that /? = I/ V2~ = 0.707. Thus a is
also 0.707 and the two states are completely mixed. This is conceptually
obvious since the matrix element is "infinitely" stronger than the initial
separation (i.e., 1/R —> °°). This seemingly trivial result also has profound
consequences. For example, it means that the mixed wave functions for two
initially degenerate states are independent of the strength of the interaction
between them. (This argument will be used in Chapter 6 to show why the wave
functions in the limiting symmetries of the IBA are independent of the
coefficients—parameters—of the Hamiltonian as long as the structure of that
Hamiltonian corresponds to the symmetry involved.)
2. The weak mixing limit corresponds to R » 1 (see Fig. 1.8, right). Equa-
tion 1.8 becomes
Hence,
since AEs is small. Frequently (for example, from measured y-ray branching
ratios) one has empirical information on /Jand therefore Eq. 1.10 (or the exact
Eq. 1.8) can be used to deduce V from the data. Similarly, for R»l, Eq. 1.6
becomes
22 Introduction
An example is useful. Suppose R = 10. Equations 1.10 and 1.12 then give
The exact results are ft = 0.0985 and A£/A£u = 0.0099. In fact, even for R = 4,
Eqs. 1.10 and 1.12 are already quite satisfactory: /? is correct to belter than 10
percent and AEs to 6 percent. Except in the case of rather strong mixing, Eqs.
1.10 and 1.12 thus provide quite accurate (instantaneous) results for iwo-slate
mixing.
There is one other important aspect of two-state mixing. Suppose we
consider two states, 1 and 2, whose energies depend on some nuclear structure
parameter x (as illustrated schematically in Fig. 1.9). For example, x could he
the deformation and the states might be two Nilsson orbits. For some x value,
xctil, the orbits cross. Now suppose that the two levels mix. They can now never
cross since they repel, and can never be closer than twice the mixing matrix
element after mixing. Thus the actual behavior of the mixed states, labeled 1
and II, is as sketched by the solid lines in Figure 1.9. The energies have an
inflection point. However, for x > jccril, the wave function of state I will have a
larger amplitude for unperturbed state 2 than for its own "parent" and vice
versa. Such behavior is very common in structure calculations and is nearly
always an indication of strong mixing. The point of closest approach of the two
curves corresponds to the point where the mixed wave functions contain equal
admixtures of each of the unperturbed states. In fact, from the separation at
this point the mixing matrix element can be derived by inspection, as one-half
the separation. This is another illustration of the usefulness of the limiting case
ofEq. 1.9.
Fig. 1.10. Illustration of two multistate mixing situ ations: (Top) N degenerate levels, all of which
mix by equal matrix elements V; (Bottom) The same, except the initial levels are equally spaced.
these levels to mix with each of the others with equal matrix elements (i.e.,
between all pairs). This idea is illustrated in Fig. 1.10. It is then easy to show
by explicit diagonalization that one state is lowered by (N- 1)V and each of the
other states is raised by one unit in V. The wave function for the lowest state
is totally mixed:
Although this is a clear case of (optimum) multistate mixing, the result for the
lowest eigenvalue is exactly what would result from applying a sequence of
two-state mixing calculations: mixing with each of the other W- 1 degenerate
states lowers this state by V, giving a total lowering of (TV - 1) V.
This feature of one state emerging with special character, low energy, and a
highly coherent wave function, is ultimately the microscopic basis for and
physical idea behind the development of collectivity. Collective states result
from many interactions of simpler (e.g., single particle or two quasi-particle)
24 Introduction
entities, and appear at low energies. As we shall see the RPA approach to the
microscopic generation of collective vibrations is a prime illustration of this
effect. So also is the effect of pairing among 0+ states that leads to the well-
known energy gap in even-even nuclei.
A second case is analogous, except that we lift the initial degeneracy and
consider a set of N equally spaced levels. This situation is depicted in Fig. 1.10
for the case of N = 6. As before, one state is considerably lowered. Of course,
the wave functions are now more complex, and are not of particular interest
here. What is interesting is that the ratio of the lowering of the lowest level in
the nondegenerate (ND) case to the lowering in the degenerate (D) case just
considered
Fig. 1.11. Mullislalc mixing: N degenerate levels in which only "adjacent" levels are mixed w i t h
(equal) matrix elements V.
Introduction 25
Finally, note that in all the multistate mixing cases considered, all of the
components of the lowest lying wave function have the same sign. Though this
result depends on the phase conventions chosen, if consistent conventions are
used for both wave functions and operators, then matrix elements (observ-
ables) will contain coherent, in-phase sums, and can be extremely large. The
wave function has coherence, and such multistate mixing can lead to collectivity
as reflected in enhanced transition rates, cross sections, and the like. Also,
note that the sum of the initial and final energies is the same, as, of course, it
must be. Since these energies appear on the diagonal of the matrix to be
diagonalized, this is equivalent to the formal statement that the trace is
conserved.
The importance and usefulness of the results in this section cannot be
overemphasized. With them, and an understanding or the basically attractive
nature of the nuclear force, and of the effects of the Pauli principle and of
antisymmetrization, it is possible to understand nearly all of the detailed
results of most nuclear model calculations in an extremely simple, intuitive
way that illustrates the underlying physics that is often lost in complex formal-
isms and computations.
Fig. 1.12. tiffed of mixing on allowed (A) and forbidden (I-') /-ray transitions.
26 Introduction
transition is normally forbidden, its strength results from mixing the two 2+
states: a similar argument accounts for the 22+-> 2* transition." At first, this
certainly sounds plausible: if the two 2+ states mix, some of the strength of the
allowed transition should be "distributed" to the forbidden one. Moreover,
the two 2+ states now share some of the same character and should be
interconnected. Let us calculate the actual E2 matrix elements for the mixed
states to see if the preceding conclusions are warranted. Using the notation of
Fig. 1.6 (Roman subscripts for the perturbed wave functions, arabic for the
unperturbed), we have
Since the 22+ —> 2* transition is assumed forbidden, the last two terms vanish
and
The 2n+-> 2j+ transition vanishes in the limit of no mixing (j3 = 0). However, it
is by no means clear that mixing will produce a strong transition. The resulting
matrix element is proportional to the difference in the quadrupole moments of
the two states. If the low-lying levels have nearly the same deformation, as is
likely in a well deformed nucleus, this difference will be very small, and thus
the second conclusion is at best risky. We have worked out this example
explicitly because the error just discussed is widespread and partly because the
derivation just given will be useful later in understanding the microscopic
structure of the ^vibration. The point can be summarized as follows: Consider
two states (of the same spin) one of which has an allowed transition to some
other level while the decay for the other is forbidden. The forbidden transition
becomes finite if the two initial states mix, and its matrix element is propor-
tional to the mixing. However, a forbidden transition between the two unper-
turbed levels becomes finite only if the states mix and if the intrinsic structure
of the two unperturbed states differs in the moment corresponding to the
operator for this transition.
With the background provided in this chapter on the properties of the
Introduction 27
nuclear force, the comments on the Pauli principle and our discussion of two-
and multistate mixing, we can now begin to develop an understanding of the
rich diversity and unity of nuclear phenomena. We start with a cursory survey
of some empirical features and then develop at some length the foundation
models through which we try to understand these data and out of which new
models and extensions arise.
2
THE N U C L E A R L A N D S C A P E
One of the difficulties often faced by the student trying to understand nuclear
models is that he/she cannot fully appreciate many of the truly simple and
beautiful results that emerge from these models because there is no reservoir
of familiar nuclear data to call upon. Therefore, when one derives the spin
sequence for a 5-function interaction between two identical nucleons in the
same orbit, the results are only of mathematical interest if he or she does not
see that this instantly explains the low lying levels of literally dozens of near-
closed shell even-even nuclei. The entire seniority scheme seems nothing but
a labyrinth of Racah algebra when one does not understand how many well-
known facets of nuclear structure are thereby trivially explained. Similarly,
the simplicity and intuitiveness of many of the results of the Nilsson model may
fall on barren ground unless one realizes the vast number of deformed heavy
nuclei that display exactly these properties.
The main purpose of this chapter is to survey the nuclear landscape to
display a few (definitely not all) typical patterns of nuclear spectra as well as
some of the systematic changes in these patterns over sequences of nuclei, so
that the reader will understand the motivation for each model and will benefit
from an empirical context for their characteristic predictions. We will refer to
the figures in this chapter frequently.
While this approach necessitates some repetition later, it allows us to see
exactly what we are trying to explain with these models beforehand, and what
kinds of data characterize atomic nuclei and are the most useful as tests of
various models.
In principle, at this stage the data should be shown "blindly," without
commentary on its meaning or implications. However, the purpose of this
book is not to develop nuclear physics ab initio and, indeed, most readers will
already be familiar with many of the major concepts and terminology. A
"purist" approach here would be needlessly tedious and artificial. In the pages
that follow, we will use many words and concepts freely that will be introduced
formally later on. Those to whom these concepts are unfamiliar should
concentrate simply on absorbing the data with the idea of using this base as a
touchstone later.
When we speak of nuclear data, we are referring to a vast, varied, and rich
reservoir of information about atomic nuclei—from the deuteron to the acti-
nides—obtained by a most diverse array of techniques. The simplest informa-
28
The Nuclear Landscape 29
tion is the mass of atomic nuclei. A more useful form for these is nuclear
binding energies, which focus on the interactions between nucleons in the
nucleus when the masses of the individual nucleons are subtracted. Still more
useful (in many cases) are nucleon separation energies, or the energy required
to remove the last, outermost nucleons from the nucleus. (We have discussed
these already in Chapter 1.) The nucleon separation energies give important
data on the surface regions of nuclei. Later, we will show that the individual
nucleons tend to orbit the nuclear center of mass in discrete shells and that, for
many applications, it is possible to neglect all the underlying shells that are
filled. Therefore, the outermost nucleons are frequently most crucial to
understanding the observed properties of nuclear level schemes.
More detailed nuclear data consists of nuclear level schemes: the energies,
and angular momenta, and parity values (J"), of the ground state and low-lying
excited states. The mirror nuclei shown in Fig. 1.5 were our first encounter
with such schemes. Careful measurements of the y-rays emitted when excited
levels de-excite to lower lying ones are fundamental to both understanding
and constructing nuclear level schemes. The crucial information here is, of
course, the y-ray energies, which help define their placements between nuclear
levels and their absolute and relative intensities, which give direct measures of
nuclear transition matrix elements.
A large amount of data has resulted from the study of scattering processes
of one nucleus on another and direct reaction processes in which two interact-
ing nuclei exchange one or more individual nucleons. Scattering experiments
often use low energy projectiles with long wavelengths comparable in size to
the nucleus itself. In such cases, they provide information on the overall
nuclear shape and on the macroscopic, or collective, excitations of the nucleus
as a whole (e.g., rotations and vibrations). Nuclear reactions, examples of
which are single nucleon transfer reactions such as (d, p) or two-nucleon
transfer reactions such as (p, t), can proceed by a direct process in which indi-
vidual nucleons are inserted into or removed from specific orbits. These
reactions provide detailed and microscopic information on the semi-inde-
pendent particle motion characterizing atomic nuclei.
Heavy ion fusion reactions, which can bring in enormous amounts of energy
and angular momentum and access neutron-deficient nuclei, or /? decay ex-
periments on fission product nuclei, which are extremely neutron rich, provide
valuable sources of information, especially on unstable nuclei.
Of course, this listing of techniques barely touches the surface and high-
lights only a few that are most useful for studying low-energy nuclear structure.
A few of these will be discussed in some detail in Chapter 10. For now, we turn
to the picture of the nuclear landscape they have provided us.
The most basic data for nuclei, of course, is a listing of which nuclei exist.
Such a list is usually presented in an JV-Zplot as in Fig. 2.1, where the hatched
area approximately outlines the stable nuclei. This is the so-called valley of
stability (i.e., the valley of energy vs. Z and N: because nature prefers to
minimize energies, the stable nuclei have the lowest energies). The general
features are that N<= Z for light nuclei, and a preference for a neutron excess
30 Introduction
Fig. 2.1. The nuclear chart showing the path of stable nuclei (crossed hatched), as well as the magic
numbers and midshell deformed regions (circles or circular segments).
in heavier nuclei. We shall see that this pattern is easily explained by combin-
ing the concepts of independent particle motion, the Pauli principle, and the
Coulomb force.
We have already seen some other examples of significant nuclear data,
namely binding energies as a function of A, separation energies S(n) and S(p)
and a couple of examples of level schemes for so-called mirror nuclei. These
Fig. 2.2. Low-lying levels of three doubly magic nuclei. Energies on the right of each line are in
MeV.
The Nuclear Landscape 31
data were shown early on because they provide basic information on the
nuclear force itself, which was already discussed in Chapter 1. Here we are
concerned with a somewhat more detailed and much more extensive survey of
nuclear excitations.
As noted in Chapter 1, there are large gaps in the nucleon separation
energies that define so-called magic numbers, and point to the existence of
nuclear shells, analogous to those in atomic physics. That is, certain magic
numbers of nucleons of a given type correspond to the filling of a set of orbits
constituting a "shell." Additional nucleons must then fill the next higher shell
and are considerably less bound. In Figs. 1.2 and 1.3 we saw such data for the
magic number 82. The complete set of the most important, empirically
observed, magic numbers is 2,8,20,28,50,82, and 126. There are also gaps or
subshell gaps at 40 and 64, especially for protons, which exist for certain
neutron numbers only (see further discussions in Chapters 3 and 6). Figure 2.2
shows examples of nuclei where both the proton and neutron numbers are
magic. Note the extremely high energy of the first excited state, and the
predominance of negative parity states. Figure 2.3 shows the Ca isotopes, two
of which,40|48Ca, are "doubly" magic, while the others are singly magic. The
abrupt change of 2+ energies in 4 2 46Ca compared to 40'48Ca is dramatic evidence
of the difference in their magic structure.
One of the great successes of the independent particle model (Chapter 3) is
the prediction of level sequences for nuclei near closed shells, in particular odd
40 48
Fig. 2.3. Low-lying levels and B(H2) values for the even-even Ca nuclei. Ca and Ca are
doubly magic.
32 Introduction
Fig. 2.4. Low-lying levels of two nuclei, 41Ca and 209Pb, with one nucleon beyond a doubly magic
core, and two other nuclei, 43Ca and 169 Yb, for comparison. Note that the energy scale for Yb is
expanded by a factor of 10. The notations (d, t) and (d, p) indicate which levels arc primarily
populated in these single nucleon transfer reactions. The S(d, p) values given for 209l)b are the
single neutron transfer spectroscopic factors; values near unity indicate nearly pure single-particle
structure.
mass nuclei where, as we shall see, the total angular momenta J (this will often
be colloquially referred to later as the "spin" of the level, although this
nomenclature is clearly inaccurate) of the ground and low-lying excited states
are given by the; values of the orbits into which the last odd nucleon can be
placed. To illustrate this, we show the level schemes of two nuclei (41Ca, 209Pb)
with one particle beyond a doubly magic nucleus in Fig. 2.4. The specific order
and energies of these levels will be easy to understand and predict after we
discuss the independent particle model in the next chapter. At this point, these
sequences appear as unintelligible jumbles.
Of course, most nuclei have more than one valence nucleon. Two examples
are included in the figure. One is 2043Ca23 with three valence neutrons. It is
similar in many ways to 4l Ca but, with three valence nucleons, a proper
treatment requires the study of multinucleon configurations and of the "resid-
ual interactions" occurring among nucleons in the valence shell.
The other is 169Yb, which is far from magic in either protons or neutrons.
This scheme, both in terms of its complexity (only hinted at in the figure) and
its compressed energy scale, sets it completely apart from the other nuclei in
the figure. We will see another example of this type of nucleus in a moment.
In treating multivalence particle nuclei such as these, a number of different
approaches are used. Close to closed shells (e.g., 43Ca), an extension of the
independent particle model that includes residual interactions among the
valence nucleons, has been enormously successful. Further from closed shells
(e.g., 169Yb), nonspherical shapes appear, and a deformed shell model (the
Nilsson model) becomes appropriate. In the case of the shell model, it is often
appropriate to consider a coupling scheme in which each nucleon has a given
total angular momentum /. The coupling of these individuals j values leads to
the final J for the state in question. The energy of such a state clearly depends
on the residual interactions among the nucleons in these orbits. It is certainly
one of the triumphs of the shell model that one can easily derive expressions
for these energies, often without a detailed knowledge of the residual interac-
tion itself, that account reasonably well for a large body of data in both odd and
even nuclei.
Figures 2.5-2.7 show the level spectra of some even mass nuclei in the
general vicinity of closed shells. Typical of such nuclei, all have 0' ground
states, first excited levels with ,/" = 2', and mostly even-parity, low-lying
excitations.
All of these features will emerge later from very general considerations of
34 Introduction
Fig. 2.7. Low-lying levels of the A =130 isobars that show the effect of adding valence nucleons (in
this case neutron holes relative lo N = 82).
ior. Figure 2.9 shows typical vibrational nuclei, especially 118Cd, where the first
excited state is a quadrupole phonon excitation of a basically spherical shape.
At about twice and three times this energy there are groups of states that can
Fig. 2.9. Low-lying levels of some typical, near harmonic vibrational nuclei.
36 Introduction
Fig. 2.10. Low-lying rotational and vibrational levels of some typical deformed rare earth nuclei.
Fig. 2.11. Level scheme of 161Dy. (Left) All levels. (Right) Levels arranged into rotational bands
with Nilsson assignments.
The Nuclear Landscape 37
Fig. 2.12. £2! values for all even-even nuclei (Raman, 1987).
left, are arranged in sequences of single panicle intrinsic (Nilsson) states, each
with a rotational hand built on top of it on the right. Another example of a
nucleus that can be classified in similar manner is 169Yb, which we looked at in
Fig. 2.4.
We note that the rotational bands in 161Dy range in character from several
examples (labeled 5/2-[523], 3/2-[532], 5/2-[512]) with regular spacings that in-
Fig. 2.13. /?2j values (in keV) plotted against ATor the/I = 100region.
38 Introduction
Fig. 2.14. EI+ values (inkeV) against N for the A = 130 region.
crease smoothly with J, to some with highly irregular sequences [e.g., 5/2+[642],
l/2-[521] and l/2-[530] (with missing 1/2- level)]. An acceptable model for odd
mass deformed nuclei must be able to account for both types of behavior. We
will discuss these intrinsic excitations and their connection to rotational mo-
tion in detail in Chapters 7 and 8.
Fig. 2.15. values plotted against N for the nuclei with N > 30.
The Nuclear Landscape 39
A dramatic way to illustrate both the collective behavior of nuclei far from
closed shells and the evolution of structure is to examine a particular property
over extended sequences of nuclei—that is, to examine nuclear systematics.
Three of the most telling data are collected in Figs. 2.12-2.16. Figure 2.12
shows the energies of the first excited 2+ states (2 + levels) in even-even nuclei
throughout the periodic table. Figures 2.13 and 2.14 show a more detailed view
of the same data in two particular regions: nuclei around mass A = 100, and
those near A = 130. As we have seen near closed shells, £2+ is rather high
lying, typically 1 to 2 MeV. In contrast, in collective nuclei, the 2t+ state can be
described as either a vibrational or rotational excitation and occurs at much
Fig. 2.16. B(E2:0+1 -> 2^) values for all even-even nuclei. (Bohr, 1975.)
40 Introduction
often assumed that all transitions are El, Ml, or E2, when assigning lvalues,
if the multipolarities have not been measured.
Since the ground state of even-even nuclei is 0+, the first excited state
(normally /* = 2+) can only decay by electric quadrupole or E2 radiation. Since
we shall see that "collective" effects in low-lying states are quadrupole in
geometric character (most deformed nuclei are prolate shaped), it should not
be surprising that E2 or electric quadrupole radiation is of paramount interest.
The usual quantitative measure of E2 transition strengths is called a B(E2)
value, defined as
in terms of the reduced E2 matrix element between initial and final states.
Figure 2.16 shows the systematics of B(E2: 2* -> Oj + ) values across the
nuclear chart in units of so-called single particle values that roughly corre-
spond to the strength expected if a single nucleon changes orbit. The most
obvious feature of the data is the relatively small values near closed shells and
the enormous ones that occur in midshell regions, such as those near mass 160
and 240. These peaks offer the most dramatic evidence known for nuclear
Fig. 2.17. Systematics of /j and 7vibrational energies in the rare earth region.
42 Introduction
Fig. 2.18. Systematics of some typical B(E2) values and branching ratios relating to rotational and
vibrational excitations in the rare earth nuclei (Casten, 1988a).
The Nuclear Landscape 43
As we will study in detail later, the structural transition involved here is one
from spherical nuclei near closed shells (E^lEZ+ < 2) toward spherical, but
vibrational, nuclei and culminates in a phase transition to strongly deformed
(nonspherical axially symmetric) nuclei whose low-lying states reflect rota-
tional behavior. If we recall the discussion of multistate mixing in Chapter 1
(for example, see Eq. 1.13) we see that both the drop in Ei\ and the rise in
B(E2) values can be understood in terms of the correlations or collectivity that
develop from multistate configuration mixing as one goes from closed shell
regions toward midshell. Many of these ideas will be quantified and specific
later.
Finally, a crucial test of our detailed microscopic understanding of the
macroscopic collective shape vibrations is whether we can understand their
systematics. Figure 2.17 shows the energies of/? and /vibrations in the rare
earth region of deformed nuclei. Figure 2.18 gives some crucial B(E2) values
relating to the lowest states in even-even nuclei and to their low-lying /3 and y
vibrations. The top panel shows B(E2: 1 + -> O/) values that describe the
matrix elements connecting rotational states. (This is a small subset of those in
Fig. 2.16.) The middle and lower panels give the ratios of y-vibrational to
ground state B(E2) values to B(E2: 2+-* 0^), and the ratio of "ft -^ g"
over "y-> g" transition strengths. The intraband values are a couple of orders
of magnitude larger than interband B(E2) values and "y-» g" matrix elements
dominate "/? -> g" values. A successful collective model must account for all
these results.
In closing this chapter it is appropriate to summarize the nuclear landscape
in a compact form that will later allow us to make instant, a priori, estimates of
the likely structure of any given nucleus. We do this by recalling Fig. 2.1, which
shows the nuclear chart in an N-Z plane. The magic numbers are indicated by
vertical and horizontal lines and the known and expected midshell regions of
deformed nuclei are encircled. Much of the rest of this book is devoted to
understanding nuclei of each specific type occurring in the chart as well as the
evolution of structure from one type to another.
This page intentionally left blank
PART II
SHELL M O D E L A N D R E S I D U A L
INTERACTIONS
This page intentionally left blank
3
THE I N D E P E N D E N T PARTICLE M O D E L
We saw in Chapters 1 and 2 that the nuclear force is attractive and short
range, and that the systematics of certain gross nuclear properties, such as
nucleon separation energies, are generally smooth, except at certain specific
nucleon numbers, called magic numbers, where they exhibit discreet jumps.
The concept of magic numbers and the shell structure that they imply is of
paramount importance in nuclear physics. Thus we summarize a bit of the
voluminous evidence of their existence. Beside the sharp drop in S(n) and S(p)
just after magic numbers (see Figs. 1.2 and 1.3), the lowest excited states in
nuclei with magic numbers of either protons or neutrons are, on average,
extremely high lying. In particular, in nuclei with even numbers of protons and
neutrons the energy of the first excited state is nearly always a 2+ state, and is
much higher in magic nuclei. This was illustrated by the Ca isotopes in Fig. 2.3:
40
Ca and 48Ca correspond to the magic numbers at 20 and 28. Across the even
^Sn nuclei that have a magic number of protons, the first excited state (2+) has
an energy £2f ~ 1200 keV (see Fig. 2.6) as opposed to E 2 f = 500 keV for the
isotones of Cd or Te (see Fig. 2.8). Even more striking, when Sn becomes
doubly magic at 132Sn, £ 2 + suddenly jumps to several MeV. Further support for
the idea of magicity stems from the fact that elements with magic proton
numbers have higher relative abundance, a larger number of stable isotopes,
and relatively low neutron absorption cross sections. The nucleosynthesis of
such elements predominantly occurs in stellar supernova explosions where an
intense neutron flux leads to rapid, successive neutron capture reactions. This
is the so-called r-process ("r" for rapid). As we shall see in Chapter 10, the
cross sections for such reactions depend mostly on the level density at excita-
tion energies near the neutron separation energy. Such level densities are
particularly low in magic nuclei. Therefore, for magic nuclei, the low neutron
cross sections imply that, once formed, it is unlikely that a sufficient number of
neutron captures take place in the short-lived astrophysical environment to
deplete their numbers. In essence, they tend to block the r-process path.
Thus, we see several lines of evidence pointing to the importance of magic
numbers. Moreover, we notice a relationship in these lines of evidence; many
stem ultimately from the difficulty of exciting a magic or closed shell structure,
and the consequent low-level density at low excitation energies.
Combining all the evidence, we can summarize the relevant magic numbers
for nuclei as
reproduce is the particular stability of nuclei with these magic numbers. One
would therefore like to construct a nuclear potential that automatically and
naturally produces gaps in single particle level energies at the magic numbers.
It is worth noting an often misunderstood point here: One often hears that
closed shell nuclei are the most stable nuclei. This is not true, however, as a
glance at the chart of separation energies in Figs. 1.2 and 1.3 clearly indicates.
As nucleons of a given type (e.g., neutrons) are added, neutron separation
energies systematically decrease. Just after a closed shell, the separation
energy undergoes a much larger drop. Thus, closed shell nuclei are only more
stable relative to succeeding nuclei.
In considering an appropriate potential for the nuclear case, a tremendous
simplification results if the potential is central, that is, if it depends only on the
radial distance from the origin to a given point. This is equivalent to requiring
that the potential is spherically symmetric. Then, the angular dependence of a
particle wave function is independent of the detailed radial behavior of the
central potential. Moreover, the orbital angular momentum operator, /, com-
mutes with the energy (tf) and is a constant of the motion. All detailed effects
of the particular choice of central potential will therefore show up only in the
radial behavior of the wave functions.
Before considering the specific choice of central potential for the nuclear
case, it is useful to summarize a few general properties of such potentials. We
denote an arbitrary central potential by [/(/•) and only require that U(r) is
attractive and U(r) -> 0 as r —> 0. The Schrodinger equation for such a
potential is
This equation is separable into radial and angular coordinates and therefore
the solutions yT rtm, can be written
n
Here, n is the radial quantum number, / the orbital angular momentum and m
the eigenvalue of its z-component, lf. It is conventional in nuclear physics to
give names to different / values following the convention:
Its solutions have some interesting properties. First, outside the potential, the
wave function decreases exponentially and therefore vanishes as r —> °°. The
quantum number n specifies the number of nodes (zeros) of the wave function
with the usual, but not universal, convention that one counts the node at
infinity but not that at r = 0, that is, n = l, 2,....
50 Shell Model and Residual Interactions
Fig. 3.1. (Top) IlJustration of a single valence nucleon orbiting a doubly magic nucleus. (Bottom)
Schematic illustration of three shell model potentials, a simple harmonic oscillator, a square well,
and an intermediate shape or modified harmonic oscillator. The latter simulates, to some degree,
the effect of an I2 term.
increase with / as well as with the number of nodes (n) in the wave function, it
is clear that, at least qualitatively, one can compensate an increase in n with a
decrease in /. The factor connecting these effects is exactly 2 for harmonic
oscillator potential: it is also 2 for a square well potential.
52 Shell Model and Residual Interactions
Fig. 3.2. Single-particle energies for a simple harmonic oscillator (S.H.O.), a modified harmonic
oscillator with / 2 term, and a realistic shell model potential with / 2 and spin orbit (/ • s) terms.
It is this grouping of levels that provides the shell structure required of any
central potential useful for real nuclei. If we recall that each energy level has
2(21+1) degenerate m states, then, by the Pauli principle, each nl level can con-
tain 2(21 + 1) particles. Therefore, if we imagine filling such a poten-tial well
with fermions, each group or shell can contain, at most, the specific numbers of
particles indicated in the figure. Hence, such a potential automatically gives a
shell structure rather than, say, a uniform distribution of levels.
Unfortunately, except for the lowest few, these shells do not correspond to
the empirical magic numbers. Therefore, while the harmonic oscillator poten-
The Independent Particle Model 53
It is clear that neither of these alternatives yet produces the magic numbers
observed experimentally. It is easy to do so, however, if one introduces a so-
called spin-orbit force. Thus far, we have not discussed the spin quantum
number explicitly. Nevertheless, it is well known that the nucleon, either
proton or neutron, has an intrinsic spin 1/2, and therefore the total angular
momentum of a nucleon in any orbit is given by the vector coupling of the
orbital angular momentum / with a spin angular momentum s = 1/2. With a
spin-orbit component, the force felt by a given particle differs according to
whether its spin and orbital angular momenta are aligned parallel or antiparal-
lel. If the parallel alignment is favored, and if the form of the spin-orbit
potential is taken as V,.t - -Vls (r)l • s so that it affects higher / values more,
then its effects will be similar to those illustrated on the far right in Fig. 3.2.
Each nl level, such as Ig, will now be split into two, Ig9;2 and lg7/2, orbits with
the former lowered and the latter raised in energy. This instantly reproduces
all the known magic numbers.
The absolute strength of the spin orbit force must be substantial (see Fig.
3.2) to produce the correct magic numbers: indeed, the splittings it produces
must be comparable to those between adjacent multiplets of the harmonic
oscillator potential. Since the constant hatoi the harmonic oscillator potential
is found to be h(0 - 41M"3 (e.g., fico^ 8 MeV for medium and heavy nuclei), it
follows that the Vls(r} must attain nearly such magnitudes.
Since the spin-orbit force is an inherently quantum relativistic effect, it is
not as easy to give a physical picture for it as for the relation between an / 2
force and the effective change in the behavior of the central potential just
discussed. It has been shown, however, to arise naturally, and with the correct
sign, from relativistic effects of the nucleonic notion. It is possible, though, to
give plausible arguments for the radial shape of the spin-orbit potential. These
rely again on the notion that, in the interior of the nucleus, a nucleon should
experience no net force. If the spin-orbit force were large in the nuclear
interior there would be a preference for nucleons with spins aligned parallel to
their orbital angular momentum rather than vice versa and therefore such a
nucleon would not be surrounded by an equal number of nucleons with all spin
orientations. This suggests, although it certainly does not prove, that the spin
orbit force is primarily a surface phenomenon. It is therefore customary to
write:
where V(r) is whatever potential is chosen for the central potential itself and
Vls is a strength constant.
It is worth pausing at this point to emphasize the importance of the spin-
orbit interaction. It is not merely a device that ensures the appropriate magic
numbers. Rather, a significant fraction of nuclear structure research in the last
two decades has relied on and exploited the particular consequences of the
spin-orbit force. To see this, it is necessary to refer more explicitly to the
concept of parity. The parity of a wave function \f/nlm = R^r^Y^O, <p) is given
by the behavior, (-1)', of the spherical harmonic Ylm under reflection, since the
The Independent Particle Model 55
radial wave function does not depend on the sign of r. Thus nritm = (-1)'. For a
multiparticle system of Af independent particles,
and the total parity is
and spacings of the single-particle levels. The reader may perhaps think that
we have already dealt with and solved this question by the sequence on the
right of Fig. 3.2. However, this ignores the fact that the single-particle energies
themselves depend on the number of nucleons in the nucleus since the single-
particle potential arises from these same nucleons. Considering nuclei with
one valence particle (beyond closed shells), this means considering the effect
of the closed shells.
The reader has undoubtedly encountered the idea that closed shells can be
ignored. There is considerable truth in this, as we shall see, but it is not the
whole story. Especially in recent years, with the advent of much new data far
off stability, this issue takes on real importance. In the next few paragraphs we
discuss the effects of closed shells on single-particle energies. An extension to
multiparticle configurations will be discussed early in the next chapter.
The basic result is absolutely trivial. Since a closed shell has 7 = 0, its wave
function is spherically symmetric. Therefore, imagine a single valence nucleon
outside this shell in an orbit; and magnetic substate m. Since the closed shell
has no preferred direction in space, its interaction with this nucleon must be
independent of m. This does not mean that the interaction can be ignored. It
can, and does, exist but it is only equivalent to a change in the spherically
symmetric central potential. A particle in a particular valence ;' shell certainly
interacts with the closed; shells below it and its single-particle energy is altered
by that interaction. The preceding argument simply means that this interac-
tion is independent of direction. [Incidentally, to anticipate our later discus-
sion of the multipole expansion of an interaction in Pk(cos&), this shift is due
exclusively to the monopole (k - 0) part of the interaction between the closed
and open shell nucleons, since that is the only multipole that is 9 independent
(P0(cos0) = constant)].
This interaction of an open shell with underlying filled shells has an impor-
tant consequence. A major shell generally consists of several constituent;'
shells. Each of the ;'s of a closed major shell can have a different (spherically
symmetric) effect on each of the valence ;' orbits. Thus, the relative single-
particle energies in a given major shell depend critically on the specific lower-
lying, filled closed shells.
We can illustrate this effect and understand its importance with some
examples. Empirically, it is possible to map out single particle energies with
one-nucleon transfer reactions. These will be discussed extensively for de-
formed nuclei in Chapter 8. Suffice it to say here that in a reaction like (d, p),
illustrated in Fig. 3.3, a neutron is stripped off the incoming deuteron into a
specific orbit around the target nucleus, leaving an outgoing proton. Clearly,
such a reaction can disclose sequences of states with single-particle structure,
each corresponding to different;' orbits. Moreover, once the "kinematical"
aspects of the reaction collision are removed, the cross sections yield a nuclear
matrix element, which is the purity of the single-particle state. This informa-
tion is embodied in the so-called spectroscopic factor: S(d, p) = 1 corresponds
to a pure single-particle neutron wave function coupled to the target nucleus.
While a reaction like (d, p) can only populate a state to the extent that the
60 Shell Model and Residual Interactions
corresponding orbit was empty in the target, "pickup" reactions like (d, t)
extract a neutron from an already filled orbit: they produce hole excitations
and sample single-particle energies below the current valence shell.
Extensive data collected over several decades, much of it from such reac-
tions, has allowed us to map out the empirical single-particle energies in a
number of nuclei that are one particle or hole removed from various major
closed shells. The recent extension of such data to nuclei far off stability has
greatly expanded the overview of single-particle energies thus provided.
We summarize some of the results in Fig. 3.4. Each panel gives the observed
single-particle or hole energies for two nuclei, along with schematic illustra-
tions of the orbits involved. The point is to compare these energies for
different systems, cores, and types of nucleon.
The top left panel of Fig. 3.4 shows proton and neutron single-particle
energies in the 82 to 126 shell extracted from the particle levels of 209Bi and the
hole states of 207Pb. The energies are nearly identical. (The slight expansion of
the proton energy scale is probably a Coulomb effect.) This is reasonable. In
each of the nuclei, the valence particle or hole "feels" interactions with
essentially the same underlying orbits. The only notable difference is that the
neutron holes in 207Pb interact with 43 other neutrons in the 82 to 126 shell,
while the proton particles in 209Bi interact with the full major shell of 44
neutrons. As individual nucleon-nucleon interactions are on the order of a
few hundred ke V, and ^'-dependent differences considerably less, this should be
a minor effect. On the top right panel in Fig. 3.4, a similar situation is shown for
the 50 to 82 shell with identical results.
However, in the bottom panel, the neutron holes in 131Sn interact with the
same 50 neutrons and 40 protons as in Zr but, in addition, with the Z = 40 to 50
closed proton shell and with other neutrons in their own shell. As a conse-
quence, their energies are very different from 91Zr.
This particular example is actually of great structural consequence. (We
The Independent Particle Model 61
Fig. 3.4. Changes in single-particle energies in different nuclei that illustrate the effect of closed
shells on the valence orbits. Since we are interested in relative changes only, one energy is
normalized in each box.
will see later that it accounts for the onset of deformation near A = 100.) The
point is that the interaction of a given shell,;',, with another (closed or open),
jv depends on the overlap of the respective wave functions; the only difference
if j2 is closed is that, then, the interaction is angle independent, so it is the radial
overlap that counts. Orbits with similar quantum numbers nlj have higher
overlaps. This is demonstrated in Fig. 3.5, where the dependence of the radial
overlaps of various orbits with a Is orbit on An = nl - n2 and A/ - l^ - 12 is
illustrated. The falloff with Arc and A/ is clear. (Incidentally, the simple
estimate that the interaction goes roughly as l/(An + A/) is not a bad guide).
In this particular case, as the proton Ig^ orbit fills from Z = 40 to 50 it exerts
a strong attractive pull on the lg7/2 neutron orbit, drastically lowering the
energy as seen so clearly in Fig. 3.4. In contrast, an orbit like 3s1/2 has poor
overlap with lg,J/2 and, relatively, its energy increases.
62 Shell Model and Residual Interactions
Fig. 3.5. Indication of the dependence of a residual interaction on the difference in principle (n)
and orbital angular momentum (/) quantum numbers of the two orbits involved. The contours
give constant values of the radial overlap integrals of a ls1/2 orbit with orbits of different n, I
(Heyde,1987).
Fig. 3.7. Ground state J* values for a number of odd-mass nuclei for comparison with the
predictions of the independent particle model (compare Fig. 3.2).
elevating a proton from below the N = 40 closed shell to pair off with the odd
g,^ proton, leaving a l/2~ hole.)
As a final illustration, if two of the three valence neutrons in 43Ca couple to
J = 0, its excited states, identified in (d, p) or (d, t), should resemble those in
41
Ca. Figure 2.4 confirms this.
From our earlier discussion of Fig. 3.4, it is obvious that deviations from Fig.
3.2 must occur. The bottom row of Fig. 3.7 is a sequence of N = 50, odd Z nuclei
from Z = 35 to 45. From Fig. 3.2 we would expect the 35th and 37th protons to
fill the f5/2 orbit with a low-lying pM excitation, the 39th to occupy the p1/2 orbit
and the 41st to 45th the g9/2 orbits. The predictions for Z = 39 to 45 are
confirmed empirically, but there is an inversion of the fs/2 and pM levels. As
another example, the level order ordinarily associated with the 50 to 82 shell is
a low-lying pair of g7/2, d5/2 orbits with a small gap before a cluster of s1/2, d3/2 and
hn/2 orbits. This pattern is seen clearly in 131Sn in Fig. 3.4. However, this figure
also shows the very different order in 91Zr. While we are able to predict the
d5/2 ground state for Zr, Fig. 3.2 would have suggested a 7/2+ first excited state,
instead of the l/2+ state observed empirically. Indeed, "Zr (with filled d^
orbit) also has a l/2+ ground state. The inversion of the ground and first
excited states of 209Bi is a third example (Compare Figs. 3.2 and 3.6). These are
not deficiencies in the model, just a fact of life in its use. As discussed in the
context of Fig. 3.4, one must deal with changes in single-particle ordering,
especially when the same major shell is inspected in nuclei that are far apart in
mass so that the interactions of the intervening nucleons will have altered the
single particle potential. Also, because of the Columb potential, proton and
neutron single-particle sequences are expected, and found to be, slightly
different, especially in heavy nuclei.
Of course, there are also real discrepancies with the predictions of the
independent particle model. Figures 2.4 and 2.11 showed the examples of
169
Yb and 161Dy, where a totally different empirical picture, completely at
variance with Fig. 3.2, is observed. We will see the reasons for this later.
There is one last prediction of the independent particle model that accounts
nicely for a characteristic aspect of doubly magic nuclei. Clearly, any excita-
tion of a nucleon to the next higher shell nearly always involves excitation into
an orbit with the opposite parity. Therefore, the low-lying excited levels of
such nuclei should be negative parity. This is empirically well known, as
exemplified in Fig. 2.2.
Although we will not discuss it here, there are many other predictions, for
example magnetic moments or single-particle electromagnetic transition rates,
that are reasonably well reproduced by the simple independent particle
approach.
Finally, we caution that, while / = 0+ is the lowest-lying configuration for the
j" configuration of identical nucleons, other / values are possible at higher
energy. Empirically, the next energy level nearly always has J = 2*. Therefore
it is risky to extend the predictions for odd mass nuclei above excitation
energies comparable to Eif in the neighboring even-even nucleus, since con-
figurations of the form j ® 2+, J) in which a particle in orbit nlj is coupled to a
"core" excitation (the first 2* state) of the under-lying even-even nuclei can
then compete with single-particle excitations.
4
THE SHELL MODEL:
TWO-PARTICLE C O N F I G U R A T I O N S
In order to proceed, we must deal more deeply and systematically with the
problem of multinucleon configurations. By this we mean "valence" configu-
rations of two or more particles outside a core, which is usually assumed to
consist of inert closed shells. There are really two issues here: First, which J
values are allowed by the Pauli principle, and, second, what is their relative
energy ordering. The first question can be answered without reference to any
discussion of a central potential or residual interaction, while the second
depends on the details of those interactions, although many properties can be
deduced from the simple fact that the radial and angular parts of a wave
function in a central potential can be separated and that predictions for the
latter are independent of / value, and for the radial part, many results can be
obtained by rather general arguments based on its short- or long-range charac-
ter or by considering a multipole expansion of it. This chapter addresses these
issues. We will start by considering the simplest case of two identical nucleons
in the same or different orbits. Then we will turn briefly to the case of two
nonidentical nucleons (proton and neutron) and the role of isospin and ex-
change terms in the residual interaction. In the next chapter we will consider
the case of larger multinucleon configurations of the form/1.
A complete treatment of many of these issues involves extensive and
sophisticated familiarity with the formalism of angular momentum and tensor
algebra. Indeed, many of the results can only be proved by rather formal
manipulation of the various angular momentum coefficients (3-/ symbols,
Racah coefficients, coefficients of fractional parentage, and so on.) and a deep
understanding of their symmetry properties. It is both contrary to the spirit of
our presentation and entirely beyond the competence of this author to present
this formal material fully. Moreover, there are numerous existing texts that
deal with it at great length in superb fashion. Two of the best are those of de
Shalit and Feshbach and de Shalit and Talmi, to which the reader is encour-
aged to refer. In the following presentation, we will make explicit use of their
results, but our principle effort is to provide a physical understanding and
motivation for the results and to present arguments for their plausibility. The
excellent recent book by Heyde also develops these ideas in a more formal
way. It is complementary to the present treatment and is highly recommended.
Before discussing the particular characteristics of multinucleon configura-
tions under the influence of residual interactions, it is crucial to discuss the role
67
68 Shell Model and Residual Interactions
of closed shells on the valence nucleons. As with the energy levels of single
nucleons in shell model orbits, it is only if the effects of closed shells can be
neglected that the study of multinucleon configurations will be applicable to
other than the lightest nuclei.
Since a closed shell is spherically symmetric, the interaction of a valence
nucleon in the state jm cannot depend on m: hence, the closed shell particles
have the same effect on all valence nucleons in a given; orbit. In multiparticle
configurations, the effects of the closed shell nucleons are independent of the
way (the relative orientations) that the individual jtn values are coupled to the
total J. Therefore, the closed shell can have no effect on energy differences of
these different J states. As discussed in Chapter 3, this does not rule out an
additive energy for the entire group of states and different; values can lead to
different shifts. Thus, the "rule" for spacings refers only to those within a given
configuration, not to the relative energies of different; shells in a major shell.
With these preliminaries in hand, we can now discuss two-particle configu-
rations. First, we need to determine which J values result from coupling of a
nucleon in orbit ;'j with one in orbit ;2. If jl *J2, these are simply the integer
values from l/j-;'J to I/, +; 2 |. If the orbits are equivalent, n,/^ = nj,jz
(which we shall often abbreviate to;'j =j2 when no confusion should arise) we
distinguish two cases— identical and nonidentical nucleons. For the latter
(proton-neutron) case, / takes on all integer values from 0 to 2j, as there is no
Pauli principle restriction on the occupation of identical m states. However,
for identical nucleons, one must explicitly consider the effects of the Pauli
principle, which requires the total wave function to be antisymmetric.
For identical nucleons, the isospin projection Tz = fz (1) + tz (2) = 1 and hence
the total isospin T = 1. (Clearly, T cannot be less than its projection, nor
greater than the sum of the individual tz values.) This is a symmetric wave
function. Hence, the space-spin part must be antisymmetric. When we impose
this requirement, only certain/ values are allowed. To see this, we have to look
at the particular m states occupied by the particles, 1 and 2. The Pauli principle
requires that
Since 2j is odd, this vanishes unless / is even: the only allowed J states for two
The Shell Model: Two-particle Configurations 69
identical fermions in equivalent orbits are those with even total angular
momentum / = 0,2,4,...(2; -1). In Chapter 5, we will see that the same result
is even simpler to obtain in the m-scheme, which provides a general, though
sometimes tedious, way of finding the possible J states for any multiparticle
configuration.
With this result for identical nucleons in hand, we have an alternate way to
look at the complete set of J values 0 - 2j, available in the p-n system. We
consider the isospin structure of this system and use the result just obtained.
The p-n system can have T = 1/2 ± (-1/2) = 0 or 1 and hence T = 0 or 1. The
T=l case is identical in all respects to the preceding p-p and n-n T=l cases
and therefore consists of the even / values 0,2,4,...(2j -1). The T = 0 case is
antisymmetric in isospin so the space-spin part must be symmetric. By a
derivation exactly analogous to that just given, one finds that the T = 0 p-n
system contains the odd /values. Together, the T(p-n) = 0 + 1 parts give ally
values from 0 to 2/. (Note that, for j * jn, each two-particle J state is now a
mixture of T= 0 and T=l parts: a total nuclear wave function of good isospin
is constructed by coupling this two-particle state to the core.)
where the last step utilizes the Wigner-Eckart theorem. Most of the discussion
in this chapter centers on the crucial properties of these energy shifts due to the
residual interactions. Because of the independence of M, we shall generally
omit the magnetic quantum numbers in the discussion that follows. We have
already stated that it is possible to separate the radial and angular coordinates
for many residual interactions. To illustrate this, it is useful to consider the
simple 5-function interaction: by definition, this interaction vanishes unless
the particles occupy the same spatial position. The reason for choosing a 5-
function residual interaction is not simply mathematical convenience. More
importantly, it is preeminently a short-range interaction, and we know that the
70 Shell Model and Residual Interactions
Fig. 4.1. Schematic illustration of the energy shifts in a configuration with particle 1 in orbit j} and
particle 2 in orbit j coupled to various spins J.
nuclear torce, including residual interactions, has just this character. At least
qualitatively, a 5-function residual interaction reproduces many observed
properties of nuclei.
Moreover, it can be shown that a 5-interaction in a/" configuration is equiva-
lent to an odd tensor interaction. Such interactions are diagonal (see Chapter
5), in the so-called seniority scheme and as such are particularly useful for
treating multiparticle configurations, since many important results reduce to
the two-particle case. Thus, a discussion of the 5-interaction in | jjj} configu-
rations has profound implications throughout the study of nuclear structure.
We can write the 5-interaction as
where the second form expresses the interaction in polar coordinates. This is
The Shell Model: Two-particle Configurations 71
particularly useful when separating the angular and radial parts. The l/r^
factor is necessary because the integration over the angular coordinates intro-
duces a factor 4;cr2.
Using the polar coordinate form and performing some straightforward but
tedious angular momentum algebra (see de Shalit and Feshbach, Chapter 5)
we obtain for the energy shifts in the identical particle configuration
where
and
FR depends only on the radial coordinates, while the quantity A results from an
integration over the angular coordinates.
Therefore, one obtains the extremely important result that the relative
splittings depend only on universal angular functions A, which are totally
independent of the nature of the central potential. They are also independent
of the principle quantum numbers nt and n2 (except for an overall scale
incorporated into the factors FR); for this reason, we shall not specify the n
values unless it is necessary in a particular case. Although Eqs. 4.6-4.7 apply
in detail only to a 5-function interaction, the general separability into radial
and angular parts is valid for any residual interaction that depends only on the
separation (r5 - r2) of the two nucleons.
We will soon discuss the meaning and implications of Eq. 4.5 at considerable
length. First, however, it is useful to give the expression for the specific case of
equivalent orbits /^ = ljr Hence /t + /2, must be even. Moreover, proper
normalization of the wave functions introduces a factor of 1/2 in the energies
that are now given by
where
and
Note that for / = 0, A <* (2j + 1)12, so that AE(fJ = 0) = const (2; +1)/2, that is,
72 Shell Model and Residual Interactions
the energy lowering of 0+ states is larger for large / and is in fact proportional
to the number of magnetic substates in the orbit/. This property is identical to
that defined for a pairing interaction. Here, its physical basis is that, for high
/, there are more magnetic substates spanning the same angular range of orbit
orientations. Hence, the wave function for a given substate is more localized
in angle. Two particles with the same I m I value thus have greater overlap and
hence a larger interaction. Similar overlap arguments will help us understand
many of the effects of residual interactions. They will be formalized later and
will pervade this chapter.
Returning to the general case of any jv j2, we note that, while the relative
energies of individual J values depend only on the angular structure of the
wave functions, the overall scale of the interaction and the average interaction
strength for particles in orbits jl and/2 depend on the radial integral (Eq. 4.6).
As we stated in Chapter 3, this will be largest for similar n^ and n2/2 values.
Shortly, we will calculate explicit numerical values of AEQJ^I), or rather, of
A (/v//)- First, however, it will help us understand two-particle configurations
and, later, multiparticle situations, if we momentarily ignore the analytic
formula and try to understand the basic results for the energies &E(j\j2J) from
simple physical arguments. It is remarkable how far this will take us.
We start with the obvious statement that the attractive 5-function interac-
tion can only be large when there is large spatial overlap between the orbits of
the particles. As we have seen, for a given ;t and;2, the overlap for different J
values depends on the orientation of the orbits in space. We shall see in Section
4.2 the explicit relation between the 3 -;' symbols in Eqs. 4.7 and 4.10 and the
relative angular orientation of the semiclassical orbit planes. For now, we
proceed more qualitatively. From overlap considerations alone, one might
think that the interaction would be largest either for J = 0 or J = Jrmax, which are
simply the two cases for which the angular momentum vectors j^ and ;2 are most
nearly antiparallel or parallel, respectively, and therefore those in which the
nucleons orbit the nucleus most nearly in the same plane.
While this simple view has an element of truth to it, the requirements of
antisymmetrization refine it considerably. Antisymmetrization, or the Pauli
principle, has enormous and profound implications throughout the study of
nuclear structure. We have seen how it determines the valley of stability,
validates the fundamental concept of independent particle motion in a dense
sea of nucleons, and gives the magic numbers and attendant shell structure.
These (or at least the first and last) are straightforward and obvious effects.
Others are subtle, even unexpected, and are certainly seldom appreciated.
Our case of two nucleons interacting via a short-range residual interaction is
just such a case. We start by considering the configuration of two identical
particles (two protons or two neutrons) in orbits jrjr
In this case, the two-particle wave function must be totally antisymmetric,
that is, antisymmetric in space, spin, and isospin. A careful understanding of
this requirement leads to some beautiful, remarkable, and profound results.
Since we are dealing with two identical particles, the states involved must
have total T=\ and hence T=l: they are symmetric with respect to the isospin
coordinates. Therefore they must be antisymmetric in space and spin
coordinates.
The Shell Model: Two-particle Configurations 73
Thus far we have carried out the discussion in what is known as a// coupling
scheme, in which the total angular momentum jl of particle 1 is coupled to the
total angular momentum j2 of particle 2 to produce a final total angular
momentum /. One can also think of the problem in terms of the so-called LS
coupling scheme, in which the orbital angular momenta of the two particles lv
12 are first coupled to total L and the intrinsic spins (1/2 ft) ^ and s2 are coupled
to S = 1 or 0. Generally the jj coupling scheme is more useful in most nuclear
structure applications, but in the present case the L5 coupling scheme sepa-
rates out the angular and spin coordinates nicely and allows a simpler under-
standing of the effects of antisymmetrization.
To see how this works, let us take a particular example that gives a simple
result. Consider the case of two particles in a 1 d5/2 t,aJ} configuration as shown
in Fig. 4.2. The /values are / 1 = 2 and / 2 - 3, respectively, and in both orbits,
j = / +1/2. The allowed / values range from 1 to 6. If we picture the vector
coupling of/, and;2 to form various J values, it is clear that the orbital planes of
the two particles will overlap the most when J = 7mix and /mjn. We therefore
expect one or both of these to be the most affected (lowered) by an attractive
short-range interaction. To proceed further, we note that Jma is greater than
^ + /2 = 5. Therefore, the /mai state can only be formed by aligning the orbital
angular momenta to L = 5 and the intrinsic spin angular momenta to S = 1, and
then aligning L and S to / = 6.
Since 5=1, the two intrinsic spins point in the same direction and thus the
spin part of the wave function is symmetric. Therefore, antisymmetrization of
the total wave function requires that the spatial part be antisymmetric. Denot-
ing the angular part of the ith particle wave function by ^(r), the requirement
of antisymmetrization in the angular coordinates is equivalent to 0|10r1)0/20r2) =
-0>1(r2)^2(r1). However, the 5-function interaction is only effective when the
particles are in contact, when tl = rr At this point, this expression for the
antisymmetrization condition, however, requires that the wave function equal
its negative [0;1(r)^2(r) = -0^(00^(0], which of course can only happen if each
side vanishes. Therefore the wave function vanishes at the only point in space
where the 5-function interaction acts, so the residual interaction has no effect
whatsoever in this particular J = 6 state. Turning the argument around, a 5-
function interaction between identical nucleons can only affect states through
amplitudes in which S = 0, in which case the spin part of the wave function is
antisymmetric and the spatial part is symmetric and need not vanish at
I f j - rj =0. Interestingly, although the 5-function interaction has no explicit
spin dependence, its effects depend critically on the relative orientations of the
spins of the two nucleons. In other words, antisymmetrization introduces an
implicit spin dependence.
Clearly, to determine which states are affected by a 5-interaction we would
like to know which have 5 = 0. Unfortunately, except for the configuration
I sz,2/= 0), where 5 obviously must be zero, all two-particle configurations that
can have 5 = 0 amplitudes will also have 5=1 amplitudes. For example, it might
be thought that the / = 1 state of the I dsa L,a J} configuration would be pure
5 = 0 since this state is made by antialigning;, = 5/2 (/+1/2) and;2 = 7/2 (/ + 1/2).
Therefore the two intrinsic spins are also antialigned, giving 5 = 0. Though this
argument does identify the main component of the configuration, it is a bit
74 Shell Model and Residual Interactions
naive: one must be careful of mentally mixing the LS and jj schemes. In fact,
the/=! state can be made in three ways—by coupling/j and/ 2 to L = l.with
S = 0 as just stated, but also with the same L value and vector coupling S = 1 to
again give J - I and, finally, by coupling /5 and 12 to give L = 2 and then
antialigning a S = 1 vector to give 7=1.
In general, the relative amplitudes of S = 0 and 1 in a given / value for the
configuration, | (/jS^jj (/2s2);2;JM) are given by the 9-j symbol
consider a couple of examples. Starting with the familiar I d5/2f7/27) case, ymln
and 7mM are 1 and 6, the parity is negative and thus only the odd / states will be
The Shell Model: Two-particle Configurations 75
Fig. 4.2. Semi-classical illustration of the coupling of intrinsic spins and orbital angular momenta
for two configurations (d$/2 f?/2) and (dJ/2 g?/2)
affected, that is J* = 1% 3~, 5~. Therefore the/= Instate is lowered the most, the
J = 3~, 5" states successively less so, and the / = 2", 4~ and 6 states not at all.
For the d5/2g7/2/} configuration, /min = 1, /max = 6, as before but here n = + so
that / = 6 is lowered the most and the order is (from lowest to highest) 6\ 4', 2+
(1% 3+, 5+ degenerate at the unperturbed position). Finally, for |g7/2 hlu2/},
/mjn = 2, /max = 9, TT = -, and so the order (again, lowest to highest) is 9~, 7 , 5 , 3
(with 2', 4-, 6 ,8~ unperturbed). We note an interesting point, namely that, just
as with S values, although the 5-function has no explicit parity dependence, the
resulting energies of states \JJ2J) are, in fact, different for the two different
parities.
Although no states (other than | s21/27 = 0)) have pure 5 = 0, a third method,
which is formally incorrect, emphasizes they; coupling picture and does always
give the right answer. To illustrate this procedure, in I g7/2hn/27), all states J = 2
to 9 can be formed with S = 0 or 1. The naive argument would say that, since
y\ = /t - 1/2 and y'2 = 12 + 1/2, the 7min = 2 state formed by antialigning j^ and y'2
would correspond to S = I and therefore be unaffected by a 5-interaction. The
same argumentation would imply that the ,/max = 9 state is S = 0, and therefore
would be lowered the most.
76 Shell Model and Residual Interactions
This approach can be used for equivalent orbit cases to obtain a very
important result. Consider a configuration 1/7) such as |(f7/2)27). The/value
corresponds to / + 1/2, and therefore the antialigned /mjn state, 7 = 0, should
have 5 = 0 and be most affected. Again, although the argument is not really
correct, the result is. Since the overlaps of the particle orbits are reduced with
increasing J, the excitation energies increase monotonically with /. Note that
the same result applies for any case of identical particles in equivalent orbits:
the J* = 0+ state will always lie lowest. Note also that, in this respect, the 5-
function interaction resembles the well-known effects of the pairing force that
is designed (defined, actually) to lower states in which pairs of identical
particles are coupled to spin 0. As pointed out just after Eq. 4.10, and as seen
in the upper panels in Fig. 4.3, this effect of the 5-interaction in \fj) configu-
ration is greater for higher; (proportional to (2/ + 1)).
This lowering of the 0+ state is an extremely important result. Ultimately, it
is the underlying reason why all even-even nuclei have 0+ ground states and,
often, large spacings to the 2+: level. It directly explains this result only for
nuclei two nucleons away from closed shells. However, we shall see in Chapter
5 that our two-particle result can be generalized to f configurations, and by
extension, to wave functions that are linear combinations of several/" configu-
rations. Further generalizations to vibrational excitations and deformed
nuclei will be seen in later chapters.
There is an easy geometrical way of viewing this case of identical nucleons
in equivalent orbits that provides a physical rationale for the lowering of the 0+
states. Intuitively, it might seem that the overlap of the two particles in the /min
and /max states would be comparable: the two orbits are essentially coplanar in
both cases. Once again, however, the Pauli principle plays a key role in
distinguishing these situations. We show a schematic illustration in Fig. 4.4. In
the / = /max state, the near alignment of the two / values implies nearly coplanar
orbits in which the two particles orbit in the same direction. The Pauli
principle, however, forbids contact. In effect, this means that the two particles
must repel each other at short distances: therefore they track each other
around the nucleus on opposite sides of an orbit so that they always remain
apart. Thus, there can be no 5-function interaction between them. For Jn = 0+,
the orbits once again are coplanar but now the two particles orbit in opposite
senses. As they do so, their separation will vary but the average separation will
clearly be much less than in the /maj state, "contact" situations occur, and a
large 5-function interaction results. The actual values oiAfj^f) for a number
of different spin combinations are summarized in Table 4.1, and Fig. 4.3 shows
several examples (including those used most often in the preceding discussion)
of AEdjyf) values under the influence of a 5-function interaction. In studying
Table 4.1, recall that the interactions are attractive so that larger values of
A(jJ2J) correspond to lower-lying levels. Perusal of the table and figure shows
that the preceding rules are always satisfied.
It is possible to summarize these results succintly, as is done in Table 4.2.
(The trigonometric functions in column 6 of Table 4.2 will be explained in
Section 4.2.) To illustrate the construction of the table, let us consider the top
row. A little thought, or working out a few examples of the preceding rules,
MULTIPLE! SPLITTINGS; 8 INTERACTION ( I d e n t i c a l P a r t i c l e s )
Fig. 4.3. Energy shifts for a 5-function residual interaction for identical nucleons in several different orbit combinations.
78 Shell Model and Residual Interactions
IDENTICAL NUCLEONS
EQUIVALENT ORBITS
Fig. 4.4. Pictorial illustration of the motions of two identical nucleons in equivalent orbits for the
cases of maximum and minimum total spin J, showing the effects of the Pauli principle.
will convince the reader that, for positive parity, the state most affected will be
•'rain tf 0\ ~/ 2 ) ig even (as m Idszgjra^)). Note that even values of (jl -/2) are
equivalent to odd values of (jl + j2). Moreover, since only even/states are
affected for positive parity states, odd values of (j\ + ;2) are the same as odd
values of (jl +j2 + J). The consistent element in the table is that, if yt +J2+J is
odd (even) for the affected states, then the lowest-lying state will be
•Anm = U ~ h I (•'ma* = A + A)- We wil1 encounter the important role of the
quantity^ + /2 + J later in dealing with interactions in p-n multiplets.
To illustrate how these ideas relate to empirical data on nuclei with two
valence nucleons, we recall Fig. 2.5, which showed the yrast states for the
typical shell model nuclei 210Po, 210Pb, and 134Te. We repeat those data here in
Fig. 4.5, along with the predictions for a 5-interaction normalized to the
Ei\ energy for the relevant configuration [(h9/2)2 for 210Po, (g9/2)2 for 210Pb
(these two cases are actually identical as seen in Eq. 4.10), and (f7/2)2 for 134Te.]
The agreement is quite good. Note the strong lowering of the 0+ ground state.
From data such as this, known for many nuclei, it is possible to estimate the
absolute strength of the interaction. One obtains, as a rough guideline, that
AE (j\ j2 J) ~ (30 MeV/A) A(Jl J2 J), where the A in the denominator is the
Table 4.1. Relative J state energy values for various identical two-particle configurations I / j J) with an attractive 5-interaction.*
J
/, A WX «f), n 0 1 2 3 4 5 6 7 8 9 10 11
d5* dM
5/2 5/2 f» + 3.00 0.685 0.286
dM ^
f» 0 0.173 0 0.457 0 1.29
S?<2 %nei
7/2 7/2 *» f» + 4.00 0.95 0.467 0.233
&7/Z f» 0 0.127 0 0.312 0 0.599 0 1.52
»»,» h1M
11/2 11/2 '»« 'iw + 6.00 1.47 0.785 0.493 0.318 0.180
*,» W 0 0.084 0 0.196 0 0.329 0 0.509 0 0.818 0 1.89
ds«
3/2 5/2 ^
P3« «H
+ 0 0.343 0 1.14
d
3* fM 2.4 0 0.686 0
PM dsa
d» 87/2
5/2 7/2 f
« + 0 0.230 0 0.518 0 1.40
d5a ^
f» 3.41 0 1.14 0 0.518 0
<» &7/2
d5* SM
5/2 9/2 f
« h» + 2.85 0 1.04 0 0.490 0
d5, hM 0 0.258 0 0.558 0 1.47
f» 8s/2
d
s. in*
5/2 11/2 h
iifl + 0 0.279 0 0.587 0 1.52
d^sa \m 2.65 0 0.979 0 0.470 0
f« \m
"The table gives values olAyjfl from Eqs. 4.7 and 4.10. Dashes indicate J values not allowed for the given configuration. States that are unaffected by the interaction (i.e.,7 even for
TT = -, J odd for x = + configurations) are given a value of 0. Other numbers are proportional to the decrease in energy of the state /.
80 Shell Model and Residual Interactions
Table 4.2. Rules for the effects of a 5-function interaction on two-particle identical nucleon
configurations
Configuration Semi-
Category States classical
Parity Affected Lowest State dependence
A+/2 h+h + J one*
Positive odd odd even/ mill
tan 6/2
Negative even odd odd/ (antialigned)
Postive even even even/ '-. cot 6/2
Negative odd even odd/ (aligned)
*6 is defined by
mass number. Recalling that AO"2 7 = 0) = (2; +1)12, this gives a typical lowering
of the ground state of several MeV in light nuclei and of 1 to 1.5 MeV in heavy
nuclei where the interaction strength is only ~ 200 ke V. Both of these are well-
known features of the data (e.g., Figs. 2.6-2.10), and again show that a 5-
function naturally produces the famous "energy gap" in even-even nuclei
usually associated with the pairing force (see the following).
In all these examples, there is another important feature we have not yet
commented on: the lowest level for a given multiple! is substantially lowered,
but the differences in interaction strength for the others monotonically de-
Fig. 4.5. Comparison of experimental and calculated low-lying even spin (yrast) states in three
nuclei with two valence nucleons. The orbits used for the two identical nucleons are indicated in
each case.
The Shell Model: Two-particle Configurations 81
crease. That is, there is a relative compression of levels near the unperturbed
position. This is not just an accident of the 3 - / symbols, but has a simple
physical origin that we shall discuss shortly. As we noted in Fig. 2.5 and, here
in Fig. 4.5 for the case of jl = 72> it is also a well-known empirical effect
characteristic of the low levels of many "shell model" nuclei.
We have been discussing the effects of a 5-interaction between identical
nucleons. Such states have iz (1) = tz (2), hence Tf = ±1, and T = 1. The
proton-neutron system also exists in a T= 1 state. By charge independence of
the non-Coulomb part of the nuclear force, the p-n T= 1 system must then also
satisfy Eqs. 4.5-4.7. Indeed, the familiar statement of charge independence
that p-p, n-n, and n-p forces are equal applies specifically (and only) to the
T = 1 mode for the p-n system.
As we have seen, however, the p-n system can also exist in a T= 0 state for
which there is no need for equality to the p-p or n-n forces. Empirically, in
fact, the T- 0 interaction seems to be significantly stronger than the T= 1 (see
the following). This T= 0 coupling is extremely important in nuclear structure,
as it is now thought to be responsible for single-particle configuration mixing
and the onset of collectivity, phase transitions, and deformations. We shall
return to these points in later chapters. For now, we are interested in simple
two-particle p-n configurations in shell model (noncollective, nondeformed)
nuclei under the action of a 5-interaction.
In order to address this issue, we must deal with a specific complication thai
arises in the p-n system. Suppose we imagine such a system occupying levels
a and b as shown on the left in Fig. 4.6a. Then, if we treat the proton and
neutron as two states of the same particle (the nucleon), the orbits or the
charges call be exchanged indistinguishably. Thus, the wavefunction for the
Fig. 4.6. "Direct" and "exchange" configurations for protons and neutrons treated as indistin-
guishable particles (a) filling the same shell, and (b) where the neutron shell corresponding to the
valence protons is already filled.
82 Shell Model and Residual Interactions
The four terms on the right correspond to the four cases of particles 1 and 2,
each occupying levels a or b, and each identified as either a proton or a
neutron. The term in a in Eq. 4.11 is a direct term that should be relatively
large since the overlap of the wave functions on the two sides is unity. The last
term keeps the particles in the same orbits (1 in a, 2 in b) as on the left side but
exchanges their type (p -» n, n -» p). This can also be large if the interaction
contains terms that can change a proton into a neutron and vice versa. The
second and third terms involve overlaps in which the particles change orbits:
therefore they are normally small. In principle, however, all must be taken into
account. Clearly, this is both complicated and tedious. The isospin formalism
for doing so is well known and is discussed in standard texts, so we will not
consider this situation. Here we are largely concerned with medium and heavy
nuclei, which greatly simplifies the problem since protons and neutrons are
usually filling different major shells. This situation is illustrated in Fig. 4.6b,
where the neutron shell corresponding to the proton shell is already filled:
hence, exchange matrix elements such as those in the 7 and <5 terms in Eq. 4.1 1
(the rightmost two terms in Fig. 4.6a) are impossible ("blocked"), since the
neutron orbits in that shell are filled. The only remaining exchange term is one
of those expected to be small. We can consider only the direct term to good
approximation in such a case. This is equivalent to treating the protons and
neutrons as distinguishable particles. Note that, while we can choose to do so
in this case, because of the blocking of exchange terms, it is not mandatory: we
could (and will momentarily) use the isospin formalism as well. The
proton-neutron and isospin formalisms are equivalent ways of obtaining the
same result. Which one is used in a given case is a choice based on practical
simplicity. When exchange effects are known, a priori, to be inconsequential,
the proton-neutron formalism may be simpler.
In the present case, treating the protons and neutrons as distinguishable
means that the Pauli principle places no restrictions on their coupling. Their
wave functions need not be antisymmetrized. We can write the two-particle
p-n wave function as y/^ = 0 0n. To see the relation to the isospin approach,
this can be rewritten as \y n = l/2[(0 <t>n + 0n0 ) + (0 0n - 0n0 )] • The first term is
symmetric with respect to interchange of protons and neutrons, the second
antisymmetric. These two terms therefore correspond to T = 1 and T = 0,
respectively, and the energy shifts in the presence of a residual interaction can
be written as the average of their values in the two isospin channels, that is, as
1/2[AE(T = 1) + A£(T = 0)]. In the like-nucleon case, antisymmetrization
requires that the S = 1 (symmetric) spin coupling be accompanied by an
antisymmetric spatial part of the wave function (which vanishes when the two
particles are in contact): such states do not feel the 5-interaction. Here, there
is no such limitation. Both 5 = 0 and 5 = 1 states can be accompanied by
The Shell Model: Two-particle Configurations 83
symmetric spatial wave functions and will be shifted by the interaction. Gen-
erally, the interaction need not be of the same strength in the two cases, and so
we specify two interaction strengths Vs_0 and Vs=l
where it is implicit that the Vs=t> (VJ=1) term acts only on the S - 0 (5 = 1)
components.
We can get the same result in terms of isospin. We saw before that the
unsymmetrized two-particle wave function can be written in terms of an
average over T = 1 and T = 0 parts. In this case, the wave function for each
isospin term must be separately antisymmetric. Specifying to a 5-function
interaction, the only states affected must have symmetric spatial wave func-
tions. Therefore, the isospin-spin part must be antisymmetric: hence, the
T = 0 part goes with 5=1 and the T= 1 with 5 = 0. Again, we get S = 0 and 1
terms as in Eq. 4.12. (Recall that the total nuclear wave function must have
good isospin, which is obtained by coupling the isospin of the two-particle
system to that of the (T * 0) core.)
Thus, the analogues of Eqs. 4.5-4.7 for a p-n system under the action of a 5-
interaction given by Eq. 4.12 become:
where
and
The 3 -;' symbol is identical to that appearing in the like-nucleon case, but now
there are two terms with different /,/ ,jn dependencies. Note that, in the first,
or Vs=0, term in Eq.4.15, only half of the levels are affected, namely those with
even / for positive parity or with odd J for negative parity. This is the same con-
dition we saw for like nucleons, as it must be since this is the 5 = 0 (T- 1) term.
The other/values are then affected by the first of the Vs^ terms. For; *jn, the
second Vs=l term itself affects all /values, and small / values the most.
As in the like-nucleon case, this equation simplifies for equivalent orbits
84 Shell Model and Residual Interactions
Here we can easily see the explicit relation to the isospin formalism. The
T - 1 p-n interaction must be identical to the p-p and n-n interactions. We
saw earlier that, for equivalent orbits, only even / states are allowed for the
T = 1 p-p and n-n systems, and hence, for a p-n system, which has all / values
from 0 to 2;', the even/values must have T= 1 and the remaining levels, namely
those with odd /, must be T = 0. Thus, in Eq. 4.16, the first term corresponds
to the interaction in the T= 1 channel and the second term to the T=0 channel.
We now need to consider the relative strengths Vfc0 and V5=1. From the fact
that the deuteron has an S = 1 ground state, it is clear that V5=1 is stronger than
Vs=0. However, there is additional evidence for this from such simple data as
neutron separation energies that is directly applicable to nuclei with all A
values. As we have just seen (Eq. 4.16) the T = 1 and 0 interactions can be
associated with the S - 0 and 1 terms, respectively. We saw in Chapter 1 from
the separation energy data that the nonpairing, like-nucleon (T- 1) residual
interaction is, on average, repulsive, where by the phrase "on average" we
mean averaged over all final / states and by "nonpairing" we mean excluding
the 0+ state (if any). So, by charge independence, the p-n T = 1 interaction
must on average also be repulsive. Yet, we also noted in Chapter 1 that both
S(p) and S(n) increase with increasing numbers of particles of the opposite
type. The interaction between protons and neutrons has both T = 0 and 7" = 1
components. So, on balance, the total (T- 0 + T= 1) p-n interaction must be
attractive. This can only occur if the T = 0 component is both attractive and
stronger than the T= 1, that is, if the \Vs=l\ is greater than Vs=0 .
Of course, the strength of the two isospin components of the interaction can
also be obtained by fitting actual p-n multiplets (groups of states with pure
proton and neutron configurations; and jn and /values ranging from \j -jn\
to jf + ;J. Schiffer and True and Molinari and co-workers have carried out
extensive surveys of this type near all closed shells from 160 to 208Pb. We will
discuss their results in Section 4.2 in terms of a simple geometrical analysis.
Here it is useful to convey a feeling as to how the data on individual isospins
can be deduced. The nuclei near 208Pb offer a nice example. Consider, for
example, the states of ^Po^ in a Ih9/2li13/z7) two-proton T = 1 multiple!.
(These can be found from the 209Bi (3He, d) 210Po reaction since 209Bi has a single
proton in the lh9/2 orbit.) The energy shifts in this multiplet can be used to
extract the (lh9/2 iiiy2) T= 1 interaction. The same multiplet exists in20883BiU5
as a particle-hole p-n multiplet. The energy shifts AEQJ-JJ) can be converted
(see end of chapter) to an equivalent set of particle-particle shifts and the total
p-n interaction obtained for each/state. The difference of the T= 1 and total
interactions then yields the net T = 0 strengths. Extraction of T = I and T = 0
strengths is even simpler in the case of equivalent orbits (/ = y'J, of course,
where the even and odd J states directly give the T = 1 and 0 interactions,
respectively. This approach is useful in light nuclei where the protons and
neutrons are filling identical orbits (e.g., the f7/2 orbit in 42Sc).
To illustrate the application of these ideas, we consider the classic example
of 1738C121. Since the N = 8 to 20 neutron shell is filled, this is an appropriate
case to ignore exchange terms. In the lowest-lying states, the configuration is
(d.,^ f7/2/j) giving states J = 2 ,3", 4-, and 5 . Since Vs^ > Vs=0, the second group
The Shell Model: Two-particle Configurations 85
Fig. 4.7. Comparison of low-lying empirical and calculated energies for 38C1. The two panels on
the right correspond to calculations with a two-body 5-function residual interaction, assuming two
different orbits for the proton. Clearly, the (d3/. f 7 _) configuration is favored. The calculation on
the left uses the empirical levels of the (d.^"1 f?/2) particle-hole configuration in 40K in conjunction
with Eq. 4.34 to predict the particle-particle levels of 38C1. (See deShalit, 1974.)
of terms in Eq. 4.15 will generally dominate and the overall ordering of levels
in the p-n system will tend to be contrary to that in the like nucleon case.
Moreover, whereas only half the states are affected for like nucleons (/ odd for
n = -\J even for n = +), all states will be shifted in the p-n case. We therefore
may expect the lowest level to be the even J state with highest overlap, the
/ = 2~ level. The 38C1 experimental spectrum and that calculated with
Vs=1 = 2F5=0 are shown on the right in Fig. 4.7. (The part on the left describes
an alternate approach to calculating 38C1, to be discussed near the end of this
chapter.) The 2-level does in fact occur lowest, and the agreement is reason-
able. The figure also shows that the calculated levels for an alternate configu-
ration with the same / values, (p3/2p_ f7/2J, have a rather different pattern since,
here, the orbital phase factors in Eq. 4.15 are different (/ + ln is now even) and
the J = 5, 3 set is lowered relative to the J = 2,4 pair in disagreement with the
data. This indicates how one can even sometimes suggest; configurations and
/"values by examining energy sequences and spacings in p-n multiplets.
Fig. 4.8. Definition and schematic illustration of some of the ideas used in the geometrical analysis
of short-range residual interactions.
or
From here on, for simplicity, we take the case of identical particles in
equivalent orbits (j\ = j2 -./) and assume that/, 7 » 1 so that terms like./(./ + 1)
can be approximated by J1. Then,
Note that 0=0° corresponds to high J and 6 -180° corresponds to low J. Thus,
for ;; = jv G = 180° corresponds to J = 0 and 0^ -» 0° to J = 7mai = 2; - 1.
Before proceeding, we first make use of some simple trigonometric equa-
tions. From sin2 0=1 - cos2 6, we obtain
The Shell Model: Two-particle Configurations 87
Hence,
Using the relations for sin 912 and tan 012, we have
or, finally
This extremely simple result expresses the shifts in different / states for a 5-
interaction between two identical particles in equivalent orbits. It was derived
for large j, J, but is remarkably accurate even for low spins (e.g., as low as
j = 3/2 and 7 = 1 but specifically not for J = 0). The function tan 9/2 is plotted
against 9 in Fig. 4.9. Since tan 9/2 ~ 9/2 for small 9 and goes to infinity for
88 Shell Model and Residual Interactions
EQUIVALENT ORBITS
Fig. 4.9. Dependence of the <5-f unction residual interaction strength (lower values correspond lo
mDreattractiveresidualinteracu'ons)fortwoparticlcsinequivalentorbits. (Left)The7"=l (./even)states. (Right)
The 7 = 0 (/odd) states. The analytic expressions are indicated above their respective plots.
8 -> 180°, Eq. 4.18 simply states that the energy shifts become large (and
negative since the force is attractive) for 9 ~ 180° (for small J where the two
angular momenta are antialigned), while the smallest effect occurs when
J = 7max = y + y, since 0= 0°. Moreover, the curve tan 0/2 becomes asymptoti-
cally flat for large J, giving a geometrical interpretation to the compression in
spacings discussed above and illustrated for 210Po, 210Pb, and 134Te in Fig. 2.5.
Note that this formula automatically reflects the Pauli principle arguments
discussed earlier, in which the 5-interaction affects only 5-0 states. The Pauli
principle appears here through the 3-y symbol and, in particular, through the
spin angular momenta of ±1/2 appearing in it.
As we have noted, identical nucleons in equivalent orbits have T= 1, so Eq.
4.18 applies to them. For proton-neutron configurations we have to consider
both 7'= 1 and T- 0 parts. We still restrict ourselves, though, to equivalent
orbits. A T= 1 p-n system, which cannot be distinguished from the T- I p-p
and n-n systems, consists of even / states and is described by Eq. 4.18 as well.
However, in a T = 0 p-n system (odd ./ states ) AE n (fj) has a different./
dependence, reflecting the different./ behavior of the second (V^) term in
Eqs. 4.15 and 4.16 compared to the first term.
For the odd 7, T = 0 case, a similar analysis gives (again for;' = y'J
The Shell Model: Two-particle Configurations 89
The behavior otAE(fJ) for T= 0 is also indicated in Fig. 4.9. We recall that
this only applies to equivalent orbits / = jn. While the T = 1 interaction is
largest (lowest-lying on the plot) for 7min (9 ~ 180°) and is smallest for
/ = /mM (9-> 0°), the T = 0 expression is large for both /min and 7mM. For T = 0
and 9~1800,&E(fJmit)~l/cos9--*°<>-,foT9=Q0,AE(j2JmJ~col9-^<>°. Both the
r = 0 and T= 1 expressions are small for 0= 90°.
All these features can be easily understood physically. The interaction
should be small for 9 = 90° for both T = 0 and T = 1, since the particles are
orbiting in nearly perpendicular planes and are seldom close enough to inter-
act. For T - 1 (which, by charge independence, means we can re-use the
identical-particle arguments), the interaction is strong when the two nucleons
orbit in opposite directions (J = 0,9= 180°). However, it vanishes when they
orbit in the same direction (Jmsa, 9 = 0°) since, then, the two particles have
identical quantum numbers and the spatial wave function is required to be
antisymmetric: it must vanish if the nucleons "touch." The Pauli principle
effectively introduces a short-range repulsion. The only way the particles can
orbit in the same direction and yet not touch is if they circulate out of phase at
opposite ends of an orbit diameter. This gives an interaction that is small for
small 9, but large for large angles in agreement with Fig. 4.9. The basic idea is
the same as for the identical particle T= 1 case (Fig. 4.4). For the T= 0 case we
treat the particles as distinct and, for both the small and large / extremes, the
orbits are nearly coplanar. Since we need not worry about antisymmetry, there
is no restriction on phasing, and "contact" is abundant, leading to a strong
interaction for both 0- 0° and 0= 180°.
Empirically, these effects are well documented as shown by the examples in
Fig. 4.10 taken from the aforementioned empirical analyses of p-n multiplets
throughout the periodic table by Schiffer and True. Note the interesting point
that for even J, T = 1, the empirical interaction is actually slightly positive
(repulsive) for small 9 (high /). A 5-function interaction cannot give this: at
best, it vanishes near J - Jmm. Such an analysis clearly shows the need for a
separate repulsive component in the residual interaction. Several studies have
successfully carried out multipole analyses of these effective residual interac-
tion, incorporating dipole, quadrupole, etc. components. Evidence for a
sizable quadrupole component has been found. This multipole varies as
P2(cos0) where, again, 0is the angle between the two orbits. As is well known,
this function crosses zero at 9 ~ 55° so that even for an overall attractive quad-
rupole term, the interaction is actually repulsive for angles between 55° and
125°. This is just the region where Fig. 4.10 shows positive (repulsive) empirical
T = 1 interactions. This repulsive aspect should not be surprising. We have
already encountered it. We noted in our discussion of separation energies in
Chapter 1 that the like nucleon (T = 1) nonpairing residual interaction was, on
90 Shell Model and Residual Interactions
balance, repulsive (S(n) decreases with increasing N). From this empirical
fact, we also deduced that the T-Q interaction is on balance stronger (more
negative) than the T=l. This is also evident in Fig. 4.10.
Finally, note that for J* = 0+, the interaction deviates from the geometric
expression. The 0+ behavior, however, is physically reasonable. As with the
like-nucleon case, the interactions are ordered by;': they are largest for large/.
The larger the) value, the more magnetic substates there are, and the smaller
the permissible angular range of an orbit for a given m. Thus the orbit planes
are more tightly defined and the overlaps of particles in ±m substates are
greater.
Fig. 4.10. F.mpirical proton-neutron multiplets for two particle equivalent orbit configurations for
comparison with the behavior shown in Fig. 4.9. The curves arc drawn through the data (Schiffer,
1971).
The Shell Model: Two-particle Configurations 91
Thus far, for simplicity, we have carried out the geometrical analysis for the
simple case of/, =/2. For/, */2 and the identical particles, we saw in Section 4.1
that either the 7^(0^ 180°) or the/ max (0^ 0°) state can be lowest depending
on the particular/values and their/ = / ± 1/2 character. Table 4.2 summarized
the different cases leading to these two situations. These two categories of
two-particle configurations should be and are reflected in the geometrical
analysis. One obtains two curves now, of which one is identical to Eq. 4.18
(AE ~ tan#/2) giving the lowest energy for the antialignment of the two values
(7^, 9 close to 180°), and the other curve goes as cot0/2 so that the lowest
energy occurs for parallel alignment (7max) and 9 close to 0°. The correspon-
dence of these two trigonometrical functions and different sets of/ : ,/ 2 values is
made explicit in the sixth column of Table 4.2. Finally, note that the equiva-
lent-orbit situation is actually a special case of this. Here, n = +,/, + /2 = 2j is
always odd and so the land/2 dependence applies and the ./min (in this case ()*)
state is lowest.
Fig. 4.11. Comparison of empirical and calculated multiple! splittings for two-particle configura-
tions of noncquivalent orbits (Sehiffcr, 1971).
92 Shell Model and Residual Interactions
Fig. 4.12. A geometrical analysis of the lh<J/2g9/2/> p-n multiple! in 210Bi. The empirical levels are
shown on the left along with thesemiclassical angle between the orbits of the two nucleons. The
right side shows that the levels split into two families, according to7 even or J odd. The solid lines
are drawn to connect the points.
For the p-n case, a similar analysis again leads to two distinct curves as in
the T= I and T- 0 cases for equivalent orbits. However, the classification is
slightly different. Recall that, for equivalent orbits, the T= I states are even,/
only. Thus /, +j2 + J-2j + J is odd. The T = 0 case, with J odd, has /, + j2 +./ =
2; + J even. It is this distinction that persists when;^ *jz for a p-n multiple!.
Again, we obtain two curves, but distinguished according to the odd or even
character o f y , + y'2 + J and describable by geometrical functions of 9 very
similar to Eqs. 4.18 and 4.19. This is beautifully illustrated by the data for
several multiplets collected in Fig. 4.11, and exemplified in depth for 210Bi in
Fig. 4.12. In all these cases the empirical energy distributions within the p-n
multiplets follow the expected energy patterns quite well.
One last point worth mentioning is that extensive surveys of empirical p-n
interaction multiplets show that the strength of the interaction, especially in
T = 0 states, smoothly decreases with increasing mass. This is quite plausible
since the average radius of shell model orbits increases with higher oscillator
numbers, while the interaction range is constant so that the average interaction
strength decreases. In heavy nuclei, typical V n interaction matrix elements
are ~ 200 to 300 keV but of course this depends on the orbits involved.
What is perhaps most important to emphasize in concluding this part of the
discussion is that, without ever having dealt with the radial parts of the wave
functions, or indeed, calculating anything, it has been possible to predict the
qualitative energy ordering of the different./ states in two-particle configura-
tions. Moreover, exact quantitative results for the relative spacings involve
only the evaluation of a single 3-/symbol. (Of course, the absolute spacings
The Shell Model: Two-particle Configurations 93
depend on the radial integrations and the strength of the interaction.) This is
but one example of how far one can go in a shell model treatment of multipar-
ticle configurations by invoking only very general arguments.
where G gives the overall strength in the interaction. Note that this interaction
is attractive and, by definition, only effective for 0+ states of identical nucleons
in equivalent orbits. It is not, however, limited to diagonal matrix elements
O^O* I Vpair I/Y^X but rather allows nondiagonal scatterings, </'t20+1V alr l/320+),
in which the pair of particles switches to another orbit as a pair. This feature is
critical to the build-up of pairing correlations and the so-called pairing gap in
even-even nuclei, and will be treated in more detail in a later chapter. For
diagonal matrix elements I/, = j2 - j3 = j4) the pairing interaction strongly
lowers the 0+ state without affecting the others.
Both the 5-function force and the pairing force are intended to represent
the short-range component of the nuclear interaction. However, the residual
interaction also contains a long-range component, that, as we shall see, is
Fig. 4.13. Comparison of levels of a (/ = 7/2)2 configuration for a 5-function and a pairing
interaction.
94 Shell Model and Residual Interactions
where 0is the angle between the radius vectors to each particle (see Fig. 4.8)
and K-is the strength of the quadrupole part. Clearly, while Kpair is short-range,
the quadrupole component simulates at least part of the long-range aspect of
the residual interaction (see the following section).
The popularity of the pairing interaction, or of any other that reproduces
the low energy of/ = 0 coupled pairs of nucleons, clearly lies in the fact that all
even-even nuclei have 0+ ground states. The quadrupole interaction is moti-
vated, empirically, by the fact that nearly all nuclei more than a few mass
numbers away from closed shells display properties that can be described in
terms of quadrupole distortions of the spherical shape. Figure 4.13 shows the
differences between the pairing and 5-function interactions. For the case of
identical nucleons in equivalent orbits, both produce a low-lying 0+ state and a
grouped cluster of states with higher angular momentum. With the pairing
interaction, this latter group is unaffected, remaining degenerate. The 8-
function interaction seems to be a better approximation, as is clear in Fig. 4.5,
as well as Fig. 2.6 for the Sn isotopes. For many-particle configurations, the
quadrupole force dominates both the pairing or «5-function interactions and
the differences in their properties are washed out.
itly, but the results are still valid for identical particles and for p-n systems
involving different major shells. If we write the wave function as a product of
radial and angular parts, we can then achieve a similar separation of the energy
shift into radial and angular parts and, moreover, separate the angular parts of
particles 1 and 2. We obtain, as in Eq. 4.5
where
and
independent of the interaction. This case of d^, L,a particles should actually
describe the lowest states of 1738CL,r The experimental spacings are in the
ratios 1:1.1:2.0. They have the same sequencing of levels as in the calculation
(2-, 5", 3~, 4~), but the magnitudes of the spacings disagree. Without ever having
specified an interaction, or evaluated any radial wave functions, we can
unambiguously conclude either that an accurate description of these 38C1 states
requires more complex configurations, or that the interaction depends on
something other than (r}- r2) (e.g., spin). Note that this is the same nucleus
that we treated earlier with a 5-function interaction (Fig. 4.7). The earlier
results are different than these precisely because we used different strengths
(i.e., a spin dependence) for the S = 0 and S = 1 terms.
Equation 4.24 applies to any interaction that can be written in terms of the
separation of the two particles, that is, as V( \ ra - r 2 1). However, one some-
times encounters spin-dependent interactions. Then we obtain a result similar
to Eq. 4.24, with the same limitation on kmax, but including odd k. This point is
very important, since the 5-interaction can be shown to be equivalent to an
interaction V( \ ^ - r 2 1) times a spin-dependent operator. Thus, a ^-function
interaction is equivalent to an odd tensor interaction, even though its mul-
96 Shell Model and Residual Interactions
tipole expansion contains only even k. The reason is beyond the scope of our
treatment but involves the fact that the interaction takes place only at "con-
tact," and therefore a 5-function interaction can be multiplied by various
"exchange" interactions that have odd tensor character. This class of interac-
tions is crucial in nuclear structure since they have the special property of
conserving seniority (see Chapter 5).
For a 5-interaction, v^r ) is given by
Note that, despite its short-range, the 5-interaction has a monopole compo-
nent (k = 0) corresponding to a part of the interaction that is independent of
angle [P0(cos0) = constant]. This is a long-range piece par excellence and it
says that a "contact" interaction has a component that pays no heed to the
angular separation of the particles! Of course, since the monopole part is
constant over all (angular) space, it is the same for all relative orientations of
j\ and j2, that is, all / values. Hence, it does not contribute to splittings of a
multiple! but does give an overall shift to the multiplet. In practical calcula-
tions, the situation in regard to allowed multipoles can be even more bizarre.
Consider a 5-interaction between two d^ particles. Not only does the k = 0
multipole contribute, but the triangle conditions on k limit the multipoles to
only the values k < 3. No really short-range multipoles appear.
This seemingly paradoxical situation of an infinitly short-range interaction
being simulated only by relatively long-range multipoles is actually easy to
understand. This understanding reveals much about the relationship between
forces and the orbits and wave functions of the interacting particles. More
accurately, it clarifies the way particles from different orbits can "probe"
different interactions. It relates in a general philosophical way to how one
determines structure in any physical system. The general rule is that in order
to sample the structure of a given scale, the probe must be comparable to or
smaller than that scale. (It is difficult to distinguish between a potato and a
carrot by bombarding either with a truck.)
Imagine a proton and a neutron in an s1/2 orbit. They have only one m state
and are spherically symmetric wave functions: the orbit is uniformly spread
out, at a given radius, over time, in a spherical shell. The two particles are
therefore always in "contact" and are simply unable to sense any details of the
residual attraction. To them, a 5-force is identical to a constant force over all
space. The higher the; value of the interacting particles, the more sensitive
they are to the details of the force simply because higher; values have more
magnetic substates. This set of magnetic substates spans the same angular
range around the body of the nucleus and hence, each magnetic substate is
restricted to a narrower angular range. Thus, two particles in; - 13/2 orbits
coupled to / = 12 can sense the fine details of an interaction: each substate M
samples a different angular range of the force. In contrast, even though the
residual interaction may be a 5-force, two low; (e.g., s1/2, p1/2, PM,...) orbits
cannot "know" this. They are the wrong probe.
Fig. 4.14. Multipole decomposition of residual interactions: (a) Legendre polynomials Pk(cos &) plotted against 8 and J value for a (9/2,11/2) configuration, (b)
Contribution of different multipoles to AE (gn \mf) with a § function residual interaction. The short-dashed, and solid curves correspond to the multipoles k = 0,2
and k = 0,2,4,6, and 8, respectively. For the complete (k = 0-8) case, the even and odd J states are separately connected by broken lines, (c) Effect of the addition
of each successive multipole on the energies of an (hj^)2 configuration. (Parts (b) and (c) are based on Heyde, 1989.) (d) Effect of forces of different ranges on
the relative energy shifts as a function of 6, that is, of /, for two different two-particle configurations. (Schiffer and True, 1976.)
98 Shell Model and Residual Interactions
Coupled with the practical historic fact that most shell model calculations
have dealt with light nuclei and, therefore, low / states, we now see why
different interactions are often used to account for the same data, and why it
often is difficult to determine the details of the interaction.
The other extreme of k values (multipole order) can also be discussed.
Neglecting the limitation given by the jjr J212 quantum numbers (or, equiv-
alently, assuming a large j shell), we can relate the largest relevant k values to
the range of the force. Consider the classical picture shown in Fig. 4.8, where
0is the angular separation of the two particles. The distance between the two
particles is given by r2 = rf + r* + 2 r/2 cos 6. If we approximate the range by
a single number rr, the integration over 0 will be limited to 9 < rjr. The
Legendre polynomial Pk (cos#) oscillates more rapidly as k increases (see Fig.
4.14a). If Pk oscillates many times within the allowed integration range, the
integral will be small because of cancellation effects. In practice, therefore, for
a given force range, k is limited to value satisfying rik « rlr. Clearly, then, as
the range decreases, more and more A: values are required. The limiting case is
of course, the 5-interaction, although other factors (/,/ values) relating to the
ability of a given wave function to "sample" an interaction will come into play
to limit the allowed range of k values.
The idea of the multipole expansion of a residual interaction can be used to
obtain some earlier results in a physically transparent way that offers new
insights. This idea can also be used to derive a famous result known as the
parabolic rule for energies of states in p-n multiplets. We have already seen
how one can understand the /-dependence of the residual interaction in terms
of the angle between the two orbits involved (see Eqs. 4.18, 4.19 and Figs.
4.9-4.12). We can apply similar arguments to specific multipoles.
Each multipole, k, has an angular dependence Pk(cos6). The lowest few are
P0 = constant, Pz = l/2(3cos20-1), PA = l/8(35cos"0- 30cos20 + 3). The Pk were
shown in Fig. 4.14a. As Eq. 4.17 indicates, for a given j\,j2 there is a specific
relation between the total angular momentum 7 and the angle 6: thus, one can
plot Pt(cos0) equally well against J or /(/+!) (as is sometimes done). This
alternate scale is included in Fig. 4.17a for the example of/^ 9/2, ;2= 11/2 (e.g.,
a
(SwAin) configuration).
Obviously, since each multipole is proportional to P^cosfl), the interaction
in the kth multipole is strongest when Pk is largest. Thus, we can tell directly
which spin states will be most affected for each multipole.
The monopole (k = 0) component is constant, affects all states equally, and
simply gives an overall shift to the entire multiple!. This is an important effect,
altering the relative excitation energies of different multiplets, but it contrib-
utes nothing to the splitting.
It is the principal origin of the mass dependence of single-particle energies
that we discussed at the end of Chapter 3. In particular, the monopole p-n
interaction leads to the changes noted there (Fig. 3.4) in neutron single-
particle energies as a function of proton number and vice versa. In this
chapter, however, we are focusing on energy splittings due to residual interac-
tions and so we turn now to other multipoles.
The Shell Model: Two-particle Configurations 99
*I am indebted to K. Heyde for much helpful advice in this section and for Figs. 4.lib and 4.14c.
100 Shell Model and Residual Interactions
multipoles, k = 0,2,4,6,8, gives the solid curve. If we connect just the odd J
states along this curve, we get a dependence just like the jl+J2+J odd curve of
Fig. 4.11. If we connect the even/values we get the j\+j2+J even curve of Fig.
4.11. Thus we see that the separation of the two curves comes from the higher
multipoles and reflects the Pauli effects since these multipoles have shorter
ranges.
Figure 4.14c shows this in a particularly illuminating way for the identical
particle I (h]1/2)27) configuration. Here, the energies obtained with the addi-
tion of each successive multipole in a ^-interaction are displayed in level
scheme form. For k - 0,2 only, a parabolic behavior is observed: the 0+, 2+, 4+
levels are lowest, the intermediate spins 6+, 8+ highest and the largest spin, 10+,
lower again. As shorter and shorter range multipoles are added, the levels
shift toward the characteristic 5-function sequence shown on the right. We
have discussed how such a sequence, particularly the weak effects of a 5-
interaction for the Jmsa states (since they have S = 1), is a specific Pauli principle
effect. Now we see exactly how this comes in through the shortest-range
multipoles, which are in fact repulsive for these spin states, reflecting the Pauli
prohibition against "contact."
Figures 4.14a and 4.14c also dramatically show why the 0+ state (or, gener-
ally, the lowest state of a given multiple!) is lowered so much. This state
corresponds to orbit planes closest to 9 = 180° (or 0°, depending on the; values
if nonequivalent orbits are involved). Here, all the multipoles contribute
coherently because P^cos 180° or 0°) is always unity. Only for other spins do
cancellation effects enter.
Finally, we can use Fig. 4.14c to better understand the relation between 5-
and pairing forces. Imagine continuing this figure further to the right. The
0+ - J separation would grow and the separation among the J* * Q+ states
would diminish relative to their separation from the 0+ level. The limit is the
pairing picture shown in Fig. 4.13. Although we do not write the pairing
interaction in terms of multipoles, we see that, in effect, it corresponds to the
dominance of very high ones. This, incidentally, is the origin and basis for the
phrase occasionally encountered that the pairing force is of even "shorter
range" than the 5-interaction, a statement that sounds paradoxical.
We have now discussed, in several ways, the relation between multipoles
and the (angular) ranges of the forces they describe and their effects on differ-
ent J states of two-particle multiplets: briefly, attractive short-range (higher
multipole order) forces tend to lower especially the Jma and Jain states (subject
always to Pauli princliple constraints). We can see this relation even more
explicitly by using forces that are finite (radial) range. To this end, Fig. 4.14d
shows the energy shifts calculated for two configurations for a Yukawa type
force with range parameter r = 0.1,1.0 and 5.0 fm. In general, the shorter the
range, the stronger the effects on the extreme J (or 0) states where the orbits
are most nearly coplanar and the particles, on average, closer. The different
patterns for r = 0.1fm for the 7=1 interaction just reflect the cot 9/2 or
tan 8/2 dependences discussed for the two different configurations (see Table
4.2). This difference, a Pauli effect, is washed out for longer-range forces, as is
The Shell Model: Two-particle Configurations 101
most of the J (or ff) dependence, since the force becomes rather insensitive to
the separation of the particles within the nuclear volume.
With this discussion of multipole contributions and force ranges, the reader
should now be in a position to estimate, at least qualitatively, the effect of any
particular interaction in a given configuation with little or no explicit calcula-
tion.
An interesting special case of the multipole decomposition is that of a
(PM^M) P~n configuration. For this configuration, Eq. 4.24 only allows k = 0,
2 values, and the k = 0 multipole is irrelevant for splittings, so the 5-interaction
in this case should give exactly the parabolic dependence on /(/ +1) character-
istic of P2(cos9). Yet, from our previous discussion, we know that the J = 0,2
(/i + J2 + J odd) states would occur on the upper curve in Fig. 4.11 and the
J = 1, 3 states on the lower curve. Although this hardly seems to yield a
parabola, careful inspection of the curves in Fig. 4.11 for the angles appropri-
ate to the | dM pM /) configuration shows that these four points have an exact
parabolic form against /(/ +1).
Actually, this result suggests why a quadrupole interaction is more impor-
tant than it might at first seem: because of the angular momentum constraints
on k , many higher multipoles that might normally contribute are elimi-
nated. Moreover, the 5-function is not necessarily the best choice of interac-
tion. Other, "finite range" interactions (e.g., Gaussian e~°(l r l ~ r z D ) are often
used and have relatively larger low-fc amplitudes. Finally, there are often other
residual interactions besides the p-n interaction, such as particle-core cou-
pling contributions that are dominated by quadrupole components. Indeed,
we saw in Section 4.2 that analyses of empirical two-particle multiplets do
suggest evidence for enhanced quadrupole interactions. (We shall see in
Chapter 6 that quadrupole core "vibrations" are the dominant low lying
"collective" modes in nuclei.)
Thus, the parabolic rule is often an excellent approximation. Figure 4.15
shows a few examples taken from Paar and introduces one final but important
point. The (g7;2d3/2) multiple! of J = 2*- 5+ states in 1Z2Sb is well reproduced by
a simple parabola in /(/ +1), as is the (g7/2h11/2) multiple!. 48Sc and 116I also show
multiplets with beautiful empirical parabolic behavior, except that they are
inverted! The reason is well understood. In 2148Sc27, the (iri2^1/2n') multiple!
really a particle-/io/e p-n configuration (ilrip f \/2n) = (iirip f~l7/2n), as is Ihe
(gra/'hj^) configuration in 49116In67. In this case, the residual interaction has
the opposite sign of a particle-particle or hole-hole multiple!. In other words,
it is repulsive, and the J states with high p-n overlap (/„,„, J^J are raised in
energy, while Jn is lowered the most. We note that this change in sign is not
always the case; in Chapter 5 we will see i! as a charac!eristic of even multipole
interactions such as the quadrupole interaction we are considering here. Inler-
estingly, it does not apply lo the ^-interaction (even though we have discussed
its expansion in even multipoles). For odd multipole interactions AE(p-p) or
AE(h-h) is identical to AE(p-h).
In Chapter 5, we will discuss at length the pairing interaction and the
concept of the quasi-particle, which is a state only partially occupied, neither
102 Shell Model and Residual Interactions
Fig. 4.15. Illustrations of the parabolic rule for a quadrupole residual interaction for several p-n
multiplets(Paar,1979).
fully particle nor fully hole. Anticipating that discussion, the occupancy of an
orbit is given by a probability, denoted V2. (the number of particles in the orbit
is (2; + l)^ 2 ). The orbit emptiness is U1. and U2.+ V2 .= 1. Thus, we can rewrite
Eq. 4.28 for the general situation of quasi-particles as
Thus, for a given proton number (particle "1"), (t/2, - V 2 ) is constant but
(f/2z - V22) changes from +1 to -1 as the neutron orbit ("2") is filled over a
The Shell Model: Two-particle Configurations 103
Fig. 4.16. Dependence of the shape of (g'lm hu/2n) multiplet parabolic splittings as a function of
the occupancy. The uppermost curve corresponds to a nearly full hn/2|i orbit and therefore a
hole-hole configuration. The lowest (h-p) curve corresponds to an hn/2ji orbit with only a few
particles and therefore a hole-particle configuration (based on Van Maldeghem, 1985).
104 Shell Model and Residual Interactions
need for prior detailed calculations. For example, let us compare the multiplets
(hn/2)2 and (dM)2. Clearly, from Eq. 4.17, the angle between the orbits in the 2+
state is less in (h,M)2 [6 = 24° compared to 49° for (d^2]. Therefore, the
overlap is greater and the 2+ energy lower so that the 0+- 2+ spacing will be less
(see Fig. 4.3). By comparing 0+- 2+-4+-... intervals, one can sometimes deduce
evidence for particular / components of the wave functions. Care must be
taken in this, however, since 2+ states, even in singly magic nuclei, often have
several significant components, exhibiting "collective" behavior and lower
energies than can be predicted for single two-particle configurations.
Nevertheless, changes in spacings of yrast states (the first states of each J
value) across a sequence of nuclei can sometimes signal the emergence of
particular shell model components.
From the properties of sums over 6 - / symbols, without ever considering the
radial functions /*, we find that by restricting ourselves to odd tensor interac-
tions,
hThat is, the total shift of all the levels in a configuration of two identical
particles in the same orbit is simply related just to the shift of the J- 0 level, for
any odd tensor interaction. Since the 5-function interaction is equivalent to an
odd tensor interaction, it can be expected that Eq. 4.31 will be approximately
satisfied in real nuclei. Finally, we re-emphasize that it is the separation into
angular and radial coordinates that allows results that are totally independent
of the details of the interaction (except, in this case, for the assumption of odd
tensors).
Thus far, we have restricted the discussion to two valence particles. Multi-
particle configurations will be dealt with shortly. It is useful, however, to
consider one simple result that allows a nice generalization of the present
discussion. Using the formalism of the seniority scheme, it is possible to relate
the matrix elements of any two-body operator in the/" configuration to matrix
elements in configurations with fewer particles. For the large class of interac-
tions that conserve seniority (including a 5-function that is a prototype of
short-range interactions) it can be shown that in a multiparticle, f configura-
tion, the more nucleons that are paired off to J - 0, the lower-lying the state will
be. Thus, for n even, the lowest state of the;'" configuration is a J* = 0+ state
with all nucleons paired to 7 = 0. (The &E(j^) for the levels of the configuration
The Shell Model: Two-particle Configurations 105
\f-J) is just a special case of this). This result is sometimes called the pairing
property, and is characteristic of any interaction that conserves seniority.
Another key result is that the order and spacing of the / * 0 states in the /"
configuration with two unpaired particles will be exactly the same as in the j2
configuration. The more realistic situation in which the valence nucleons
occupy several; orbits, jv ;2>... in a multiparticle configuration I/T1,/'"2,- • • /)
with n; even, is just a generalization. By the first result, the / = 0 combination
will lie lower than one in which a pair of nucleons is coupled to / * 0. Thus we
can generalize these conclusions and state that the ground state of even-even
nuclei will always be 0+. Moreover, since the odd J levels of I ;"7) are unaffected
by a 5-interaction (n even), the low-lying positive parity states should all have
even spin and should increase in energy with spin. Remarkably, in view of the
simplicity of the argument, this is an almost universally observed situation.
Furthermore, we noted that in n=- two-particle configurations of identical
particles, I/^T), it is the odd spin levels that are lowered. A simple generaliza-
tion using the seniority scheme shows that the lowest-lying negative parity
levels in even-even nuclei should have odd spin. This is also almost always
observed, and is known as the Talmi-Glaubman rule.
One can go even further. Consider an arbitrary shell in heavy nuclei
consisting of several states (/ orbits) from one shell and one unique parity orbit
from the next shell. Recall from Fig. 3.2 that the highest normal parity /value,
/roax, in each succeeding major shell increases by one. The corresponding
highest j value is jma[ = /max -1/2. (The ;' = /maj +1/2 is the orbit brought down by
the spin orbit interaction into the next lower shell.) The next highest / is
lma - 3/2. The unique parity orbit is the / +1/2 coupling from the next highest
shell, and so has ;unique = /mai + 1 + 1/2 = /max + 3/2. For example, in the 50-82
shell, /mM = 4, so the highest normal parity orbits are the g7/2, ds/2 and the unique
parity orbit is h11/2. In the g^ - hn/2 configuration, our rules (see Table 4.2)
show that the /=;',+ /2 = 7/2 + 11/2 = 9" state is brought lowest and the 2~ is high-
lying at the unperturbed energy. In the d5/2-hn/2 case, the J" = 3~ is lowest. A
little thought shows that this relationship of/ values is generally true for any
shell in heavy nuclei. Therefore, we can obtain, without calculation, two
general predictions. First, the lowest-spin, negative parity, two-particle state
in an even-even nucleus will be a 2~ level: there is no simple spin combination
that gives a 1" level. However, since this 2~ level always results from a/ = M72
normal parity orbit coupled antiparallel to a ;' = I + 1/2 unique parity orbit, it
will always occur rather high in energy. Therefore, the lowest-lying, low-spin,
negative parity state will be the 3" level. This prediction is borne out in nearly
all heavy even-even nondoubly magic nuclei (in closed shell nuclei two differ-
ent shells are involved in n= - excitations). Examples were shown in Figs. 2.5
(210Pb) and 2.6 (Sn). The main exceptions to all these predictions are them-
selves illuminating: they concern nonspherical, or deformed, nuclei in which,
to borrow terminology from a later chapter, one sometimes encounters two-
quasi-particle, negative parity excitations involving a unique parity Nilsson
orbit (e.g., the low-lying 4- state in 168Er). But, as is well known, the nonspheri-
cal character of these nuclei is itself induced by a strong residual quadrupole
interaction: being an even rank tensor, it does not conserve seniority and
106 Shell Model and Residual Interactions
therefore, a priori, one would not expect the preceding arguments to apply.
A final topic to deal with before considering more complex multiparticle
systems is the spectra of hole and particle-hole configurations, where, by the
former, we mean those of the type (/'-") = (/2'+1 -"). We have briefly mentioned
these in our discussion of multipole forces. Here, we obtain a few more explicit
results.
For hole states, a diagonal matrix element of any single particle operator
acting on an n-hole state \j~"JM) will be equal in magnitude to its value in the
n-particle system \j"JM). The sign relation will depend on the odd or even
tensor character of the operator. This is easy to see by considering all the
possible m states of the n-hole configuration. There will be n unoccupied m
states. The expectation value of a single-particle operator Om (which we
denote O for simplicity) will be the sum of its expectation values over all the
particles:
where or" simply indicates the n-hole state, but where the calculation is carried
out over the 2; + 1 - n particles. This is clearly equal to
that is, equal to the sum over the entire shell minus that for the n missing
particles. But, we have shown above that the first term must vanish since the
closed; shell can have no preferred direction in space and, therefore
Note that the n-particle and n-hole states are not exactly equivalent. They
have different total M values. For n particles in a given / state occupying m
states ml,mv...m ,M ="ZMm.. For wholes in the same orbit, one clearly must
have Mh = -Mf (e.g., 1/2 + 3/2 = -[5/2 + (-1/2) + (-3/2) + (-5/2)]). So the above
matrix element can be written (in simplified notation)
where the subscripts indicate the n-hole and n-particle configurations. Using
the Wigner Eckart theorem for a tensor operator O*of rank k gives
and hence we can now relate the matrix elements of any single-particle opera-
tor in states of the same JM for n-hole and n-particle configurations:
The Shell Model: Two-particle Configurations 107
This gives the critical result that the expectation value of any odd tensor single-
particle operator (e.g., magnetic moment) in an n-hole state is the same as the
corresponding n-particle expectation value, while for even rank tensors (e.g.,
quadrupole moment), the two expectation values are the negatives of each
other.
In Chapter 5 we shall encounter more general results of this special case.
This has immediate consequences of great importance. It implies, for ex-
ample, that magnetic moments or dipole transition rates are the same for
corresponding particle and hole configurations, while quadrupole moments or
E2 matrix elements change sign. Therefore, such matrix elements must vanish
at mid-shell. Of course, these features are well known empirically, although
they are partly obscured by configuration mixing effects.
A nice extension of this is to the interaction energies in particle-hole
configurations relative to those in particle-particle ones. Compare a
proton-neutron particle-particle configuration with a proton hole-neutron
particle configuration. The interaction is a product of proton and neutron
tensor operators. The effect of the neutron operator on the neutron wave
functions (a particle in both cases) is the same for both configurations. Only
the proton operator acts on different configurations (particle in one case, hole
in the other) and the above results for one-body operators then apply. Thus,
for an odd tensor interaction (product of odd tensor one-body operators), the
energy shifts AE(j -';„ /) = A£(/p jnf) while, for an even tensor interaction,
AE(jf-ljn /) = -A£(yn /). The U, independence of a quadrupole force leading
to the up-and-down pointing parabolas for p-p (or h-h) and p-h configura-
tions in Eq. 4.30 is an example of this sign change. For the more general case
of a mixed-interaction, the results are slightly more complicated and are given
by
Note that this gives the energy differences for different J values. The energies
themselves can have a constant additive term that is/-independent (but that
can depend on/). Thus, while Eq. 4.34, or the simpler relations for odd and
even tensor interactions, relates p-h and p-p spectra for a given j shell, it does
not relate such spectra for different; orbits in a major shell.
Since, physically, any very short-range interaction cannot give results quali-
tatively different from that of a 5-function (equivalent to an odd tensor
interaction), one expects that the odd tensor result will be closer to that
observed empirically. In other words, the energies (order and spacing) will be
identical in the p-h and p-p configurations. Nevertheless, Eq. 4.34 is very
important because it is so general. It is valid for any two-body interaction in a
//-coupling scheme. Moreover, it has a practical use from a different, empiri-
cal, point of view. Without knowing anything about the actual interaction, the
knowledge of the order and spacings in a p-p configuration allows one to
predict those in the corresponding p-h configuration. A classic example of this
is the comparison of ,738C121 and 1940K21: the former is expected to be a (dMpf7;2J
configuration, while the latter is (d^f^J = (d'^f^J. The comparison is
108 Shell Model and Residual Interactions
included in Fig. 4.7, where the empirical "°K energies were used to calculate the
level energies in 38C1. The agreement is excellent.
Finally, we consider the interaction energies of the (j*f) configuration
relative to those in the (j"J) configuration. By arguments similar to the
preceding in the m-scheme, it is easy to show that the interaction energy for any
two-body operator in an n-hole state will be equal to that in the «-particle state
plus an additive constant. Therefore all spadngs in n-hole configurations
\i~nJ) should be identical to those in the corresponding n-particle states \ff).
Note an important point analogous to a comment we stressed in regard to the
effect of closed shells on open shells in Chapter 3. As used here, the term shell
refers to a single ; shell, not a major shell consisting, generally, of several;
values. The additive constant relating «-hole to n-particle configuration need
not be independent of/. Therefore, the full set of spacings in a configuration
such as jini jl "2 ---j] will not generally equal those in j\l j"2 • • • J\. The equal-
ity of spacings applies to each /, individually. If the realistic wave functions
contain admixtures of several / values, there is no simple general relation of
particle and hole energies.
A final point: thus far we have only discussed diagonal matrix elements
(energy shifts) due to a 5-interaction, but off diagonal effects are equally
important. These, of course, give rise to the mixed wave functions character-
istic of realistic shell model calculations. Expressions for the off-diagonal
matrix elements (JJ2\S\ /3;4) are similar to Eq. 4.7, except there is now a
second 3 -;' symbol involving/,, ;4. The strength of a given matrix element
then depends on the angular correlation of the two particles in both initial and
final states, as well as on the radial overlaps.
109
110 Shell Model and Residual Interactions
A=7/2 ;2=7/2
m, m
2
Af /
7/2 5/2 6"
7/2 3/2 5
7/2 1/2 4
7/2 -1/2 3 6
7/2 -3/2 2
7/2 -5/2 1
7/2 -7/2 0.
5/2 3/2 4"
5/2 1/2 3
5/2 -1/2 2 4
5/2 -3/2 1
5/2 -5/2 0
3/2 1/2 2"
3/2 -1/2 1 2
3/2 -3/2 0.
1/2 -1/2 0] 0
"Only positive total M values are shown, the table is symmetric for M < 0.
For the (ds/2)3 case just discussed, this gives Jnm = 3; - 3 = 9/2.
Thus, we see an example of how a consideration of the possible m substate
occupations can give information on permissible total angular momenta J.
Basically, the m-scheme is a systematic set of procedures for doing this in a
general case.
The m-scheme is best described with a detailed example. Earlier we
showed that only even J values were allowed in the j1 configuration of identical
nucleons. This was argued in terms of the symmetry properties of spherical
harmonics. However, it can also be seen by inspecting the possible magnetic
substates. Table 5.1 summarizes the allowed magnetic substates for a two-
particle configuration (7/2)2. (Actually, the M < 0 cases are omitted since they
are completely symmetric to the Af > 0 cases.) We construct such a table
starting with the highest magnetic substate for particle 1 and list all of the
possible substates for particle 2 allowed by the Pauli principle, then carrying
out the same procedure for the next lower magnetic substate (in this case 5/2)
for the first particle. It is necessary to recall that the two nucleons are indistin-
guishable so a combination 5/2,7/2, for example, is not allowed, since the 7/2
and 5/2 combination has already been listed. Continuing in this way, we obtain
all of the possible m values for the two-particle system. Since a given total
angular momentum / must have magnetic substates M = J,J- !,...-(/ -1), -J,
it is now trivial to deduce the possible final / values using the table. In the
present example, there must be a / = 6 state, since there is a M = 6 configura-
tion. This 7 = 6 state has magnetic substates M = 6, 5, 4, 3, 2, 1, 0, and the
corresponding negative values. Thus the top seven magnetic substates listed in
the table must be used up for this single / = 6 state. There is no M = 5 magnetic
substate left over and thus there cannot be a J = 5 state. There can, however,
be a / = 4 configuration that consumes the M = 4,3,2,1,0 magnetic substates
Multiparticle Configurations 111
"The full set of allowable m. combinations that give M > 0 are obtained by the conditions nij > 0, m3 < m2 < mt
and no two m.values identical.
listed next in the table. By continuing this argument, one sees that there is no
J = 3 state, but there are a / = 2 and a J = 0 state, thus proving in a different way
the result obtained earlier that two identical particles in the same orbit can
couple only to even total angular momenta /.
The case for (5/2)3 configuration is shown in Table 5.2, which will not be
discussed in detail although the reader may go through the example and verify
the results just as we did for the (7/2)2 case. Clearly, for multinucleon configu-
rations where «is large, this procedure can be lengthy. Other techniques are
available. However, the m-scheme is important because it shows in a transpar-
ent way how the physical effects of the Pauli principle arise.
We noted above that the m-scheme gives a rule for the maximum permissible
/value in a/" configuration of identical particles very simply. It also gives, in an
equally simple way, the result that a/1 configuration can never have a state with
/ = Jm!at - 1. We will show this only for the case of two particles, but the
generalization is straightforward. (Although the following considerations are
general, reference to the specific example in Table 5.1 will clarify the argu-
ments.) The maximum Jinaf configuration is /mM = 2; -1. A state J = /max - 1
would have (if it existed) J = 2(j -1). One value of M = Jmia - 1 must be used
for the J - /mM state. Therefore, in order to have a state with J = Jmax -1 = 2(/ -1),
there must be a second permissible M = Jma - 1 = 2(/ - 1) state. This cannot
involve a particle in an m = j state, since that state is already consumed for the
/ = /max level. Therefore, the only way to make another magnetic substate
M - 2(j - 1) is to have two particles with m - j - 1. But this violates the Pauli
principle, and therefore is impossible, proving that a / = /mai - 1 state never
exists in a j2 configuration. As noted, this can be generalized to the ;'"
configuration.
on the energies of such / states and dynamic matrix elements involving such
configurations.
Clearly, when considerations such as those discussed in Chapter 4 are
attempted for multiparticle configurations, the situation rapidly becomes
much more complex. To see this in a simple example, consider a (d5/2)3
configuration of identical nucleons. A J = 5/2 state can be made in three
distinct ways by first coupling two particles to an intermediate J' = 0,2, 4:
and
However, the (5/2)3 configuration has only one J = 5/2 state (see Table 5.2). Its
wave function must therefore be a totally antisymmetric linear combination of
these three basis states. The normalized coefficients in this linear combination
are called coefficients of fractional parentage (CFP): their squares give the
probability that a given final state is constructed from a specific "parent"
configuration—in this case, a two-particle state. The relative magnitudes of
the three CFP's for the (d5/2)3 configuration are not arbitrary, but are given by
certain angular momentum coupling coefficients. CFP coefficients can be
constructed not only for three-particle configurations, but for any n.
Unfortunately, the complexity of the possible couplings makes the notation
for the coefficients rather complex and this has deterred many nuclear physi-
cists from delving into the subject; the formalism can be terrifying. Formal
textbooks can be filled with page after page of long, daunting expressions in-
volving sequences of CFP coefficients, angular momentum coefficients, sum-
mations over them, and the like. Here, we attempt to cut through much of this
by summarizing some of the essential results and their motivations with a few
simple examples. This is not entirely satisfactory, since it deprives the reader
of an appreciation of the beauty and power of the formalism. Moreover, as the
author can personally attest, while a simple presentation of the final results
saves the reader from the tedium of struggling through their derivation, it also
confers on them an element of mystery—the reader is left with a sense of
wonder at how one can start with such general interactions and general
configurations and end up with very simple final results. He or she may glance
back over the imposing derivations in the hopes of seeing where some simpli-
fying assumption or some restrictive case has been invoked. The real power
and beauty of the method, however, is that such assumptions are usually not
required: very general results that enormously simplify the treatment of many-
particle shell model configurations can often be obtained.
In any case, we will introduce the formal notation for CFP coefficients, but
will avoid their manipulation as much as possible. We will derive one simple
result that illustrates their power and economy.
Multiparticle Configurations 113
and determines, as in the preceding three-particle case, the probability that the
wave function | fJM1) can be written in terms of the (n-2)-particle configuration
\j"^(Jn_2Mn 2)> coupled to a two-particle configuration |/172M2). The defining
equation is therefore
Clearly, then,
We can see how the concept of CFP coefficients and the parentage of n-
particle configurations is useful. Consider a configuration \ff) and ask what
the energy shifts, AE(jnJ), are for each final /value for an arbitrary interaction.
(Note that, as usual, we drop the magnetic quantum numbers to simplify the
notation.) First, we note that, since the particles are indistinguishable, the total
interaction energy in any final state J is given simply by the interaction energy
for any pair of nucleons (say, particles 1 and 2) times the total number of
possible pairs n(n - l)/2. However, the two-particle matrix element Vn can
only depend on 72, and not on the way in which 72 is coupled with Jn 2 for the
other n-2 nucleons to give the final /. The total interaction energy for particles
1 and 2 is just the sum of the interaction energies for each two-particle angular
momentum ]^ multiplied by the probability of each 72 in the state \ff). We
denote this probability W(fJJ2). Thus, we can immediately write the interac-
tion energy from particles 1 and 2 in the state I/1/) as:
where the W coefficient is the sum of the squares of the CFP coefficients for a
given J2 over all possible values Jn 2. That is,
The interesting point here is that there are in general fewer values of 72 than
there are of 7. For example, in the (7/2)3 configuration, J can be 15/2, 11/2,
9/2,7/2,5/2,3/2 while (7/2)2 can only couple to 72 = 0,2,4, and 6. Thus, by Eq.
5.2, the six energies of the configuration j"J are given in terms of the four
matrix elements
114 Shell Model and Residual Interactions
Fig. 5.1. Comparison of the low-lying empirical levels of 51V with calculations obtained by
coupling an (J/2 proton to an (f?ffl)2 two-particle configuration (right) and by coupling an f?/2 proton
to the empirical levels of 5aTi. (See deShalit, 1974.)
The beauty of this is that these matrix elements are usually easy to calculate
for a known interaction and, even when the interaction is not known, empirical
values for them can be obtained from the neighboring even-even nucleus
(with n = 2). This can then be used to calculate the energy levels of the adjacent
odd mass nucleus.
We have discussed the (7/2)3 example here because it is treated in detail in
de Shalit and Feshbach, where the low-lying (f7/2)3 energy levels of 51V are
calculated in terms of the empirically known (f7/2)2 levels of 50Ti (0+:0,2': 1.55,
4+:2.68, 6+: 3.2 MeV). The results are shown in Fig. 5.1; the agreement is
remarkably good for such a simple approach. Note once again that nowhere in
this discussion has any aspect of the interaction been specified, except to
assume that it is two-body only. We could also have calculated 51V with the
same formulas using a 5-function interaction to simulate 50Ti, that is, to define
the (f?/2)2 matrix elements. Normalising the 8-function strength to the 0'-6f
spacing in 50Ti gives calculated 50Ti energies of (M), 2+:2.68, 4+:3.0, and 6':3.2
MeV. These have a different distribution than the empirical levels and, when
applied to 51 V, give the fit on the right of Fig. 5.1. Clearly, this approach is not
Multiparticle Configurations 115
nearly as successful. The point is that the empirical 50Ti spectrum automatically
includes all relevant interactions in the (f,^)2 system. The CFP techniques
relate this directly to51V, independent of a knowledge or guess of the interac-
tion. Thus, an understanding of the makeup of an n-particle configuration in
terms of its (n-2)-particle structure can greatly simplify the treatment of
nuclear spectra in complex systems. The present results can be generalized to
n > 3, and provide comparable, and even greater, simplifications.
(f^)4 configuration. From the m-scheme and the simple formula derived
earlier, Jnm = 4j - 4(3)12 = 8. This state can only be made by maximizing the
alignment of all; = 7/2 angular momenta as allowed by the Pauli principle. The
7 = 8 state therefore has seniority 4; there are no particles coupled in pairs to
/ = 0. On the other hand, /=2,4, and 6 states can be made by first coupling one
pair of particles to/ = 0 and then using the remaining | (7/2)2/) configuration to
produce angular momenta of 2,4, or 6. Such states have seniority v = 2. Finally,
the J = 0 state of the (f7/2)4 configuration obviously has seniority 0, that is, all
particles are coupled in pairs to / = 0. (Note that there may be other J = 0,2,
4,6 states of the (f^)4 configuration, all with v = 4.) What we have shown is that
/ = 0,2,4,6 states of v = 0 or v = 2 can be constructed.
The seniority concept is important for several reasons. First, it leads to
many simple, powerful results under very general conditions. For example,
various interactions and matrix elements can be classified in terms of whether
or not they conserve seniority. As will be seen, they have very different
properties as the number of particles in a shell increases. Secondly, and
perhaps most importantly, it seems that many realistic residual interactions
conserve seniority, so this scheme gives reasonable predictions for actual
nuclei. It is impossible within the scope or philosophy of this book to derive all
the results of the seniority scheme without adding an undesirable complexity.
Such derivations are available in many detailed textbooks on the shell model.
The complexity of these derivations often tends to obscure some of the simple
ideas lying behind them. It is these ideas that we wish to emphasize here. We
will derive or motivate a few crucial results; the others can be obtained by
analogous, though more tedious, manipulations.
Perhaps the most important ingredient in understanding the results of the
seniority scheme is the following: consider the / 2 configuration and the matrix
element of any odd tensor interaction. (The introduction of the concept of
tensors and their rank here should not be intimidating. The spherical harmon-
ics of order k, Y^ simply form the 2k + 1 components of a tensor of rank k. An
example of an odd rank tensor is the magnetic dipole operator. The quad-
rupole operator is an even rank tensor. As commented eariler, the 5-function
interaction is equivalent to an odd-tensor interaction.)
For the case of a one-body odd-tensor operator acting in they 2 configuration
The proof of this is trivial. We recall that in the two-particle configuration only
even J values are allowed. Therefore, J on the left side must be even and, by
conservation of angular momentum, there is no way that 7 = 0 can be coupled
to an even J by an operator carrying odd multipolarity.
Equation 5.4 simply states that all matrix elements of one-body odd-tensor
operators vanish in the y2 configuration. This includes the 7 = 0 case. Odd
tensor operators cannot "break" a/ = 0 coupled pair, nor can they contribute
a diagonal "moment." The significance of this simple equation cannot be
overemphasized.
In many-particle systems, it has three enormously important consequences.
For such configurations, one-body operators are normally expressed in terms
Multiparticle Configurations 117
seniority scheme, but rather connect states with seniorities v and v ± 2. Using
arguments such as these, it is therefore clear why Ml transitions in even mass
nuclei are rare—they can only connect states of the same seniority—while E2
transitions dominate even in near-closed shell nuclei. Therefore this domi-
nance is not necessarily a demonstration of collectivity, but a reflection of the
seniority structure of low-lying states in ff) configurations.
Thus far in our discussion of seniority, we have considered single-particle
operators representing moments or transitions. Equally important are two-
body interactions, which can be either diagonal or nondiagonal. Both are
important, although we will emphasize the former since they determine the
contribution of residual interactions to level energies. A key example is the 5-
function interaction. Clearly, interactions can be written as products of single-
particle operators. We saw an example of this earlier in discussing multipole
expansions of arbitrary interactions. We now turn to consider the properties of
various interactions in the seniority scheme.
Consider an arbitrary odd-tensor two-body interaction V12. This can be
taken as a product of one-body operators, Stodd fj*f2*. As with one-body
operators, it is extremely useful to be able to relate the two-body interaction
matrix elements of seniority v states in the f configuration (n even) to the
matrix elements in a f configuration. Deriving this desired result is trivial.
Consider the matrix element (a subscripted k labels particles, not rank)
where the sum is over the n-particles, and where a and </ denote any addi-
tional quantum numbers needed. Since the states have seniority v (even),
there are (n - v) particles paired off to / = 0. By the same reasoning that led to
Eq. 5.4, the terms in "I. _, kVa, that act on these particles cannot change their
coupling. All that this part of the sum can do is contribute a diagonal matrix
element of the form Q2J = 01 V.J // = 0). But this is just the lowering of the 0+
energy in a;2 two-particle configuration. We define this energy lowering by V0.
The sum contributes this for each such (J - 0)-coupled pair, of which there are
(n - v)/2. Having thus separated off these particles, we are left with a sum over
v particles of the same interaction. Thus, we obtain,
This interaction matrix element may be either diagonal or nondiagonal (in a),
but it cannot change v since it is of odd tensor character. In either case, it is of
absolutely central importance in nuclear spectroscopy. As with the case of
one-body odd-tensor operators, we have an equation relating matrix elements
of a two-body interaction in they" configuration to those in the/" configuration.
Here, however, these matrix elements are not constant across a shell, but linear
in (n - v)/2, the number of nucleons paired off to J = 0. Such matrix elements
peak at midshell. This feature is sometimes known as the pairing properly.
To understand other important implications of this, let us first consider
Multiparticle Configurations 119
diagonal matrix elements where a = «'. The second term on the right in Eq. 5.6
is simply the number of pairs of particles coupled to / = 0 multiplied by the
interaction energy, V0, for each pair. Recalling that we are dealing with
attractive residual interactions (larger matrix elements imply lower-lying
states), then states with lower seniority v will lie lower in energy. The v = 0
states, which must have /* = 0+, will lie lowest. Immediately, this accounts for
the well-known empirical property that the ground states of (spherical)
even-even nuclei all have /* = 0+.
Similarly for odd mass spherical nuclei, the ground state will usually be a
v - 1, J=j state in which all but one nucleon is paired off in \fj = 0) combina-
tions.
It is worthwhile to explicitly write Eq. 5.6 for a/" configuration in the v = 0,
7 = 0, and v = 1, J = j states. For both situations the first term vanishes since
there cannot be a two-body interaction in a ;v=0 (no particle) or ;'v=1 (one
particle) system. Therefore, the energies are given by the second term:
These equations simply state that the ground state energies in the respective
systems depend solely on the numbers of pairs of particles coupled to J = 0. In
the odd particle case, the unpaired nucleon is, from this point of view, just a
spectator. Indeed, as de Shalit and Feshbach emphasize, the nuclear force
effectively measures the number of pairs of particles coupled to J = 0, at least
insofar as it can be approximated by odd tensor interactions.
One of the most crucial uses of Eqs. 5.6 and 5.7 concerns the energies of
seniority v = 2 states (the following argument applies to higher seniority states
as well, but these are less often identified experimentally). Let us consider the
energy difference E(J"v = 2,7) - E(j"v - 0, / = 0). Simplifying the notation by
denoting the interaction by V, Eq. 5.6 and 5.7 give
Once again, the power of the seniority scheme allows us to link matrix
helements in the configuration/" to those in the configuration;". The square of
Eq. 5.13 gives the behavior of the transition rates induced by the operator
throughout a shell. For large/ and n (/', n » v), this transition probability goes
as (f(l - /)) where / = nl(2j + 1) is the fractional filling of the shell. This
expression at first increases as/, then flattens out, peaking at midshell. More-
over, it is clearly symmetric about the midshell point. Probably the most
common and important application of this concerns E2 transition rates in-
duced by the operator Q = r2Y2. The important quantity (2 + I Q I IO^) 2 is
proportional to the E2 transition rate from the first 2+ state to the ground state
in an even-even nucleus, and can be written for the /" configuration as
[assuming the 2 +(0 +) state has v = 2(v = 0)]:
For shells that are not too filled, so that (2/ ± 1) » n, this becomes
is just proportional to the number of particles n in the shell, for small n. For
large n,n-> 2j + 1, it falls off, vanishing, as it must, at the closed shell. For
/, n » 2, we see that, as given in the general case above,
This behavior is commonly observed in real nuclei, with B(E2:21+ -> 0^)
values rising to midshell and falling thereafter. Data beautifully illustrating
this are shown for the Z = 50 to 82, N = 82 to 126 region in Fig. 5.3. (The peak
regions of the B(E2) values in Fig. 2.16 are additional examples of this in
condensed form.) In part, this behavior is due to coherent effects involving
single-particle configuration mixing of different/ values in the wave functions
122 Shell Model and Residual Interactions
Fig. 5.3. Saturation of empirical B(E2) values in the rare earth region that illustrates Eqs. 5.13 and
5.16. The numbers on each line give the neutron number.
for each particle, but the overall behavior still reflects a generalization of this
simple result for the seniority scheme.
For transitions induced by even-tensor operators of rank k > 0 that do not
change seniority, the expression corresponding to Eq. 5.13 is
zation to many-/ shells suggests that such moments should have opposite signs
at the beginning and end of major shells. Although the trends in realistic cases
are complicated by the different; shell degeneracies, this qualitative feature is
a well-known empirical characteristic of heavy nuclei, and it contrasts mark-
edly with that for odd-tensor operators that are independent of n (see Eq. 5.5).
We shall see another important application of Eq. 5.17 in our discussion of
the p-n interaction in Chapter 6. Moreover, the decrease toward midshell and
symmetry about that point will see important reflections even in deformed
nuclei (systematics of $ Fig. 6.11, p. 164) where seniority is strongly broken.
Finally, we turn to two-body interactions for even-tensor operators. Some
of these interactions can change seniority, connecting states with v and v -2.
For this case, the result is trivial to derive. An even-tensor two-body interac-
tion connecting states with seniorities v and v - 2 must be a product of two one-
body operators—one that conserves seniority, another that connects v and
v-2. In the reduction to a matrix element in the/" configuration, the first gives
a factor identical to that in Eq. 5.17, the second gives the factor in Eq. 5.13.
Thus, their product yields the result
Once again, we note the factor (2/ +1 - 2n), which vanishes at midshell and has
opposite signs in the first and second halves. For;', n » v, this interaction
energy varies across a shell as (1 - 2/)/(l -/): at first this increases with/but it
peaks well before the shell is one-quarter filled, tapers off, and crosses zero at
midshell; in the second half of the shell, it is symmetric to the first half except
for a change in sign.
We will not give the general expression for seniority-conserving matrix
elements in the f configuration since they are more complex, involving not
only matrix elements of the interaction in the f but in the ;v+2 configuration as
well.
It is useful at this point to summarize some of these important results. This
is done in Fig. 5.4, which shows the behavior of both seniority conserving and
nonconserving matrix elements for one-body operators and two-body interac-
tions across a/ shell under the assumption (where applicable) that; and n are
large and much greater than v. For the v -> v - 2 even-tensor case the square
of the matrix element is given since it is directly proportional to the most
common example of such behavior, B(E2: 2^ -> O/) values. Each panel also
gives the (sometimes approximate) analytic formula.
To recapitulate: one-body odd-tensor operators (e.g., magnetic moments)
conserve seniority and are constant: one-body even-tensor operators may
change seniority, with v —»v - 2 transition matrix elements (e.g., B(E2: 2^ —»0, + )
values) peaking at midshell, while seniority conserving matrix elements (e.g.,
quadrupole moments) vanish at midshell and are negatives of each other for
124 Shell Model and Residual Interactions
Fig. 5.4. Summary of the behavior or various operators and interactions across a shell in the
seniority scheme. Note that the middle panel gives the square of the matrix element since this
corresponds to the physically interesting case of B(E2) values.
Thus far, the discussion has focused on the 5-interaction. A very popular
alternative is the surface ^-interaction (SDI), which, as its name implies, acts
only at the nuclear surface. It is equivalent to the angular part of the 5-
interaction and to the assumption that all radial integrals are equal. Though
this is a simplifying assumption, it permits an important generalization of one
of the key preceding results: for degenerate orbits, the SDI conserves seniority
in the multi-;' configuration I/'"1;'"2- • • » v , j } . (In a single;' shell the 8 and SDI
interactions are identical.) The SDI also has off diagonal matrix elements in
multi-/ situations that give rise to mixed wave functions. These matrix ele-
ments are generally larger than for the volume 5-interaction: the reason is
simply that, when the interaction can occur throughout the nuclear volume,
the effect of non-complete overlap of the particle wave functions is larger.
why seniority is often a reasonably good quantum number even in rather com-
plicated configurations.
In any case, the theorem implies that seniority must be a good quantum
number for the (f^)" states in the Ca isotopes, independent of the interaction.
This does not mean that the interaction energies are independent of the
interaction since, as we saw, the matrix elements of these interactions depend
on their odd- or even-tensor character. Nevertheless, knowing that seniority
must be a good quantum number, we can confidently choose some reasonable
interaction and inspect the predictions of the seniority scheme. We assume
that an odd-tensor interaction is a good choice since the 5-function can be
written in this form. Then, from Eq. 5.6 or, more explicitly, from Eqs. 5.9 and
5.10, the excitation energies of the / = 2+, 4+, and 6+ levels should be constant
across the Ca isotopes.
Let us inspect the data in Fig. 2.3. The yrast energies are high at 40Ca and
48
Ca since both of these nuclei are doubly magic. However, "^.^Ca have, as
expected, much lower 2+l energies and these are indeed relatively constant.
The B(E2) values in the Ca isotopes are also roughly consistent with the
seniority picture since they peak near midshell, although the symmetry about
midshell is not particularly evident.
The constancy of v = 2 level energies in singly magic nuclei has long been
emphasized by Talmi. His classic example is the Sn isotopes, whose energies
were shown in Fig. 2.6.
One can also use the seniority scheme to look at nuclear binding energies,
B.E.^), (n even) for a series of nuclei in which an orbit / is filling. These
binding energies are the absolute energies of the I f , v - 0, J = 0+> ground states
(we assume an odd-tensor interaction). Of course, this is only the residual
interaction energy, to which must be added the n single-particle energies Enl..
To be slightly more general, there is one other interaction that is diagonal in
the seniority scheme—the trivial case of a scalar interaction. Since this is a
constant, the interaction matrix element will be equal to that constant multi-
plied by the total number of pairs of nucleons that can interact, n(n - l)/2.
Combining these three terms, one then has for the binding energy in a /"
configuration, relative to the closed shell,
This formula, which is valid for any interaction that is diagonal in the seniority
scheme (and therefore, for any interaction for; < 7/2) displays the well-known
parabolic behavior of nuclear masses. We note that the parameter B always
turns out to be negative: the quadratic term is always repulsive. Evidence of
this can be seen empirically from the data presented in Chapter 1, which
showed that binding energies are approximately proportional to A. If B > 0
(attractive), binding energies would increase quadratically, instead of linearly,
with A.
This is the same point we have made before, that the nonpairing, residual
interaction between like nucleons is repulsive. We can now carry this one step
further. Since the quadratic term stems from the scalar part of the interaction,
Multiparticle Configurations 127
Thus, these excitation energies are identical to those in the/ 3 configuration and
are independent of n—that is, they are constant across a shell. Clearly this also
means that the spacings between v = 3 states are n-independent. This can of
course be seen by explicit calculation:
Fig. 5.5. Low-lying levels of 142Nd in comparison with analytic and numerical shell calculations.
(The empirical level scheme is based mainly on Wirowski, 1988.)
To start, we note that the lowest shell model orbits in the Z = 50-82 shell are
2d5/2 and Ig^. There are many ways that the protons can be distributed over
the two positive parity orbits. If we assume that the splitting of states of a given
seniority v is small compared to the spacings between states of different
seniorities (we are neglecting the first term relative to the second in Eq. 5.6),
then the v = 0 / = 0+ state will, as usual, lie lowest, by an amount V0 below the
v = 2 states. The v = 2 states, in turn, occur, on average, this same distance
below the v = 4 states. Therefore, we can assume that all of the J * 0 low-lying
positive-parity states (J* < 6+) are v = 2. Since we are dealing with (positive
parity) j values < 7/2 we know that seniority is a good quantum number
regardless of the interaction. Since I (ds/2)" v = 2} does not give a 6+ state, let us
assume for simplicity that the ds/2 shell is filled, leaving four protons in the g7Q
orbit. By Eq. 5.6 and the preceding discussion, the J = 2+, 4+, 6+ v = 2 excitation
energies should be identical to those in a (g7/2)2 configuration, and as such can
be estimated theoretically using the results for a 5-function interaction. We
postpone doing so for a moment until we determine how best to estimate the
absolute strength of the interaction. The next positive parity state is 8+ and
must have v = 4. By our earlier arguments it should lie higher than the v - 2
states by roughly \V0\. The lowest-energy way to construct it is to couple the
two v = 2 states with/^ 6+ and 2+ together. Similarly, the 10f level must also
have v = 4. The easiest way to form it is by coupling the J" = 6+(v = 2) and
j" = 4'(v = 2) states.
Multiparticle Configurations 129
The negative parity states must involve the hn/2 orbit. The two lowest con-
figurations should involve eight protons coupled to / = 0 in the d5/2 and g7/2
orbits and an (h^ g7/2) or (hn/2ds/2) pair. The former gives spins/ = 2~ - 9~ while
the latter yields / = 3~ - 8\ From the rules developed in Chapter 4 for the
ordering of different / states in two-particle configurations under the influence
of a 5-interaction, we find that the 9" state should be the lowest in the
| hjj^gj^/} configuration while the 3~ should lie lowest in the I hn/2d5/27) con-
figuration. In both cases the even-spin negative-parity states are unaffected by
the interaction. Empirically, the lowest negative parity state is indeed 3~, sug-
gesting the (h^d^) assignment. The lowest T and 8" levels must also belong
to this multiplet (since, if they were part of the (h^g^X/ = 2~ - 9~) multiplet
they would lie above the 9~). The 8~ level gives the unperturbed position of the
(hu/2d5/2) multiplet, and the energy difference 8"- 3~ gives the absolute scale of
the interaction strength, thus allowing us to predict (using Table 4.1) the 5" and
T energies, as well as the spacings among the positive parity levels. The J = 9~
state is then the lowest member of the (hn/2g7/2) multiplet. (The lower-spin
members would not have been detected in experiments carried out for 142Nd.)
Finally, the J* = 10~ and 11- states cannot arise from either (hn/2g7/2) or
(hn/2d5/2) couplings. Since, empirically, the latter lies lower, a reasonable con-
figuration for the 10" level is I (hn/2d5/2)/ = 8~ ® 2+, )10~, meaning an (hn/2d5/2)
pair coupled to /* = 8" built on the seniority v = 2 2+ level of the remaining (n-2)
particle system. The energy difference £(10~) - £(8") ~ £ 2 J is approximately
satisfied experimentally. The lowest 11~ levels can be made either by coupling
the h1]/2g7/27 = 9")state to the v = 22+l level or the hn/2d5/2./ = 7~) level to the
v = 24\ level.
All these results are incorporated now into Fig. 5.5, where it is evident that
the agreement of this extremely simple calculation with experiment is actually
remarkably good. The ordering and energies of most of the levels are correctly
predicted analytically, in agreement with experiment, and with only three
parameters: the single-particle energy differences ehn( 2 -e g 7/ 2 and ed 5 | 2 -e g7 / 2
and the strength of the residual interaction. Figure 5.5 also shows an actual
detailed 10-particle diagonalization of the 142Nd level scheme, using a surface
5-interaction (which gives relative spacings, within a configuration Iff), which
are the same as for a volume 5-interaction and which is otherwise similar as
well, although different in some details) with strength 0.4 MeV and with the
single-particle energies (in MeV): eg7|2 =0, EdS|2 =0.7, £ im/ 2 = 2.5. The
calculation also shows reasonable agreement with the empirical scheme, but
more importantly, it shows that our simple analytic interpretation is a re-
markably accurate approximation of a complex shell model diagonalization.
This kind of interpretation highlights the power of the methods we have
discussed and shows how far we can go in a relatively simple shell model
interpretation of rather complex level schemes. The principle difference in the
numerical diagonalization is that additional components, such as (d5;2)2 and
(d^^) or even (ds/2g37/2), come into play. Similar analyses can be applied to
countless other nuclei (e.g., Sn) and greatly help to understand the results of
complex realistic calculations.
Improved calculations, compared to Fig. 5.5, have also been carried out.
130 Shell Model and Residual Interactions
They allow (hn/2)2 as well as (ds/2)nl and (g7/2)"2 configurations in the positive
parity levels. The inclusion of these amplitudes mainly affects the required
hj^ single particle energy and, of course, changes the strength of the interac-
tion needed to fit the data. The reason is easy to see from Eq. 4.10. Since the
lowering of the 0+ state in a configuration \j 2J) is proportional to (2/ + l)/2
multiplied by the interaction strength, V0, the inclusion of (h11/2)2 means that a
smaller \V0\ is required to maintain the same / rast- /0+ spacing. The extra
lowering of the 0+j level effectively raises the excitation energy of all the
others. To regain a fit to the data for the negative parity states, a lower ftn/2 is
required.
This illustrates two important points. First, the choice of effective residual
interaction, single-particle energies and the shell model space are intimately
linked. One should be wary of conclusions regarding any of these if there is not
supporting evidence for the choices concerning the others. Secondly, despite
the beauty of analyticity, a realistic treatment of complex nuclei still requires
detailed explicit calculations if really quantitative results are desired.
Finally, note that we have completely ignored core excitations of the pro-
tons or neutrons. These can be significant and their effects can vary signifi-
cantly for different states. In particular, even in singly magic nuclei (e.g., 142Nd
or Sn), the lowest 2+ and 3" states are often rather "collective" with a number
of major components, including core excited particle-hole components (e.g.,
(g^-1 hn/2) / = 3~ for neutrons in Nd or protons in Sn). To some average extent,
their ignored effects are mocked up by the choices of the single-particle
energies of the valence orbits and of the interaction strength. It is no wonder
that residual interactions are often called "effective interactions" and that an
extensive theory of such interactions has been built up. Indeed, as discussed at
the end of Chapter 4, a number of alternates to the <5-function interaction are
often used. Often, these are more structured, finite-range, interactions such as
Gaussian forms. Various so-called Skyrme interactions are also popular, as are
interactions defined explicitly in terms of sets of two-body empirical matrix
elements. It is clear from the discussion at the end of Chapter 4 that such finite-
range interactions have different (often more parabolic) ./-dependencies than
the 5-function.
heavy nuclei to describe such behavior more directly in terms of the so-called
pairing interaction. This is defined (see Eq. 4.20) to be an attractive interaction
acting only on two identical particles in total angular momentum 0+ states.
States with /* * 0+ are unaffected. A comparison of the pairing interaction with
a 5-function interaction for an (f7/2)2 configuration was shown in Fig. 4.13. The
sequence of energy levels for the 5-function interaction is the familiar result we
have been studying and simply reflects the properties of the 3 -/ symbol in Eq.
4.10. The pairing force presents a similar overall pattern but with a degenerate
multiplet of J* * 0* states at the unperturbed energy. As we have seen,
inspection of the empirical level schemes of two-particle states in singly magic
nuclei such as 210Po, 210Pb, 134Te (Fig. 4.5), and many others, shows that the 5-
interaction reproduces empirical spectra much more accurately. Even for
multiparticle (but still singly magic) cases, such as the Sn isotopes, the 5-
function provides a more realistic interpretation (see Fig. 2.6). Nonetheless, it
has become nearly universal practice to invoke a pairing interaction and to
refer to the lowering of the 0+ state as a pairing effect. (Indeed, we already
referred to the second term in Eq. 5.6 as the "pairing property" in analogy with
the effect of the pairing force.) Part of the appeal of the pairing force is the ease
with which it can be extended to multiparticle systems, where it leads to the
desired result of 0+ ground states in even-even nuclei without the need for the
seniority apparatus and angular momentum algebra we have discussed.
There are a number of experimental facts that motivate the introduction of
a pairing force and the concept of pairing correlations. We have mentioned
some, but it is useful to summarize them here. The best known is the simple
fact that the ground state of all even-even nuclei has J* = 0+. A related point is
that this 0+ state is normally far below other noncollective intrinsic states. This
is the so-called pairing gap.
Perhaps the most direct evidence for a pairing interaction is the so-called
odd-even mass difference. This simply refers to the fact that when nucleons
are added to a nucleus, the gain in binding energy is greater when an even-even
nucleus is formed than when the neighboring odd mass nucleus is formed. This
empirical fact can be inferred from the data of Figs. 1.2 and 1.3 by comparing
the absolute values of S(n) or S(p) for adjacent odd and even nuclei. An extra
attractive interaction that couples pairs of like-nucleons together can accom-
modate this fact, and indeed, the separation energy data suggest a strength for
the pairing interaction of -1-2 MeV.
There are three other features that can be seen, at least in retrospect, as
clearly pointing to the need for a pairing interaction. The pairing interaction
clearly favors sphericity since it favors the formation of pairs of particles
coupled to a total magnetic substate M = 0. Therefore, near closed shells, the
presence of a strong pairing interaction will inhibit the tendency to deform.
Instead of the smooth transition toward deformation that would normally
occur as valence nucleons are added, one typically sees, empirically, a se-
quence of more or less spherical nuclei followed by a rather rapid transition
region to deformation. Secondly, for a deformed nucleus of a given shape, it is
easy to calculate the moment of inertia. This in turn determines the spacing of
132 Shell Model and Residual Interactions
rotational levels (see Chapter 6). Empirical moments of inertia extracted from
those energies are systematically lower than calculated in the shell model. The
inclusion of pairing correlations essentially removes the discrepancy. Finally,
the energy distribution of Nilsson orbits near the ground state in deformed odd
mass nuclei is, on average, denser than expected from the Nilsson diagram.
This will be seen to be a "compression" effect that occurs in the transformation
from particle-hole to quasi-particle energies.
The pairing interaction was defined in Chapter 4 in terms of its matrix
elements as:
where G is the so-called strength of the pairing force and the rest of the
expression has an obvious meaning. Note that the pairing force is independent
of orbit but, since it is identical for each magnetic substate, scales for an orbit
/ as (2/ + 1). It is therefore stronger in high j orbits. Although G is orbit
independent, it decreases with A in heavier nuclei where the outer nucleons
are generally further apart and so spatial overlaps are likely to be less. G may
be different for protons and neutrons, being lower for the former because of
Coulomb repulsion. Commonly used prescriptions are
Fig. 5.6. Definition of several quantities used in the calculation of pairing correlations. (Left)
Idealized set of equally spaced single-particle shell model levels. (Right) Resultant orbit
occupancies V2. The calculation is general since the single-particle energies are given in units of A.
the basic results of pairing theory. We will not derive these standard results but
will try to highlight their key effects.
To proceed, we refer to Fig. 5.6 where several quantities important in
discussing the pairing interaction are defined: the Fermi energy, denoted A;
the single-particle energies, e, with ea being reserved for that level closest to
the Fermi surface; and A, the so-called gap parameter, defined in terms of a
hsum over orbits i, j as
where the usual U and V factors are the so-called emptiness and fullness
factors that pervade the study of heavy nuclei. They are given by:
134 Shell Model and Residual Interactions
These equations can be solved for A, U., and V. for a given set of single-particle
energies e. Or, A can be estimated from empirical mass differences between
adjacent nuclei with odd and even numbers of nucleons. The behavior of V?
against the ratio (e. - A)/A is shown in Fig. 5.6. We see that Vi -> 1 for
(e. - A) « 0 and vanishes for levels far above the Fermi surface. The opposite
applies to Uf. Both fall off rapidly for single-particle energies e. ~ A. Also, from
Eqs. 5.23, V2 + U2 = 1. Uf is the probability that the orbit i is empty, whereas
V.2 is the probability that it is filled. Pairing smooths out the level occupancies
near the Fermi surface over a range ~A, which, in turn, is proportional to G, the
"strength" of the interaction. If there were no pairing, the Fermi surface A
would coincide with the last filled orbit being filled and (e. - A.) would be the
excitation energy required to excite one of the nucleons in this last orbit to one
of the higher orbits er In the presence of pairing, however, this single-particle
excitation energy (e. - A) is replaced by a quasi-partide energy E. given by
Fig. 5.7. Effects of pairing in odd and even nuclei. (Left) Level compression near the ground state
for odd mass nuclei. (Right) The energy gap in even-even nuclei.
as well. It is immediately clear that the minimum energy is Ej = 2A, giving the
famous "pairing gap" that is a nearly universal feature of even-even nuclei in
which few two quasi-particle excitations appear below -1.5 - 2 MeV. In fact,
this empirical energy gap is one way of extracting A experimentally. As noted,
typical values of A range from about 700 keV -1 MeV.
Although there is a large gap between the ground state of an even-even
nucleus and the first excited two-quasi-particle state (2A), above this energy
the quasi-particle level density will be at least as large as in a neighboring odd
mass nucleus.
The energy gap and the high-level density just above it are evident in
virtually any even-even nucleus. Figures 2.5, 2.6, and 2.10 show several
examples. However, there is another very simple and elegant way of demon-
strating the pairing gap empirically without the need for in depth study of each
individual excited state. The («, A) reaction proceeds by the capture of a
neutron of energy En into a target nucleus, forming a residual nucleus at an en-
ergy Ec - S(n) + En. This "capture" state then decays by the emission of yrays
to low-lying levels. Each y-ray thereby defines an excited state by the equation
Exl = Ec - Ey. Under appropriate experimental conditions (See Chapter 10)
the (n, y) reaction provides an a priori guarantee of observing all final states in
certain spin and excitation energy regions. Thus, a spectrum of y-ray transi-
tions will reveal a sequence of peaks, each corresponding to a specific excited
level, and together, displaying all the excited states with certain /* values.
Because of the "completeness," the pairing gap should be immediately identi-
fiable by a sudden increase in peak density. Figure 5.8 shows such a spectrum
for the case of 196Pt, and includes an insert portraying the reaction process
schematically. One sees the ground state, a few low-lying excited states, and
Kg. 5.8. Spectrum of primary frays following average resonance capture (ARC) into 196Pt. The
inset schematically illustrates the reaction process. Two peaks in the spectrum are labeled by their
y-ray energies in keV: the rightmost peak corresponds to a transition to the ground state, while
that at 6100 keV populates a state at 1823 keV The increased level density at the pairing gap is
immediately obvious at approximately this excitation energy (Cizcwski, 1979).
Multiparticle Configurations 137
C O L L E C T I V I T Y , PHASE T R A N S I T I O N S ,
DEFORMATION
This page intentionally left blank
6
COLLECTIVE EXCITATIONS IN E V E N - E V E N
NUCLEI: VIBRATIONAL AND ROTATIONAL
MOTION
the most popular, successful, and widely tested to date has been the interacting
boson approximation (IBA) model of lachello and Arima. This model will
also be discussed later in this chapter. Other algebraic models, which empha-
size the fermions (nucleons) directly, currently offer enticing initiatives that
promise to be both more general and more microscopic.
Since the shell model picture of nucleons orbiting in a spherically symmetric
central field must ultimately be relied upon for the microscopic justification of
collective behavior in nonspherical nuclear shapes, it is worthwhile to see how
this model is capable of producing nonspherical configurations. To many
students this is a mysterious point: the explanation is actually quite simple
while the confusion stems from a semantic misunderstanding. This discussion
will form at least a qualitative justification for the basic concepts of the Nilsson
model—the assumption of a deformed shell model potential—as well as
highlight the central importance of the proton-neutron (p-n) interaction in
the development of collectivity and deformation in nuclei. And finally, it will
help to foster an appreciation of the key role played by the distribution of
valence nucleons between protons and neutrons, as opposed to a simple
consideration of the total number of valence nucleons.
There is really nothing mysterious in the idea of generating deformed
shapes from the spherical shell model. The model is spherical in the sense of
the shape of the central potential, not the resulting shape of the nucleus. A
spherical potential allows no distinction between magnetic substates of a given
orbit, all of which are degenerate. A single-particle in a shell model orbit
occupies all such substates equally, and overall, its wave function will be
spherically symmetric. However, any mechanism that yields an unequal
occupation of m states gives a nonspherical shape. (An orbit may be circular,
but it is a circle in a specific plane and therefore gives the nucleus a "bulge" in
that plane.) Such mechanisms abound in the shell model without introducing
anything fancy (even without residual interactions). Consider, for example, a
configuration of two identical nucleons in equivalent orbits, say I (f7/2)2/),
outside a magic core. The 0+ ground state is spherical, of course, but let us look
at the construction of the 2+ state in terms of m substates. From Table 5.1 we
see that many possible mv m2 values, such as m^ = 7/2, m2 = 1/2 or m^ = 5/2,
m2 = 3/2 give M = ml + m2> 2 and cannot contribute to the 2+ state. This state
is then formed from a nonuniform distribution of mv m^ components and must
be nonspherical. Indeed, this is why it has a quadrupole moment (see Eq. 5.5).
A nonuniform magnetic substate distribution is in fact so characteristic of
deformation that one of the best known features of the Nilsson (deformed
shell) model is a filling of orbits based on their m values instead of their;
values. But that jumps ahead of the discussion. We must discuss how and why
/ itself nay not always be a good quantum number. This is an essential point
since "configuration mixing" of single-particle; values ensures an unequal m
substate distribution and is therefore tantamount to ensuring deformation.
To do this we will show that there is a fundamental difference between the
occupation of valence orbits by like nucleons (e.g., two protons or two neu-
trons) and unlike nucleons (one proton and one neutron). Consider then, and
by way of example, a nucleus with two valence nucleons filling the lower part
144 Collectivity, Phase Transitions, Deformation
of the 8-20 shell in the ld5/2 and ldw orbits, which are separated by =5 MeV.
We want to consider matrix elements that can admix d5/2 and d3/2 compo-
nents. In Chapter 4 we discussed diagonal matrix elements of short-range two-
body residual interactions. Now, we are dealing with nondiagonal ones
(although the diagonal elements still play an important role in modifying the
unperturbed energies of the states that mix). Nevertheless, the basic idea is the
same: if the particles are not close to one another in the two two-particle
configurations, the matrix element will be small. The Pauli principle must also
be considered. In addition, as opposed to the diagonal case, here we also need
to consider the unperturbed (initial) energy spacing of the two configurations.
The whole issue then is just one of two-state mixing. We consider the possible
matrix elements:
we shall see in a later chapter that the widely used and extremely important
RPA and TDA techniques are just such descriptions. Nevertheless, an alter-
nate viewpoint, that approaches the nuclear structure more macroscopically,
emphasizes the nuclear shape and excitations of that shape, providing a much
simpler, physically transparent approach.
In this chapter we shall discuss a sampling of the most important models for
collective excitations in even-even nuclei. As always, the emphasis will be on
the physical ideas.
To begin, we recall some of the systematics shown in Chapter 2. Figure 2.8
showed the energy levels of the Sn, Xe, Te, and Cd nuclei. Sn, with Z = 50, is
singly magic and displays a typical shell model behavior regardless of the
number of valence neutrons. The 2^ energy remains high and the 4^, 6^ levels
cluster. As soon as valence nucleons are added, for example in Te and Cd
(where the two valence protons are counted as holes), ZJ2j- drops sharply. The
decrease grows as the number of valence neutrons increases. The drop is even
faster for Xe, which has four valence protons. Figure 2.15 showed the sys-
tematics of the energy ratio E* {I E^\. It ranges from values < 2 for shell model
nuclei through ~2 for nuclei reasonably close to closed shells, then increases
sharply towards the limiting value of 3.33 near midshell. As we shall discuss,
values near 2.0,2.5, and 3.33 are all typical of different types of macroscopic
collective shapes: spherical harmonic vibrator, axially asymmetric rotor, and
axially symmetric rotor, respectively.
Generally, there is a smooth progression from one to another of these
idealized collective limits. However, inspection of Figs. 2.13 and 2.14 shows
that the systematics is anything but simple. At the end of this chapter we shall
see some easy, physically transparent, ways of understanding this complexity
and of parameterizing the behavior of heavy nuclei. Appropriately enough,
this approach will be based on a recognition of the importance of the residual
p-n interaction among the valence nucleons.
Here, though, we discuss models for each type of behavior, turning later to
their evolution from one into another. We start the discussion with the least
collective nuclei, which occur soon after closed shells: spherical-vibrational
nuclei. The generic concept of vibrational motion in nuclei is widespread and
encompasses a great richness of phenomena. We speak here of a particular
kind. To put this in context, suppose we expand the residual interaction among
the valence nucleons in multipoles, the first few terms will correspond to
monopole, dipole, quadrupole, octupole, and hexadecapole components. Each
of these carries a parity n= (-1)A where A is the multipolarity involved.
The electric dipole mode corresponds, geometrically, to a shift in the center
of mass, and therefore plays little role in the low-lying spectrum of even-even
nuclei. At higher energies, however, it induces the well-known giant dipole
resonance, which can be pictured as an oscillation of the proton distribution
against the neutron distribution. As this mode involves a rather large scale
displacement of major components of the nucleus, it requires considerable
energy, typically between 8 and 20 MeV. Since it is also a negative parity
excitation, and since most of the orbits in a given major shell consist of the
same parity, it necessarily involves excitations of the particles from one major
Collective Excitations in Even-Even Nuclei 147
shell to the next and, once again, we see why it is generally high lying. We shall
not discuss the giant dipole resonance, or indeed, other giant resonances, any
further. This is in no way meant to minimize their importance: indeed, they are
a major focus of current work in nuclear structure physics. Their neglect here
stems rather from the emphasis in this book on the low-lying nuclear structure
spectrum and from the author's feeling that he has nothing particularly new or
innovative to say about the subject.
There is, however, a low-lying magnetic dipole excitation that has recently
been discovered in beautiful electron scattering and yray inelastic scattering
experiments. It occurs, for example, in heavy deformed nuclei at roughly 3
MeV and corresponds to a vibration in which the proton and neutron distribu-
tions osillate with respect to each other with a scissors type of motion, as
opposed to the linear vibrational motion of the giant electric dipole resonance.
The idea is illustrated in Fig. 6.1. This mode, characterized by strong Ml
electromagnetic transitions to the ground state and first 2* state is now known
in a number of nuclei and an interesting systematics has been established. It
has been studied from both geometric and algebraic (IBM-2) viewpoints.
Further discussion of this active area of research is beyond the scope of this
book.
Quadrupole Vibrations
The next vibrational mode, that we shall consider in detail, is the electric
quadrupole or E2 vibrational mode. It appears in different guises in different
categories of nuclei. Near closed shells, where the nuclei are spherical in their
ground state, the action of a quadrupole residual interaction causes the nu-
cleus to oscillate in shape, taking on a range of quadrupole distortions as a
function of time. The Hamiltonian for such a state can be schematically
written as
where Eg is the zero-point energy and the operators b1^ and b2^ create and
destroy this quadrupole vibration: y/^ = b^lO).
For simplicity of notation, and to keep the essential physics to the fore, we
148 Collectivity, Phase Transitions, Deformation
shall henceforth usually drop the subscripts on the operators b-. In the same
spirit, summations over the components /a will usually be implied rather than
explicit. Since we shall make frequent use of phonon or boson creation and
destruction operators, we pause for a moment to recall some key properties of
such operators in the formalism of second quantization. The basic defining
rules for arbitrary creation and destruction operators b and b f are:
and
or
So, b% simply counts the number of b-type bosons. Thus we now recognize
that the second term in //is just the energy, relative to the ground state energy
E0 needed to create the quadrupole phonon excitation, which naturally carries
a spin and parity 2+.
There is no reason, except the limitations provided by the Pauli principle
when the microscopic structure of these vibrations is considered, that prevents
more than one phonon excitation from simultaneously existing. These mul-
tiphonon states y/N h = (b^lO) will correspond to higher and higher nuclear
levels. From Eq. 6.4, the second term in H is the product of the number of
quadrupole phonons and the energy of each. Clearly, at this stage in the
Hamiltonian one has a purely harmonic vibrational spectrum, where the
excitation energy is linear in the number of phonons: for an N A-phonon state,
Ex = Ha (Nph + 5/2), since the quadrupole mode is a 5-dimensional oscillator.
2 2 2 6
2 2 1 5
"1
Z,
2 2 -1 3
2 2 -2 2
2 1 1 4
2 1 0 3
2 1 -1 2
2 1 _2 1
2 0 0 2
-1 •\ 1 1 A 1, £D
/. —1 1
—— n
2 0 -2 0
2 -1 -1 0 "
1 1 1 3
1 1 0 2
1 1 —1
1
1 1 —2 0
1 0 0 1
1 0 -1 0
0 0 0 0 13 o
7 = 6,4,3,2,0
"Only positive total M values are shown; the table is symmetric for M < 0. The full set of allowable m. values
giving M > 0 is obtained by the conditions ml a 0, m} S m2 < m t .
To continue, we must now turn to the question of which spin states are
allowed in multiphonon excitations. For the two-quadrupole phonon case, it is
clear that the maximum possible spin is 4+. But it turns out that only a triplet
of levels with spins J* = 0+, 2+, 4+ is allowed. There are many ways to derive this
result. Perhaps the most elegant is the use of Young tableaux, but here we shall
use the simpler and more straightforward, though more tedious, method of the
m-scheme. The essential difference between the use of the m-scheme for
phonon excitations and for single-particle excitations is the recognition that
phonons, involving particle-hole excitations and integer spins, behave essen-
tially as bosons. Therefore, the Pauli principle is not applicable and the wave
functions must be totally symmetric. This means that all combinations of m
states are allowed. Table 6.1 shows the m-scheme counting of substates for the
case of two quadrupole phonons and shows that the allowed spins are as stated
previously. The m-scheme analysis for the three-phonon case is given in Table
6.2, which shows that this excitation comprises a quintuplet of levels (at three
times the single phonon energy), with spins J* = 0+, 2+, 3+, 4+, 6+. This harmonic
picture of single- and multi-phonon excitations is illustrated in Fig. 6.2.
To pursue the study of multiphonon states, it is necessary to delve more
deeply into their structure. Consider the three-phonon levels. As Fig. 6.3
illustrates, the 6+ state can only be made in one way: by aligning the angular
momentum of a single phonon state with that of the 4+ two-phonon level.
150 Collectivity, Phase Transitions, Deformation
Table 6.3. Relative coefficients of fractional parentage for three-phonon quadrupole vibrator
states*
V2*
^ °* 4+
+
6 V3
4+ vn/7 VlO/7
3+ VT5/7 -V6A7
2+ V7/5 V4/7 V36/35
0+ VI
The normalization is to Vjvf For the three-phonon states, / the squares of the coefficients for each / also
give the relative values of B(E2:/3 —^Ji2 ). For example, the B(E2:4*3. —»2 + 2 .) and B(E2:4+ — > 4 + ) values are
in the ratio 11/10 = 1.1. See text.
states. The wave functions for the 2+, 3+, and 4+ three-phonon states are
therefore linear combinations of two terms, and the relative amplitudes are
simply phonon coefficients of fractional parentage whose squares give the
relative likelihoods that the three-phonon state is made in a certain way. It is
often useful to know these coefficients, so we give them for the N^ - 3 states
in Table 6.3. We will encounter two applications of these coefficients momen-
tarily.
An important aspect of the vibrational model centers on electromagnetic
transition rates since they are particularly sensitive to coherence properties in
nuclear wave functions. We saw, for example, in Fig. 2.16, the systematics of
B(E2:Oj+ -> 2j+) values throughout the periodic chart. Although small and
comparable to single-particle estimates in light nuclei, they attain values
orders of magnitude larger in heavy deformed regions. Intermediate values
characterize the realm of spherical-vibrational nuclei we are presently consid-
ering.
In general, radiation can be given off when any nucleon changes its orbit.
For example, changes in single-particle orbits in shell model nuclei are often
accompanied by the emission of y-radiation. While collective excitations are
clearly not of single-particle nature and the destruction of one does not
correspond to a single change of orbit by an individual nucleon, we will see in
Chapter 9 that their wave functions can be represented as coherent linear
combinations of single-particle-hole (or, equivalently, two quasi-particle)
excitations. Therefore, not only are y-ray transitions between phonon levels
permitted, but the coherence can make them particularly strong. Since a two-
phonon excitation involves a superposition of two linear combinations of one-
body excitations, the destruction of two-phonons would require a simultane-
ous destruction of two particle-hole excitations or four quasi-particles. There-
fore, such transitions are forbidden and one has the characteristic phonon
model selection rule AW h = ±1, where N h is the number of phonons. The
argument for this selection rule (obtained here for quadrupole vibrations of
spherical nuclei) is rather general and applies to any phonon structure de-
scribed as a linear combination of one-particle excitations. We will encounter
it repeatedly in various applications.
152 Collectivity, Phase Transitions, Deformation
Let us now consider the magnitude of these B(E2) values between phonon
states as illustrated in Fig. 6.2 and 6.4, where we assign a value of unity for the
decay of the one-phonon 2+ state to the ground state. Since, in first order,
multiphonon excitations simply consist of the piling on or superposition
of more than one identical phonon, it might seem that the B (E2) value
for the decay of the two-phonon state would also be unity. However, this
neglects the fact that there are two phonons in the initial state and that either
one of them may be destroyed. This gives twice as many decay possibilities and
therefore B(E2:(W = 2) -> 1) = 2, as indicated in Fig. 6.2. Continuing this, one
can state a general expression for the decay of the N ^-phonon state to the
(A^A-l)-phonon state. A transition N^ -> Wph-l must be accomplished by an
E2 operator of the form b, that is, a one-phonon destruction operator. By Eq.
6.2
phonon levels can each decay to two or more of the two-phonon states. We
shall now show that the proportionality of these B (E2) values to the number of
phonons in the initial state actually refers to the sum of the B(E2) values from
a given Nflt-phonon level to all possible (N^-lj-phonon levels. It is trivial to
work out the relative B (E2) values for each of these decay routes: this exercise
is in fact one of the promised applications of the phonon CFP's in Table 6.3.
Consider as an example the decay of the N^ = 3,4+ level to the 2+ and 4+ two-
phonon states.
From Table 6.3, the wave function for the three-phonon 4+ level can be
written, in obvious notation, as
Then, the E2 matrix element connecting this level to the 2+ two-phonon state is
Similarly,
and we see that the three-phonon —> two-phonon B(E2) values are propor-
tional to the squares of the three-phonon CFP coefficients in Table 6.3. The
table can be used to obtain the B(E2) values we have not worked out. The
results for the decay of the 4+ and 2+ three-phonon states are illustrated in Fig.
6.4. We also see an example of our general result, namely,
Fig. 6.5. Energy anharmonicities in the vibrator model assuming arbitrary two-body residual
interactions.
phonon state can only be made by antialigning one phonon with the 2+ two-
phonon state. Again, there are three ways to do this, and the anharmonicity in
the three-phonon 0+ energy will be triple the anharmonicity in the two-phonon
2+ level, or 3e2. The other three-phonon states, which can be made in two or
more ways from the 2-phonon levels, will have total energy anharmonicilies
given by the relative proportions of their wave functions arising from the
various two-phonon states. These relative proportions are given by the CFP
coefficients of Table 6.3, and the resulting energy anharmonicities are shown
in Table 6.4 and illustrated in Fig. 6.5.
It is worthwhile to reiterate what has been derived here. We have never
specified the structure of the phonon itself. We have also never specified the
nature of the residual interaction except to state that it is two-body. Neverthc-
156 Collectivity, Phase Transitions, Deformation
Table 6.4. Energies of the three-phonon quintuplet states in terms of the two-phonon anhar-
monicities.
N. J Energy*
(relative units)
0 0 0
1 2 1
2 0 2 + efl
2 2 2 + e2
2 4 2 + E4
3 0 3+3e 2
3 2 3 + 7/5e0 + 4/7e, + 36/35e4
3 3 3 + 15/7£2 + 6/7e4
3 4 3 + ll/7£2+10/7e4
3 6 3+3£ 0
*£a, £[, and £4 are defined as the deviations of the 0*, 2*, and 4* level energies of the two-phonon triplet from
106 IO8 110 112 111 116 118 120 122 124
CADMIUM SYSTEMATICS
Fig. 6.6. Energy systematics in the Cd isotopes. Note the intermingling of two-phonon triplet
states with extra levels (Aprahamian, 1984).
nuclear regions far from stability. An understanding of them is, in any case,
essential to disentangling the empirical features of many vibrational nuclei.
Moreover, the discussion will introduce some basic ideas relating to the p-n
interaction, which we shall return to later in this chapter. We have already
alluded several times to the idea that this interaction is essential to the
development of collectivity and deformation in nuclei. Here we will encounter
our first specific example of this.
Cd has Z = 48, and therefore two proton holes relative to the Z = 50 magic
number. It is of course possible to excite the Cd nuclei by elevating two
protons from the Z = 28-50 valence shell into the next higher, Z = 50-82 shell.
The idea is sketched in Fig. 6.7. Normally this requires considerable energy
since it involves raising two nucleons across a major shell gap. However, the
residual p-n interaction is strong and attractive, and as such introduces a major
modification to this first-order energy. In a simple picture, one can view the
"normal" states of Cd as consisting of two proton holes interacting with some
number of valence neutrons. In the intruder state, there are four proton holes
in the Z = 28-50 shell plus two proton particles in the Z = 50-82 shell. In a
sense, there are six valence protons that can now interact with the same
number of valence neutrons. In this picture the intruder states in Cd are
analogous to the normal states in Ba, as suggested in Fig. 6.7.
We have already seen that the more valence nucleons there are of both
kinds, the "softer" the structure will be. Sufficient numbers of valence protons
and neutrons lead to deformed shapes. Therefore, in this schematic view,
intruder levels should be more deformed than the "normal" levels. Moreover,
since the attractive interaction is three times greater than in the normal levels,
158 Collectivity, Phase Transitions, Deformation
Fig. 6.7. Schematic illustration of intruder excitations and normal states in Cd. The normal
states of Ba are shown for comparison with the Cd intruders.
the intruder state excitation energies are lowered relative to their unperturbed
(no p-n interaction) value. This lowering increases approximately linearly
with the number of neutrons, and therefore, one expects the energies of these
intruder states to drop from approximately twice the shell gap for 7V-50 or 82
toward midshell, where the p-n interaction strength is maximum. Inspection
of Fig. 6.6 shows that this simple picture is at least qualitatively correct.
As suggested, this intruder state model is quite general and such excitations,
once thought to be rare, are now known to abound throughout the periodic
table. Perhaps the best known example is in the Pb region, whose systemalics
are shown in Fig. 6.8.
Most intruder levels observed to date are proton excitations. This is related
to the role of a strong p-n interaction in lowering these levels. Since there is a
Collective Excitations in Even-Even Nuclei 159
Fig. 6.8. Systematics of 0+ intruder levels in the Pb isotopes (Van Duppen, 1985).
neutron excess in heavy nuclei, the excited valence protons in the intruder
state occupy the same shell as the neutrons, thus enhancing the p-n interaction
(see Fig. 3.5).
The concept of intruder states is far more important than the explanation of
a fewbothersome levels. It is closely connected with the origin of deformation
itself, as we shall discuss towards the end of this chapter.
These comments relate to our discussion of Cd vibrational states because,
as the TV = 82 shell closure in Cd is approached, the expected rise in intruder
energies should leave behind a reasonable vibrational spectrum. This has led
to experiments on 118>120Cd. The data included in Fig 6.6 seem to confirm this
expectation, although studies of absolute B(E2) values show that the interpre-
tation is not quite so simple. The level scheme for 118C is illustrated in Fig 6.9,
which includes the known information on y-ray transition rates and relative
B(E2) values. There is a triplet of levels near 1200 keV in which the energy
separation is much less than Ei\. Furthermore, and most remarkable, an
entire closely spaced quintuplet of candidates for the three-phonon multiplet
was identified. (In fact, candidates for even higher four-phonon states have
been suggested.) Although there are significant deviations from the expected
160 Collectivity, Phase Transitions, Deformation
Fig. 6.9. Level scheme of 118Cd showing the one-, two-, and three-phonon states as well as an
intruder (T level at 1615 keV and possible candidates for four-phonon excitations above 2.2 MeV.
On the right are shown the average (A/V h = 1)/(A/V h = 2) branching ratios. On the left are the
predictions for the three-phonon states assuming the empirically observed anharmonicities in the
two-phonon states. These are the same predictions one would obtain in the U(5) limit of the IB A
(Aprahamian, 1987).
Fig. 6.10. (a) Equal potential surfaces for different multipole distortions, (b) Schematic illustra-
tions of various quadrupole shapes (prolate, oblate, axially asymmetric) as well as of /and /3
vibrational motions.
these two coefficients represent only the motion of the nuclear center of mass,
and write a0 = /fcosyand c^ - a2 = /feiny. The nuclear shape is then specified
in terms of ft andy. j3 represents the extent of quadrupole deformation, while
ygives the degree of axial asymmetry. Most nuclei are axially symmetric, or
close to it, at least in their ground states. For an axially symmetric nucleus, the
potential has a minimum at y= 0°. [It is unfortunate that no single notation for
deformation parameters exists. /? is quite common, but we shall also encounter
e and 8, especially in Chapters 7 and 8. Often, a subscript "2" is appended to
explicitly denote quadrupole deformation.]
The relation between fi, y, and the nuclear radii can be seen by evaluating
the change in radius (Rx z- R0) in Cartesian coordinates as a function of j3 and
To see the shapes implied by these expressions, Table 6.5 gives the values of
these correction terms to a spherical shape for four y values in units
of f5/47r RJ5- Values greater than zero in the table indicate an elongation in
the direction concerned; those less than zero indicate a compression. Note
that, for y values that are a multiple of 60°, two Rvalues are always identical
since the nucleus is axially symmetric for these y values. For y= 0°, the nucleus
is extended in the z-direction and compressed in x and y. This is a prolate
(American football) shape. Oblate (disk-like) nuclei correspond to y= 60° and
180° and are compressed in the y- and z-directions, respectively, and extended
in the xz and xy planes, respectively.
The essential difference between prolate and oblate shapes is that the form-
er is extended in one direction and squeezed in two, while oblate shapes are
extended in two and compressed in one. Intermediate values of y(y* nnlY),
such as the y = 30° example in the table, correspond to axially asymmetric
shapes, that is, to a flattening of the nucleus in one of the two directions
perpendicular to the symmetry axis. Then all three radii are different.
An attempt has been made to depict several nuclear shapes in Fig. 6.10b (as
well as [} and y vibrational motions to be discussed). These pictorial images,
Table 6.5. Changes in the radius of a quadrupole ellipsoid in the x, y,, z directions for several y
values and fixed /}. *
r
0° 30° 60° 180°
SK
< +1 +0.866 +1/2 -1
«;
8R -1/2 0 +1/2 +1/2
-1/2 -0.866 _] +1/2
*AU numbers are in units of VS/4^R0/3.
164 Collectivity, Phase Transitions, Deformation
Fig. 6.11. Empirical systematics of quadrupole deformation parameters /3 in the rare earth region.
while crude and too classical, should be helpful to readers unfamiliar with the
shapes involved.
The systematics of Rvalues (effectively, quadrupole moments) for the rare
earth region is shown in Fig. 6.11. The qualitative behavior is easily under-
stood in terms of a generalization of the seniority argument of Chapter 5 (see
Eq. 5.17). Early in a major shell, when softness to deformation first appears,
the individual; orbits are still nearly empty; hence the quadrupole moments
for the nucleons in these orbits are positive. Then a large positive Q(ft) builds
up rapidly. As the shell fills, however, the contribution of successive /' shells to
the total quadrupole moment decreases, vanishes, and ultimately turns nega-
tive (see Fig. 5.4). On account of these negative contributions, the summation
over the individual quadrupole moments steadily decreases and may even go
negative (as in Pt, not shown in Fig. 6.11) near the end of the shell.
Two important quantities for a quadrupole deformed nucleus are the
moment of inertia and the quadrupole moment of the ellipsoidal shape. Both
can be written in terms of ft for axially symmetric nuclei. For an ellipsoid, the
so-called rigid body moment of inertia is / = 2/5 Mi2. Integrating the radius
over the nuclear surface gives (to first order in ft)
to second order in p.
Note that since Ra <* /I1'3, / = /I5'3 and is a linear function of ft. Qa is
directly proportional to ft in leading order. For the ft values typical of actual
deformed nuclei, ft- 0.3, the higher-order terms are rather small.
Collective Excitations in Even-Even Nuclei 165
Note that for K - 0, only even J values are allowed, so wave function collapses
to a single term
and so on. Thus, the energy ratio E^lEz^ = 3.33. This simple formula is one
of the most famous results of the rotational model and still remains one of the
best signatures for rotational motion and deformation. We have already seen
examples of nuclei that behave according to this relation in Figs. 2.10 and 2.15.
Combining Eqs. 6.8 and 6.13 gives two characteristic features of transitional
and deformed nuclei. For a given mass region (A = const), h2/2I decreases as
j3 increases, leading to smaller and smaller rotational spacings as a deformed
region is entered. This behavior is one of the signatures of nuclear transition
regions, as we pointed out in Chapter 2 (Figs. 2.13 and 2.14). Second, since
nuclear radii increase as Av3, I = A5'3 and W1I = A~s'3, for constant ft.
Rotational spacings should therefore decrease for heavier nuclei. Figure 2.10
illustrated exactly this behavior with examples taken from Zr, the rare earth,
and the actinide regions.
As we stated earlier, in a geometrical picture (that is, a macroscopic one in
which we do not worry explicitly about the Pauli principle), there is no reason
why rotational motion cannot be superposed on intrinsic excitations, whether
of collective vibrational or two-quasi-particle character. Now consider such an
intrinsic excitation in a deformed nucleus.
Each intrinsic excitation carries intrinsic angular momentum Jo. It can be
partially characterized by the projection of that angular momentum onto the
symmetry (z) axis. Since for axially symmetric nuclei, any rotation of the
Collective Excitations in Even-Even Nuclei 167
The energy expressions, Eqs. 6.13 and 6.14, are quite accurate for low spin
states in deformed nuclei, thus affirming the basic validity of the rotational
concept. An example is shown in Fig. 6.12 for a typical deformed nucleus, 164Er
(the nature of the intrinsic excitations indicated in the figure will be discussed
momentarily). The energies for each band are normalized to the bandheads in
order to isolate the rotational behavior. The predictions from Eq. 6.14 are
reasonable, but there are also clear deviations as J increases. Also, note the
changes in the inertial parameter, h 2/27, from band to band. Apparently the
deformation is not completely constant.
An understanding of the physics involved in these deviations is extremely
Fig. 6.12. Ground, /, and J3 band levels of a typical deformed nucleus 164Er. For each band the
symmetric top rotational energy predictions (Eq. 6.14) are shown.
168 Collectivity, Phase Transitions, Deformation
transitions, while collective, are much weaker than y-» g transitions, typically
a few single-particle units. We illustrated these points in Chapter 2 (Fig. 2.18)
by summarizing the transition rate data for deformed nuclei in terms of ratios
of y-> g and /? -> g B(E2) values to B(E2:2* -> Oj+) values. One remarkable
feature is the relative constancy of the y-» g B(E2) values. This surely points
to a collective, slowly evolving structure.
To further consider y-ray transitions, both within rotational bands (intra-
band transitions) and between intrinsic excitations (interband transitions), we
must be a bit more formal. The basic results are extremely simple to derive.
Taking the wave function of Eq. 6.10 for deformed nuclei, the E2 transition
matrix element (up to constant factors) is:
where we have neglected the cross terms of the form <JOE2II-.K.), which
normally vanish by conservation of angular momentum, and have utilized the
fact that the Kf -> K. and -Kf -» -K. matrix elements are identical. The
separation of the wave function into rotational and vibrational components
thus gives a separation of the matrix element into an intrinsic part (second
factor on the right in Eq. 6.15) dependent only on % and a "rotational" part,
which in turn depends only on the angular momenta involved and is propor-
tional to the Clebsch-Gordon coefficient (JK2 AK\JKf}.
The diagonal matrix elements with j. = %f, J. = /.give the intrinsic quad-
rupole moments Qo of the excitation involved. Note that these quadrupole
moments are given in the intrinsic body-fixed frame. The observed quad-
rupole moments, that is, the so-called spectroscopic quadrupole moments,
involve a transformation to the laboratory frame, giving the well-known result
The dependence on K and / reflects the fact that the perceived shape of a
rotating nucleus is not the same as the shape in the intrinsic frame. This is easy
to visualize. When a prolate deformed nucleus rotates about an axis perpen-
dicular to the symmetry axis, the time averaged shape looks more like a disk
(an oblate nucleus), which would have a quadrupole moment of the opposite
sign. This effect is exacerbated for higher rotational velocities and, indeed, for
J(J +1) > 3K2, the spectroscopic quadrupole moment does have a sign opposite
to the intrinsic quadrupole moment. In fact, for K = 0 this is always the case.
Note that for J = 0 (which implies K = 0 since K < J), Q = 0: a state of zero
angular momentum can have no preferred direction of the time averaged
distribution in space and therefore no quadrupole moment.
For the important case of matrix elements diagonal in % but not in J (or
transitions within a band), we have nearly the same result except for a Clebsch-
Gordon coefficient connecting 7. and 7- Thus
172 Collectivity, Phase Transitions, Deformation
Since the intrinsic quadrupole moment Qo<* P(l + 0.16/3), B(E2:0^ -^ 2/) ~ I?.
The large Rvalues (of about 0.3) that characterize deformed nuclei can lead to
a one to two order of magnitude increases in this B(E2) value above that of
near spherical nuclei (j3 ~ 0.05). This explains the systematics we showed in
Fig. 2.16, which provides the most obvious evidence of deformed collective
behavior in nuclei.
Frequently one can extract very sensitive and critical information on struc-
ture effects and rotation-vibration interactions from B(E2) values for a pair of
transitions connecting the same two intrinsic states. These transitions can be
either both intraband or both interband with the same initial and final bands.
Then the intrinsic matrix element will clearly be identical for both transitions
and will cancel in their ratio. Such branching ratios depend only on the squares
of Clebsch-Gordon coefficients, and are therefore model independent in the
sense that they do not depend on the microscopic structure of the excitations
involved. They depend on the assumption of the separability of rotational and
vibrational motions. They are known as Alaga rules.
Specifically, we have
Note that, since the intrinsic structure has canceled out in such ratios, they
are equally valid for transitions involving any intrinsic states (e.g., two-quasi-
particle states) as well as for those involving vibrational excitations.
As examples of these ratios, we have
Equally simple but numerically different results are obtained for other
transitions. There are two important uses of such ratios. First, since they
depend on the K values of the initial and final states, they can sometimes be
empirically used to assign K quantum numbers to different intrinsic excita-
tions. Secondly, as we shall see momentarily, small admixtures of different
intrinsic excitations (bandmixing effects) can induce enormous changes in
these branching ratios, so the empirical ratios can provide very sensitive tests
of small details of the nuclear wave functions.
It is easy to look up or calculate values of the Clebsch-Gordon coefficients
Collective Excitations in Even-Even Nuclei '• 173
Table 6.6. Some useful Alaga rules for E2 transitions in deformed nuclei*
(J.K2AK \fffi
/. J
f K^Kf
0->0 2->0 0-»2 2->2
0 2 1.0 1.0 —
—
2 0 0.200 0.200 — —
2 0.286 0.286 0.286 0.286
3 — — 0.500 0.500
4 0.515 0.014 0.215 0.215
3 2 — 0.358
— 0.358
3 — — — 0
4 — 0.143 — 0.343
5 — — — 0.300
4 2 0.286 0.120 0.008 0.120
3 — — 0.112 0.267
4 0.260 0.351 0.351 0.042
5 — — 0.389 0.234
6 0.455 0.031 0.142 0.340
5 3 — 0.191
— —
4 — 0.319 — 0.191
5 —. 0.093
— —
—
6 — 0.182 0.167
7 — — 0.360
—
6 4 0.315 0.098 0.021 0.235
5 — — 0.154 0.141
6 0.255 0.364 0.364 0.130
7 — — 0.347 0.124
8 0.431 0.039 0.116 0.371
*The entries are the squares of the Clebsch-Gordon coefficients for each indicated transition. Thus, relative
B(E2) values connecting states /., Jf in bands with K., Kf are B(E2 JX.-tJfa) ~ (J.K2&K l^)2.
Table 6.7.Comparison of some relative B(E2: y—»g) values in deformed rare earth nuclei with the
Alaga rules*
Relative B(E2:JJf)
U4 168 178
J. Jf Alaga Gd Er Hf
2 0 70 43 54 88
2 100 100 100 100
4 5 14 6.8 5.8
3 2 100 100 100 100
4 40 105 65 52
4 2 34 16 20 18
4 100 100 100 100
6 9 14 —
—
—
5 4 100 100 100
6 57 — 123 107
6 4 27 — 12 18
6 100 — 100 100
8 11 — 37 —
"One transition is normalized to 100 for each initial state. The Alaga rule entries are relative values from Table
6.6 for the K = 1 -> K = 0 case.
tions from them and their study greatly deepens our understanding of
deformed nuclei.
2. The deviations increase substantially with increasing spin.
3. Transitions in which the spin increases (Je > /) are nearly always empiri-
cally larger than the Alaga rules, while spin decreasing transitions (Jf<J)
are nearly always smaller.
4. The deviations can become quite large, leading to factors of three or four
discrepancies from the predictions.
5. The deviations are, on average, larger in Sm than in Er and larger in Er
than in Hf.
Combined with all the evidence from rotational energy sequences, meas-
urements of quadrupole moments, and the like, point 1 provides a vast body of
evidence that supports the idea of a superposition of rotational and intrinsic
motion and the approximate separability of the two. Point 5 suggests that this
separability is most applicable in midshell and least just after the transition
regions from spherical to deformed nuclei. This is reasonable, of course, since
the energy scale of rotational motion decreases systematically toward mid-
shell, and therefore, the distinction in energy between rotational and vibra-
tional behavior is larger there than closer to the vibrational regions at the
beginning and end of major shells. We shall soon see more dramatic evidence
of this point shortly in terms of a systematic measure of the rotation-vibration
coupling. First, however, in order to understand points 2, 3, and 4, and in
particular why point 4 does not indicate a serious breakdown of the rotational
description, we must introduce the concept of bandmixing and discuss a
quantitative formalism to treat it in a simple way.
Collective Excitations in Even-Even Nuclei 175
where e' is the small mixing amplitude of each band in the other. It is
convenient to separate the spin dependent and spin independent parts of the
mixing by writing
where K is the lower AT value. We now interpret y-> g mixing as proceeding via
a two-step Coriolis effect through an intermediate K = 1 band (which need not,
and generally is not, known empirically). For weak mixing, this can be viewed
as a sequence of two separate two-step mixing effects.
We know from the discussion in Chapter 1 of weak two-state mixing that if
three states 0,, 02, 0, mutually mix, the mixing of states 02 and 03 gives
Then, if the already mixed state 02' mixes with state 0P we have
Thus, the overall mixing amplitude of state 03 in state ^ is simply given by the
product of the individual two-state mixing amplitudes J312 and /323.
176 Collectivity, Phase Transitions, Deformation
Applying this to the present case, we have the mixing sequence (K = 0) ->
(K = 1) -»(K = 2). Hence the spin dependence of the AAT = 2 mixing amplitude,
/XO is
Similarly, for mixing between the j8 (K = 0) and ground bands, the mixing
sequence is (K = 0) -> (K = 1) -»(K = 0). Then/),(./) is given by
Note that bothf^f) and/r(/) -» /* for large J: the band mixing increases rapidly
for high spin.
The spin dependence of e' explains point 2 concerning the increase of the
deviations from the Alaga rules with increasing spin. However, we have yet to
explain why these deviations can be so large without implying a corresponding
destruction of the entire rotational picture on which the Alaga rules and the
present formalism are based.
We can now calculate the interband E2 matrix elements very simply using
the admixed wave functions of Eq. 6.22.
The first term in Eq. 6.27 is the direct matrix element in the absence of
mixing. Thus, the perturbed matrix element can be written as a sum of a direct
term plus a contribution proportional to e'. In deriving this expression we
have dropped terms in e' 2 since the mixing is assumed to be small. Each of the
two terms multiplying er" is proportional to a Clebsch-Gordon coefficient
multiplied by the intrinsic quadrupole moment of the 7 or ground band.
Therefore, even if we assume these intrinsic moments to be equal (as is
commonly done since the deformation does not differ much from band to
band), the K dependence of the Clebsch-Gordon coefficients prevents this
term from vanishing.
In the case of /} -> g mixing, exactly the same formalism applies with a
substitution of ff for/ One interesting result for the special case of transitions
that do not change spin (/. = / ) follows immediately. For identical quadrupole
moments, the two terms multiplying e' are identical and vanish: The /? —» g
bandmixing has no effect on transitions which do not change the spin. This is
a special case of the result derived in Eq. 1.17.
Incorporating the spin dependence of € and expressions for the Clebsch-
Gordon coefficients in Eq 6.27 leads to a general form for the effect of band-
mixing on interband B(E2) values. We obtain in this way the well-known ex-
pressions:
Collective Excitations in Even-Even Nuclei 177
Table 6.8. Correction factors F (7., J.) and F^ (7., 7) for y-> g and ft -> g reduced E2 matrix
elements due to y- g and /3 - g bandmixing
7. ] Correction factor
—— j-—
7 1 + (27 + 1 ) 2 1 + 2 ( 2 7 - 1 ) 2 ,
/-I / ! + (/,+2)2 -
/ I, l+2Z r 1
X+i f, i-(J}-i)zr -
7,+ 2 7) l-(27j+l)2 y 1-2(27,+3)2,
*Riedinger, 1969.
^~ 10.0(14) 20 6.7(18)
2 2
r-* g
— 2.38(18) 1.43 8.8(14)
lyr-tOg
~——
z >4
r~ « 4.16(61) 14.0 7.6(1.1)
A plot of the left side against the spin function on the right is a straight line with
intercept Ml at J = / and slope M2. From such a plot, called a Mikhailov plot,
one can extract directly from the empirical results both the direct intrinsic
unperturbed AK - 2 matrix element and the mixing amplitude e (<^M2), pro-
vided the data can be fit by a straight line. Deviations from a straight line can
arise from several sources: unequal quadrupole moments of the bands, more
than two-bandmixing, undetected Ml components in the interband transi-
tions, or two-bandmixing that follows a different spin dependence than that
given by the fp(J) and/(7) functions.
An example of a Mikhailov plot, for 168Er, is shown in Fig. 6.14 (the dashed
line labeled IB A will be discussed later). It is clear that the data points are very
well approximated by a straight line, thus validating the use of the Mikhailov
formalism. Such data are typical of deformed nuclei.
The use of the bandmixing formalism in either the Zror Mikhailov forms is
a powerful tool for analyzing deviations of relative E2 transitions from the
rotational (Alaga) values and for studying rotation-vibration interactions.
Empirically, it is invariably found that M2« Ml and negative (by convention,
Ml is positive): therefore Z is positive. In this way, Z values have been
extracted for a number of rare earth nuclei. The results are summarized in Fig.
6.15.
These systematics exhibit a parabolic behavior that minimizes at midshell.
This smooth pattern highlights, indirectly, the collective structure of the unper-
turbed states involved: if the mixing were with single-particle excitations, it
would surely be more erratic. As we shall see in Chapter 9, however, the
microscopic structure of collective vibrations changes smoothly and slowly
with N and Z. The minimum in Z at midshell, and the generally small values
Collective Excitations in Even-Even Nuclei 179
Fig. 6.14. Mikhailov plot for y-> g transitions in 168Er (Warner, 1982).
Fig. 6.15. Empirical systematics of Z^in the rare earth region (Casten, 1983). N is half the number
of valence nucleons.
180 Collectivity, Phase Transitions, Deformation
where g(J) and h(J) are positive functions of the final spin /.
It is worth working through an explicit example to see how the bandmixing
technique is used. We take the case of y-ground mixing in 168Er and use both
the analytic approach with Table 6.8 and the Mikhailov formalism.
According to Tables 6.6 and 6.8,
where the first factor is the unperturbed (Alaga) ratio. From the experimental
value of 1.85, we obtain Z = 0.044. Similar values are obtained from other
transitions. A good average value is Z ~ 0.038. The small magnitude of Zr
confirms the adequacy of the two-state bandmixing approach.
Turning to the Mikhailov approach, which generally is easier and yields
interesting physics more directly, Fig. 6.14 gives
Using Qg = 7.61 eb, the spin independent part of the mixing amplitude is
Collective Excitations in Even-Even Nuclei 181
The full mixing amplitudes e'(J) = VI e f(J) are then -0.0053 (2+),
-0.012 (3+), and -0.021 (4+).
We can now calculate the actual mixing matrix element since for such small
mixing we have, from Eq. 1.10
If we now assume that the mixing interaction, though possibly large, is not
very spin dependent, then the composition of the mixed wave functions will
also not depend much on spin and therefore, a~ a 'and fi~ fi'. But, then, the
factor (aa' + flfi') ~ 1 by orthonormality and the intraband transition has an
identical B(E2) value as in the unmixed case.
Extending this argument to multiband mixing, the factor (aa' + /J/T) in Eq.
6.37 will simply be replaced by (aa' + ftp' + 77' + 88' +...). If, again, the
primed and unprimed mixing amplitudes are approximately equal, this is just
the orthogonality sum, which is unity. (For interband transitions, the ampli-
tude sum is (0,0,' + /3j/32' + ...), and in this case nearly vanishes by the same
orthogonality argument.) Thus we see that, although the mixing is large,
z'm/'aband transitions are barely affected and retain their normal rotational
strengths.
This has many repercussions, two of which are worth citing briefly. It
Collective Excitations in Even-Even Nuclei 183
means, for example, as we argued already, that Alaga rules for intraband
transitions are essentially unaffected by mixing. Thus, observed deviations
from the Alaga rules can be ascribed to other mechanisms (e.g., Ml compo-
nents) and can be used to estimate these. Second, consider heavy ion reactions
that bring large amounts of angular momentum into the nucleus, which then
decays by a series of cascade transitions (see Chapter 10). It has been observed
that these cascades flow through many rotational bands, but that the popula-
tion within a band tends to remain intact as J decreases, even though these
relatively high-lying quasi-particle excitations are expected to mix considera-
bly. The preceding derivation provides a simple explanation: the mixing can
indeed be strong, but as long as it does not change rapidly with J, the z'mraband
transitions are only slightly affected and remain dominant.
In closing this section, we note that extensions of the formalism to include
j9 - 7 mixing have also been developed and are available in the literature. One
point that will be useful in our later discussion of the IBA can be deduced
immediately without a formal development of the mixing expressions. The
effects of/?- 7 bandmixingon ft —>gand 7—> g transitions are second order
and generally weak; however, since j3 -> 7 transitions are forbidden in the
absence of mixing, such mixing can strongly break this fundamental selection
rule. The expression for /} —> 7 transitions in the presence of j3 - 7 mixing is
analogous to those we generated for the ft — » g and 7—> g cases, except that
there is no longer a direct term and hence the entire transition strength arises
solely from a mixing term proportional to Z . Thus, in ratios of /3 —» 7
transitions, Z. cancels out. Therefore, although the finite transition matrix
elements arise from mixing, branching ratios are independent of the strength of
that mixing and are given only by ratios of functions of 7, and 7 . We note for
future reference that in the IBA model, ft -> 7 transitions are, in contrast,
allowed for deformed nuclei but their branching ratios depend on the detailed
structure (in effect on the value of the asymmetry parameter 7). We will
discuss this further later in this chapter.
Having discussed the low-lying, intrinsic excitations of axially symmetric
nuclei, we can return to the question of rotational energies and corrections to
the simple first-order expressions in Eqs. 6.13 and 6.14. It was useful to discuss
these intrinsic excitations, in particular, the bandmixing between them, first,
because the corrections to the symmetric top formula are intimately connected
with excursions from axial symmetry and rotation-vibration coupling. Indeed,
the first order rotational expression makes several implicit assumptions, the
most important of which are that there is no coupling between rotational and
intrinsic degrees of freedom and that /? is independent of J. These two
assumptions are, in fact, related. As the nucleus rotates, it experiences a
centrifugal force that tends to increase the deformation and moment of inertia
and decrease the rotational spacings, and leads to an enhanced coupling to
vibrational modes (recall Eqs. 6.25 and 6.26, which show that this increases
with spin). There are several ways of incorporating these effects into a
rotational energy expression. One of the first and most common is simply to
expand the rotational energy in powers of 7(7 + 1) and keep the second term.
184 Collectivity, Phase Transitions, Deformation
Fig. 6.17. Empirical ground band levels of Yb compared with various models. The labels ab and
A, AB, ABC refer to the coefficients in Eq. 6.40 and in the expansion of rotational energies in
powers of J(J +1) (see Eq. 6.40 and following discussion).
where A - h 2/2I. (We will derive this formula in a moment.) From our earlier
comments, we know that empirical values of B are negative. If they are also
small (BIA « 1), the expansion converges rapidly and Eq. 6.38 will be a
significant improvement.
In some cases, still higher-order terms such as CJ3(J + I)3 are necessary to
produce adequate fits for higher J values. Rather than explore this, we shall
turn shortly to an alternate expression that automatically includes Eq. 6.38 and
all higher-order terms. First, we show an example in Fig. 6.17 of the ground
state rotational band of 168Yb compared with the energies calculated from Eqs.
6.13 and 6.38, as well as other expressions to be discussed. Evidently, the first-
order expression (Eq. 6.13) is reasonable only for very low-spin states. Equa-
tion 6.38 (AB in the figure) is an improvement for higher spins, although it too
encounters serious difficulties for still larger./. A fit with the CJ3(7 + I) 3 term
(ABC) included further improves the predictions, but is also inadequate for
large 7: the opposite signs empirically deduced for B and C tend to produce
wild oscillations in predicted energies (compressions of levels, even spin
inversions) at high enough./ values.
An alternate approach to incorporating rotation-vibration or centrifugal
Collective Excitations in Even-Even Nuclei 185
effects into the rotational energy expression is to make the moment of inertia
spin dependent. This approach is known as the Variable Moment of Inertia
(VMI) model and has enjoyed considerable success. In general, its predictions
are better than those of Eq. 6.38, and it is not limited to the realm of strongly
deformed nuclei. Figure 6.17 includes VMI predictions and shows their
advantages. We shall not dwell on this approach, as it has been extensively
covered in other literature.
Interestingly, it is easy to see how both effects (a change in the moment of
inertia and the addition of a higher order term) result immediately from the
effects of f-g bandmixing.
We have seen that the mixing is generally small so we can use the approxi-
mation of Eqs. 1.12 to write the energy shift (lowering) of the ground state
band as A£ *sb(J) = V2/AE2 . But, from Eq. 1.10, the mixing amplitude
er' =V2e/r(y) = VIAE^. So, AEf = 2 eff(J). Hence, from Eq. 6.25,
The second term is the promised correction to the standard rotational formula,
and can give the variation with / of the inertial parameter ft 2/27, while the first
gives the required second-order correction term. From this derivation it is
clear that Eq. 6.38 is, as we implied earlier, ultimately connected with the
concept of rotation-vibration coupling (bandmixing) and also that it implicitly
assumes small mixing.
When / becomes large enough such that er/r(J) ~ 1 we must anticipate a
breakdown of Eq. 6.38 and thus a need for many higher-order terms or an
alternate formula. We have seen this effect empirically in the failure of Eq.
6.38 for J> 14 in Fig. 6.17.
However, there is a much superior rotational expression that is valid for
even higher spins that unfortunately has not been discussed much in the
literature. It automatically gives Eqs. 6.13 and 6.38 as limiting cases, automati-
cally includes all the higher order correction terms, and moreover, contains a
specific relationship between the coefficients of each successive term. One
simply writes the two-parameter formula
or
Hence
which is simply Eq. 6.40 with a = a/2/3 and b = 2h 2fi/a2. Note that energy ratios,
such as Ejl E2\, depend only on the single parameter b in Eq. 6.40. Neverthe-
less, this formula is far more accurate than any of the expressions we have con-
sidered, as shown in Fig. 6.17 for 168Yb where the predictions are compared
with one-, two-, and three-term expansions in / (J + 1) and with the VMI
model. Its success extends to softer (transitional) nuclei (e.g., 152Sm, 184Pt).
Since the expression works so well for higher/, we anticipate a later discussion
to caution that it is only applicable below any "backbend" that may be present.
Aside from its empirical success, Eq. 6.40 is interesting because, for rela-
tively low spins such that bJ(J + !)«!, expansion of the square root naturally
recovers the second (and higher) order terms in the rotational formula of Eq.
6.38:
Naturally, this would lead to more radical departures from the energy and
transition rate expressions we have considered. In fact, in certain limiting
cases of large asymmetry, new selection rules appear. In another sense,
however, such models for larger asymmetry are extensions of the small excur-
sions from axiality dealt with so far, and their predictions go over into the latter
as y-» 0°. It also turns out that many predictions of models for large, fixed
asymmetries y are identical, or nearly so, to models incorporating dynamic
fluctuations in y so long as %. d in one equals yms in the other.
The best known model of fixed stable asymmetry (triaxiality) is that of
Davydov and co-workers developed around 1960. Here, the potential V (y) is
envisioned to have a steep, deep minimum at a particular value of y so that the
nucleus takes on a rigid shape with that asymmetry.
We have seen that, if the rotational and vibrational motions are not com-
pletely decoupled, and there is an interaction (mixing) between the y and
ground bands, the latter will acquire a finite yrms and K will no longer be a good
quantum number. Therefore, it is not surprising that in the Davydov model K
is not a good quantum number either. Here, however, since y can be large, the
K admixtures can reach levels far beyond those we have encountered.
The relation between the Davydov model and models with axially symmet-
ric but ysoft potentials runs deeper than this. In a nucleus with such a
potential, the greater the softness the lower the y vibration will lie, and the
larger ymt will be in the ground state. In the Davydov model there is no
distinction in intrinsic structure between what is normally called the ground
state rotational band and the y vibrational states. The levels of these two
bands simply become the so-called normal and anomalous levels of a new
ground state band whose energies depend explicitly on y, which can take on
values fromO 0 —> 30° (prolate symmetric —»maximum asymmetry: 30° < y< 60°
corresponds to the "oblate" region of asymmetry). Figure 6.18 shows the
lowest levels as a function of y and clearly illustrates the descent of the y
vibrational levels. Indeed, for y> 25°, E2+<E4+. In contrast, as y-» 0°, the
normal levels 0+t, 2+v 4+v 6+p which are rather insensitive to y, go over into
those of an axially symmetric ground state band, while the "y band" energies
increase rapidly. An important feature of the anomalous levels is their energy
"staggering": They tend to be grouped into couplets as (2+, 3+), (4+, 5+)...
The behavior in Fig. 6.18 is easy to understand. As y increases, the nucleus
becomes increasingly flattened. Therefore, states whose wave functions are
predominantly aligned in the direction of the flattening attain lower energies,
since the nuclear force is attractive and they are, on average, closer to the bulk
of the nuclear matter. This is exactly the case for the K - 2 (and higher K)
levels which, therefore, rapidly decrease in energy with increasing y. They also
mix with the normal ground state band levels (yrast states) and, as we have
seen, K is no longer a good quantum number. The leveling off of the energies
of the anomalous levels for y >25° is easily understood in terms of that mixing.
In Table 6.10 we give a number of interesting quantities relating to the
Davydov model, including the amplitude for K = 0 in the 4^ and 42+ states as a
function of y. For y between 25° and 30°, the major amplitudes in the 4 + state
188 Collectivity, Phase Transitions, Deformation
Fig. 6.18. Normal and anomalous levels of the triaxial rotor (Preston, 1975).
Table 6.10. Some useful predictions of the asymmetric rotor (Davydov) model*
where
The individual energies are equal to the respective numerators and denomina-
tors multiplied by (9/sin23y). Note that X-> 1 for y-»0°. Thus, £ 2 £/£ 2 ;->°°
for y-> 0° as seen in the Table and figure. Also, X -> 1/3 for y-> 30°. Table 6.11
gives the empirical values of E^ I Ei\ for a number of heavy rare earth nuclei
as well as the associated y values. If, on account of the anomalously low
E^*^IE^\ ratios, these nuclei are considered to have large rigid triaxiality (this
is the term usually used for the concept of fixed asymmetries), then these y
values are the only input needed to make Davydov model predictions for other
observables. Such predictions are included in Table 6.10 for several y values.
The branching ratio,
is also a useful indicator, as shown in the same table and in Fig. 6.19. R2 can be
written analytically and calculated for any y from
190 Collectivity, Phase Transitions, Deformation
where the numerator and the denominator are the individual B(E2) values.
Note that both B(E2) values in R2 vanish for 7= 0°, yet they have a finite ratio
that is the Alaga rule: the vanishing is reasonable since for 7= 0°, E2$ I Ei \ -» °° ,
corresponding to infinite rigidity in the 7direction and to vanishing vibrational
amplitude. R2 increases rapidly with 7 and R^ —> °° for 7= 30°. This latter result
is identical to the selection rule for an alternate model of axial asymmetry that
we will soon discuss, the 7 flat or y-unstable model of Wilets and Jean in which
the 2 transition is forbidden.
Some other B(E2) values and branching ratios are given in Table 6.10.
Those corresponding to rotational transitions in either the 7 or ground bands,
are nearly 7 independent. Others vanish at both 7= 0° and 30° but attain small,
Collective Excitations in Even-Even Nuclei 191
Table 6.11. Values of E2+/E2+ for some deformed and transitional nuclei and the corresponding
y values (rounded to nearest degree)
Nucleus E 7
^'EA
152
Sm 8.9 13°
160
Dy 11.1 12°
168
Er 10.3 13°
172
Yb 18.6 9°
176
Hf 15.2 10°
182-^y 12.2 12°
184
0s 7.9 14°
188
0s 4.1 19°
192
0s 2.4 25°
196p, 30°
1.94
finite values for intermediate 7 values. These are transitions between normal
and anomalous levels. The 22+ —> 2,+ and 3,+ —> 4^ transitions form a third
category: small at 7 = 0° and rising rapidly toward 7= 30°. This behavior is
easily understandable if we note that, in the 7= 30° case, the Davydov model
has the same selection rules as both the quadrupole vibrator model and the
Wilets-Jean 7-unstable model. For example, at 7 = 0°, the 22+ -> 2^ and
3a+ —> 4j+ transitions are interband (7—> g): given the built-in stiffness in 7, they
must be forbidden. At 7= 30°, the 3:+ state is analogous to a three-phonon
level, the 4^ level to a two-phonon excitation, so the transition becomes
allowed. Similarly the 22" —» 2 T + transition is analogous to a 2 —> 1-phonon
transition.
Since finite 7 values correspond to mixed K values, one might expect a close
relation between B(E2) values for finite 7 in the Davydov model and in the
bandmixing formalism. This is indeed so. Consider, as an example, the case we
worked out earlier of 168Er. We found that the Mikhailov plot analysis gave a
full mixing amplitude in the 4+ states of e '(4*) - 0.021. In the Davydov model,
the experimental ratio R^ = 1.85 yields 7- 11° (see Table 6.10). This, in turn,
corresponds to a y-» g mixing amplitude of 0.03, which is quite close to the
bandmixing result. The agreement is not exact since the comparison is not
quite on an equal footing. In the bandmixing case, the spin independent
mixing amplitude was deduced from a Mikhailov plot, which gives an overall
average value for all transitions, while the Davydov mixing value was deduced
from R2 alone. In any case the essential point is that both the bandmixing
formalism and the Davydov model lead to K mixtures in 7 and ground bands,
and give comparable mixing amplitudes and B(E2) values for small 7. (The
bandmixing formalism is a first-order perturbation treatment and is therefore
inapplicable for large 7.) Though the physical pictures are different, predic-
tions for many observables are nearly identical.
When we turn to the large 7extreme, one might think that the extremely
low-lying 7 band levels of the Davydov model would be an appropriate
signature for stable axially asymmetric shapes. However, we can also envision
large 7 values as dynamic quantities by picturing a deformed nucleus that is
192 Collectivity, Phase Transitions, Deformation
where -d is a constant analogous to hz/2l and the levels are now classified
according to the quantum number A. This classification scheme is given in Fig.
6.20. The yrast levels have / = 2A. Note that each A value (for A > 1)
corresponds to more than one level and that the A values 2, 3, 4, ... include a
low-lying set of levels analogous to the 7 vibrational band and to the anoma-
lous levels of the Davydov model for large 7.
Now that we have the three basic extreme geometric models, the harmonic
vibrator, the axially symmetric deformed rotor, and the 7soft axially asymmet-
ric deformed rotor, it is interesting to compare their rotational energies. The
results are shown in Fig. 6.21. The spacing of the normal rotational levels of
the Wilets-Jean rotor are quite different than in the symmetric rotor. This is
easy to see if one notes that for these levels (the yrast states) A = J/2, so that
Ew^(/-yrast) = (-Q/4)J(J + 6), which increases with / considerably slower than
the /(/+ 1) law for a symmetric rotor. For example, Et\lEi\ = 2.5 compared
to 3.33 for the rotor and E6IE = 4.5 instead of 7. We note for later reference
Fig. 6.20. Ground band levels in the y-unstable or Wilets-Jean model. (Left) In terms of A
multiplets; (Right) Displayed in analogy to the quasi-band structures of a normal rotor.
Collective Excitations in Even-Even Nuclei 193
Fig. 6.21. Dependence of ground band energies on spin for different models. An identical set of
curves is obtained in the U(5), O(6), and SU(3) symmetries of the IB A (see Eq. 6.72).
that the 7 unstable limit of Wilets-Jean is very closely related to the 0(6) limit
of the IB A Model.
Another characteristic of the Wilets-Jean model concerns E2 selection
rules. The allowed transitions must satisfy AA = ±1. Comparison with Table
6.10 shows that the Davydov model goes over to the same selection rules as the
Wilets-Jean model for 7= 30°, again highlighting the similarity of a y-flat model
with ymt = 30° to a rigid asymmetric model with fixed j- 30°. The Wilets-Jean
picture also resembles the phonon model in its E2 selection rule and gives
some identical predictions. For example, the 22+ -> 2/ transition
(A = 2 -^ A = 1) is allowed, as is 4^ -> 2+, but the crossover 2^ -»0^ transition
is forbidden (AA = 2), as is the 4t+ —> 22+(AA = 0). These are the same results
one would have if the O/, 2* and (4^, 22+) doublet were treated as the zero-,
one-, and two-phonon vibrational states. This is not surprising since the
potential for the spherical vibrator, while parabolic in ft, is independent of y, so
it is trivially 7-unstable as well. The real difference between the vibrator and
Wilets-Jean limits is that in one /3ave = 0, whereas the other has a deformed
minimum. Another difference between the two models (with the vibrator now
considered in its harmonic, degenerate multiple! limit) is that the vibrator has
a two-phonon triplet of levels 0+, 2% 4+ while the Wilets-Jean scheme has only
a 2+, 4+, doublet, the first excited 0+ state having A = 3. Thus, not only is it
higher in energy but its allowed decay is to the second2t level (2+2) rather than
the first as in the vibrator; more precisely, B(E2:02+ -> 220 /B(E2:02+ -» 2t+) -»°o
in the 7-unstable limit but is zero in the vibrator. The reader is warned,
however, that these differences do not necessarily persist in a sufficiently
anharmonic vibrator. We will encounter this point, and this close relationship
between vibrator and 7-unstable models, again, in our discussion of the U(5)
and O(6) symmetries of the IB A model later in this chapter.
Despite the obvious similarities between the Davydov and Wilets-Jean
models, there is one outstanding difference by which to distinguish them
194 Collectivity, Phase Transitions, Deformation
For example,
These double energy differences involve three levels with spins J, J-l , and J-2,
and we use the convention that the level of spin J (the starting level) is always
of even spin. The usefulness of Es in comparing different models is evident if
we relate it to the energy of the first 2+ level, Etf. Although the harmonic
vibrator is not a rotational scheme, the levels of the "quasi-/ band" in that
scheme also display a staggering similar to the /-unstable model and include
the same degenerate couplets. We can therefore include this model in this
intercomparison as well. For the four cases of the symmetric rotor, the triaxial
rotor/ Davydov model with /= 30°, the 7-unstable or Wilets-Jean model, and
the harmonic vibrator, the following analytical expressions result:
(symmetric rotor)
(yrigid, 7= 30°)
(7unstable)
Generally, if Es(4) <Ez+/3 (7-soft case), the even spin states are depressed
relative to the odd spin; if £,(4)) > E^\l!> (7—rigid), the odd spin levels are
depressed relative to the even spin ones.
Before inspecting empirical values of Es(4), we note another interesting
feature of the vibrator limit. As we have seen, empirical values of the energy
ratio R = E2 h/Elph for vibrational nuclei are typically -2.2 rather than the strict
harmonic limit of 2.0. Table 6.4 gave the energy levels of the three-phonon
states, which include the 3+ and 4+ levels of the "quasi-7band" in terms of the
two-phonon levels. Es(4) cannot be defined analytically in this case since the
unknown energies of these 3+ and 4+ levels implicitly involve the energy of the
4^ level. However, if we make the approximation that the two-phonon 2^ and
4j + levels are degenerate (i.e., e2 = e4 in the notation of Table 6.4), then the 3+
and 4+ quasi-7band states will still be degenerate at an energy 3E2$ (R - 1).
Collective Excitations in Even-Even Nuclei 195
Fig. 6.22. Systematics of Es(4)/£2t in the rare earth region. Predicted values for several models
are given at the right.
characterizes the O(6) symmetry of the IBA and nuclei in O(6) -> rotor or
SU(3) transitional regions. To anticipate that discussion briefly, it is interest-
ing to note that the classic O(6) nuclei, 192~196Pt, have larger Es(4) values than
some of the other Pt isotopes, even though they are supposedly completely y-
soft. We can already guess from Fig. 6.22, however, that nuclei with nearly y
independent potentials but with shallow minima at 30° (that is, with the
addition of a small component of triaxiality at y= 30°) might have Es(4) values
higher than the extreme y-unstable limit. This is in fact the case, as will be
commented later in the discussion of the IBA.
The y-rigid-y-soft ambiguity is not the only one that can obfuscate an
interpretation of the structure. Figure 6.22 also shows the difficulty in distin-
guishing near harmonic vibrational structure from y softness. There is one
useful, albeit qualitative, indicator stemming from the systematic behavior of
certain absolute energies that can sometimes clarify whether a transitional
region is vibrational ~» symmetric rotor or axially asymmetric —> symmetric
rotor in character. This indicator is the relation between the energy of the 2 +
(normally 22+) level and the \+ state of the quasi-ground state band.
In a transition from vibrator to symmetric rotor, the 2 + level starts off
initially degenerate with the \+ level at rather high energies (since E 2 f is then
also rather high). As the phase transition develops, the 4,+ level drops rapidly
(as the 25+ energy drops) to become a member of the ground state rotational
band, while the 2* level remains rather high in energy (although it may drop
slightly). In a transition from an axial asymmetric rotor (either soft or rigid) to
symmetric rotor, the 2r+ and 4/ levels start out nearly degenerate (for large y)
and rather low-lying, but the former rises rapidly as y -»0° while the 4^ level
drops slightly as the deformation increases. These contrasting systematics are
shown for the Sm and Os nuclei in Fig. 6.23. This signature, while valid, cannot
be quantified as well as the energy staggering since the two levels involved
have intrinsically different structure in the rotational limit: the 4^ state be-
longs to the ground state band while the 2* state is an intrinsic excitation
whose energy depends on details of its microscopic collective wave function
Fig. 6.23. Systematics of low-lying levels in the Sm and Os nuclei. The symmetric rotor limit is at
the right in each case.
Collective Excitations in Even-Even Nuclei 197
(see the discussion of the RPA in Chapter 9). In any case, it is evident that the
Ba-Gd region represents a vibrational-rotational transition, whereas the
Pt-Os region is axially asymmetric —> rotational.
This completes our brief survey of some of the essential features of geomet-
ric models for deformed even-even nuclei. These models have been
enormously successful, especially when the first order energy and B(E2)
predictions are modified by incorporating the higher order terms that reflect
changes in shape and adiabaticity of the rotational motion as J increases.
These models are all truly phenomenological in that they are applied to real
nuclei by inspecting empirical properties in order to assign the appropriate
shapes (J3,7, or higher-order moments). The models then provide a framework
for the prediction of numerous other observables.
In closing this section, it is important to stress that we have hardly exhausted
the collective vibrational modes in even-even nuclei. For example, we have
not touched on the most basic low-lying negative-parity excitations—octupole
vibrations—nor have we considered hexadecapole excitations. Just as quad-
rupole vibrations in deformed nuclei can have K = 0 or K = 2, octupole
excitations exist in K = 0,1,2, and 3 forms. In many deformed nuclei, several
of these have been identified, primarily via strong El and/or E3 transitions to
the ground state hand. The ordering of these octupole K values changes
systematically through a shell: near the beginning, the low K modes are lowest,
while they are highest toward the end. The basic physics, especially concerning
7-ray transitions into and out of these excitations, is similar in spirit to what we
have already outlined for positive parity levels. Though these excitations
account for most of the known negative parity states below the pairing gap in
deformed nuclei, they have been less well studied than their positive parity
quadrupole counterparts. Recently, however, they have taken on a new
interest in conjunction with, or rather as an alternative description to, the
possible existence of a-like cluster states, which are expected to have some of
the same properties. Hexadecapole vibrations are much less well known:
perhaps the most likely manifestation of their particular characteristics are in
the low-lying K = 4 bands in the Os isotopes.
We will not pursue a discussion of these higher multipole excitations here;
instead, we will return to the question of hexadecapole excitations and stable
hexadecapole deformations after our discussion of the Nilsson model in Chap-
ter 8, and to the structure and systematics of octupole excitations in Chapter 9
after our treatments of the TDA and RPA approaches to the microscopic
derivation of collective vibrations.
of these models, of course, had numerous offshoots: the shell model for
spherical nuclei led to the development of the Nilsson model for deformed
shapes and the collective models developed refinements that incorporated
axial asymmetry, higher moments in the shape such as hexadecapole compo-
nents, and the like. The link between these models was provided in the early
1960s by the development of microscopic approaches to collective motion
utilizing such techniques as the TDA and RPA, which we will discuss in
Chapter 9.
In 1974 the Interacting Boson Approximation (IB A) model was proposed,
which is based on a third approach that is group theoretical or algebraic and
that recalled methodology developed in the 1950s by Elliott and co-workers
for light deformed nuclei. The IBA has been extraordinarily successful, and
has also generated its own family of offshoots inspiring alternate, sometimes
competitive algebraic approaches such as the interesting pseudo-SU(S) and
symplectic group studies of Draayer and co-workers.
The IBA is a model for collective behavior. It has become customary to
refer to collective models of the Bohr-Mottelson type as "geometric" models
and those of the IBA or other group theory-based approaches as "algebraic"
models. Today one has a situation in which there is a triad of models—shell,
geometric, and algebraic—with which one can attack the basic problems of
nuclear structure. These models are not generally incompatible, although
there are differences in certain important details, but rather reflect three
approaches to nuclear structure that emphasize different complementary
aspects of that structure.
As we have commented repeatedly, the shell model rapidly becomes intrac-
table far from closed shells. In order to circumvent this, two basic alternatives
have been tried. In one, that of geometric models, the whole microscopic
approach is abandoned and replaced by a macroscopic one involving an
assumed or deduced nuclear shape, with rotations and vibrations about that
shape. The other, of which the IBA is an example, seeks to effectively truncate
the shell model space: the practical utility of such an approach depends on the
extent of the truncation, while its success depends on the appropriateness of the
truncation in isolating the key configurations involved (at least in the low-lying
states).
The truncation inherent in the IBA is shockingly extreme. For example, it
reduces the 3 x 1014 2* shell model basis states in 154Sm to 26! It is a wonder that
such a scheme can work at all, much less have the extensive and repeated
success it has enjoyed.
The basic idea of the IBA is to assume that the valence fermions couple in
pairs only to angular momenta 0 and 2 and that the low-lying collective
excitations of medium and heavy nuclei can be described in terms of the
energies and interactions of such pairs. These fermion pairs, having integer
spin, are treated as bosons (called 5 and d bosons for obvious reasons).
More formally, the model is founded on and embodies the following as-
sumptions and ideas:
• Closed shells of either protons or neutrons, and excitations out of them,
are neglected.
Collective Excitations in Even-Even Nuclei 199
many key predictions, often different from otherwise closely related geometri-
cal models. As we shall see, it confers a microscopic aspect on an otherwise
basically phenomenological model.
We now present a simplified outline of some key elements of the IBA-1
model (protons and neutrons treated together). We first discuss the bosons
and the basis states that can be constructed from them, and then a suitable
IBA-1 Hamiltonian. We then turn to a discussion of the group theory of the
IBA and its symmetry structure. Finally, we consider realistic (nonalgebraic)
calculations for actual nuclei and a simplified approach to many of these, the
so-called consistent Q formalism. Throughout, we give a number of concrete
examples of IBA predictions and stress the relationship to the geometrical
models discussed earlier in this chapter.
The basic entities of the IBA are s(J = 0) and d(J = 2) bosons, which are
assigned energies e and ed. (Note that it is conventional in the IBA literature
to use L for angular momentum both for the individual bosons, s and d, and for
the total spin of a state. Here, we keep to the convention of this book and use
J for these quantities.) A given nucleus with N + Nn valence protons and
neutrons (each counted to the nearest closed shell) has N = (Nf + Nn)/2 s and
d bosons. For example, 152Sm has N = 6 + 4 = 10 and both ^Ba^and 196Pt have
N = 3 + 3 = 6. No distinction is made whether the valence nucleons are par-
ticles or holes. Ground and excited states are formed by distributing the bo-
sons in different ways among s and d states and coupling them to different total
J. The level structures that result depend on these distributions and couplings.
The simplest situation is to imagine all TV bosons in s boson states. By
convention, es = 0, the absolute ground state. The lowest excited state will have
(N - 1) s bosons, one d boson, and an energy E = ed. The next states, in this
simplest case, will be a group with two d bosons (nd = 2). Clearly, as in the
phonon model, the two d bosons can couple to J = 0, 2, 4. This triplet will be
degenerate. Higher d boson multiplets will also occur up to nd = N. This is a
purely harmonic spectrum identical to the harmonic vibrator except for the
limitation due to finite boson number.
Since the IBA is configured explicitly in terms of s and d bosons, most of the
formalism is phrased in terms of creation and destruction operators for these
entities, s, s f , d, d f , and combinations thereof. The basic rules for operating
with these are the same as for the phonon operators b, b f used earlier in this
chapter (Eqs. 6.2-6.4). The Hamiltonian for the harmonic system just de-
scribed, is simply
that is, the energies are E = end, where for simplicity here and henceforth we
drop the subscript "d" on e.
Different states in a multiple! can be distinguished by their angular
momentum / and by the number np(n&) of d bosons coupled pairwise (tri-
pletwise) to J = 0. Sometimes one specifies not nft, but the number of bosons
not coupled to / = 0, and denotes this "boson seniority" quantum number by v.
These states form a convenient basis set for the IBA and are illustrated in Fig.
6.24. Note that for nd> 4, more than one state of a given / can occur.
Collective Excitations in Even-Even Nuclei 201
Having defined the basis states in this way, we can now consider more
general IBA Hamiltonians composed of creation and destruction operators
for s and d bosons limited to a maximum of two-body (boson) interactions. If
we keep only those terms relevant to excitation energies (i.e., if we ignore
terms contributing to binding energies), we can write
where e ', Cp v2 and v0 are six free parameters. As we have discussed, the first
term simply counts the number of d bosons and multiplies it by a d boson
energy. This gives the unperturbed energy of a state with nd noninteracting d
bosons. The second group of three terms introduces interactions between
pairs of d bosons that depend on the angular momentum to which they are
coupled but that do not change the relative numbers of s and d bosons nor mix
the basis states. The other terms have the property of changing the number of
d bosons by An^ = ±1, ±2. These terms mix different basis states of a given /
and, as in the analogous case of the shell model, it is this configuration mixing
that leads to a build-up of collectivity and to the appearance of rotation-like
behavior.
One often sees another equivalent form of the IBA Hamiltonian,
where
202 Collectivity, Phase Transitions, Deformation
The operators in Eqs. 6.51 and 6.52 are convenient combinations of those in
Eq. 6.50 that have simple physical interpretations in terms of, for example,
boson pairing and quadrupole operators. The most important point to note in
Eq. 6.51 is the And character of the various terms; those in nd, J2, T32 and T42
have &nd = 0, P TP has And = 0, ±2 contributions, while Q f has An, = 0, +1, ±2
parts.
An important aspect of IBA predictions focuses on E2 transitions. The
relevant operator, T(E2), is simply related to Q in Eq. 6.52, by
where eg is a boson charge similar to the effective charge for fermions and is
often treated as a free parameter. In the original IBA formalism, the parameter
X in Q is fixed at % - - fJ/2 in the Hamiltonian and treated as a free para-
meter in T(E2). An alternate formalism, the consistent Q formalism (CQF),
uses the same % in both //and T(E2), which leads to certain simplifications and
to a clearer physical picture of this model. This formalism will be discussed
shortly.
We have stated that the s and d boson structure of the IBA leads to a six
dimensional space and hence to a description in terms of the unitary group
U(6). We shall not delve into the group theory of the IBA in any detail, but a
few ideas are useful to understand how the symmetries so characteristic of this
model arise. Much of the following discussion is based on a recent review by
the author and D.D. Warner to which the reader is referred for additional
material on the IBA and its literature.
The basic concept underlying the group theory of the IBA is that of the
"generators" of a group. These are sets of operators that "close on commuta-
tion" (i.e., the commutator of any pair [A, B] = AB - B A either vanishes or is
proportional to another member of the group, or a linear combination thereof).
For the IBA, the 36 operators sfs, s^, d^/s and (d^dp7 where / = 0,1,2, 3,4
and l/il <J satisfy this condition and are the generators of U(6). As an example,
we show this closure for the particular pair dfs and s ts.
Using Eqs. 6.2-6.4 we have
or
Collective Excitations in Even-Even Nuclei 203
The other commutators can be similarly evaluated and indeed close on com-
mutation. This set of 36 generators of the group of transformations of U(6) is
said to form a Lie algebra.
Another key concept is that of a Casimir operator of a group. This is an
operator that commutes with all of the generators of the group. Such opera-
tors can be composed of linear or higher order combinations of the generators
and are appropriately called linear, quadratic,... Casimir operators.
The linear Casimir operator of U(6), which commutes with all 36 genera-
tors, is the total boson number operator N = d T d+ s*s whose eigenvalue is N.
This result follows trivially from the fact that all 36 combinations of the s and
doperators must conserve the total boson number. For example,
Suppose now that some smaller set of operators also closes on itself under
commutation. This set forms the generators of a smaller subgroup of U(6). It
will have linear and/or quadratic Casimir operators associated with it that
commute with all the generators of the subgroup. There are several subgroups
for U(6), so the reduction process continues until the rotational subgroup O(3)
is reached.
It is now necessary to find the quantum numbers that label the states. In
general, the generators of a group may change some quantum numbers (e.g.,
nd) but there will be one (or more) that are not changed by any of the
generators. For U(6), the 36 generators always conserve N. The set of basis
states that have a particular fixed value of an unchanged quantum number (or
numbers) is called an irreducible representation of the group.
Since the generators of a given group cannot connect different irreducible
representations, the Casimir operator(s) of a group that commute with all the
generators by definition must be diagonal and therefore must conserve all
quantum numbers, including those of the subgroups. Indeed, each Casimir
operator has eigenvalues that are functions only of the conserved quantum
numbers of the particular subgroup. Thus we have the central result that a
Hamiltonian consisting of Casimir operators of a group and subgroups cannot
mix different representations of any of the groups involved. Furthermore, its
eigenvalues are simple linear combinations of its component Casimir operator
eigenvalues and are functions of the quantum numbers characterizing each
group and subgroup. Since the quantum numbers characterizing a subgroup
are constant for all the states of the particular representation it defines, all the
states of that representation must be degenerate. This degeneracy is broken
only by the next step in the chain, which subclassifies the levels according to
another quantum number (for a subsequent subgroup). This whole process is
illustrated for one of the group chains (the so-called O(6) limit) of the IB A in
Fig. 6.25. The precise meanings of the quantum numbers and eigenvalue terms
will be clarified shortly. The key point here is the successive degeneracy
204 Collectivity, Phase Transitions, Deformation
Fig. 6.25. Illustration (using the 0(6) symmetry) of the successive degeneracy breaking in a
dynamical symmetry group chain (Casten, 1988a).
energy spectrum can be written down immediately and each state can be
labeled by appropriate quantum numbers even though these symmetries may
correspond to a complex physical situation and, in terms of Eq. 6.51, to a
complex Hamiltonian. Since transition operators can often be written in terms
of the group generators, transition selection rules appear naturally, and the
rates for allowed transitions can be written analytically. Moreover, many
ratios of transition rates depend only on general characteristics of the symme-
try (group chain) and are parameter free. This should not be surprising: the
Alaga rules for E2 branching ratios in deformed nuclei are a familiar geomet-
rical analogue.
Returning now to the basic problem in the IBA, there are exactly three
group chains of U(6) that end in O(3), which is the rotation group. This group
is a necessary subgroup in any physical chain because it provides for rotational
invariance. The three group chains can be written, along with their relevant
quantum numbers (see discussion to follow) as:
I. U(5)
The U(5) symmetry is the IBA version of a vibrator. Its representation
labels were already introduced, since this limit provides the basis states used in
most treatments of the general IBA Hamiltonian.
The eigenvalues of U(5) are
where a, /?, 7, and <5are parameters. A harmonic version of U(5) was illustrated
hin Fig. 6.24. Note, however, that U(5) is a very rich symmetry and allows much
anharmonicity. The degenerate multiplets with a given value of nd include
levels with different values of v, J and the energies can depend on these
quantum numbers. The U(5) wave functions, of course, are trivial. Since they
are themselves the normally-used IBA basis states, each wave function has but
a single term. Even highly anharmonic U(5) spectra maintain the same simple
wave functions: the anharmonicity is a diagonal effect on the energies and does
not lead to mixing of the basis states.
An interesting result concerning the anharmonicity follows from the Hamil-
tonian for U(5) written in the form of Eq. 6.50. Here U(5) includes all terms
with And = 0. Although the interactions that break the degeneracies may
206 Collectivity, Phase Transitions, Deformation
The sum on the left side of Eq. 6.59 accounts for the distribution of strength
from a given initial state if the angular momentum selection rules allow decay
to more than one level of the next lower multiplet. This sum contains more
than one term only for decay of nd> 3 states, and is identical in origin to the sum
in the phonon model expression, Eq. 6.6.
The factor (nd + 1) in Eq. 6.59 is analogous to the phonon model result
proportional to (N H +1). The factor (TV- nd) in the IBA case has no analogue
in the phonon model and arises specifically from the finite boson number. Its
origin can easily be seen. The matrix element (nd, ns s *d I nd+ 1, ns-1) can be
calculated as:
6.60. In the phonon model, a creation (or destruction) of a phonon takes place
in isolation: the transition rate is related to the number of phonons available.
However, in going from one U(5) representation to another in the IBA, the
creation (or destruction) of a d boson must involve the destruction (or crea-
tion) of an s boson to conserve N. As nd grows, there are fewer available s
bosons, and the 5 boson factor ns (or N- n^ decreases. In an (nd+ 1) —> nd
transition in the IBA, larger nd values facilitate the transition (there is more
freedom in choosing a particular d boson to destroy) but the smaller number of
s bosons hinders the transition. These two counterbalancing aspects are
reflected in the two factors in Eqs. 6.59 and 6.60.
Equation 6.59 gives, for the transitions between the lowest levels,
and
Since U(5) is usually relevant only near closed shells where N is rather small,
differences with the geometric model can be significant. For example, Eq. 6.63
gives R = 1.6 for N = 5, compared to R = 2.0 for the geometric picture. Finally,
when the initial state is the fully aligned J = 2N excitation, the factor (N- nd) is
reduced to unity. This is an example of the well-known cutoff effect in B(E2)
values involving high spin states, which is another characteristic distinction of
the IBA from geometric models.
Whenever some model predicts a symmetry, it is always a critical test to
search for empirical examples. This is particularly true for the IBA since it is
so intimately connected with the concept of dynamical symmetry. Searches for
U(5)-like nuclei naturally focus on those regions where the geometric vibra-
tional model is also appropriate. The nucleus "8Cd has recently been proposed
as a near-harmonic empirical manifestation of U(5). Its level scheme was
shown in Fig. 6.9. The E2 selection rules for U(5) (identical to the vibrator) are
clearly well-satisfied, and the small anharmonicities in the two-phonon states
reasonably account for the three-phonon anharmonicity. It is also possible to
fit many nuclei with more highly anharmonic U(5) energy spectra but strong
disagreements with the U(5) E2 selection rules and branching ratios nearly
always result.
II. SU(3)
This symmetry is the IBA version of a deformed rotor, but with special
characteristics that distinguish it from its geometric analog. The SU(3) limit is
208 Collectivity, Phase Transitions, Deformation
obtained when the Q2 term dominates in Eq. 6.51 (a J2 term may also be
present). Thus
Table 6.12. Wave functions expressed in the U(5) basis for the first three 0+ states in each limit of
the IB A
Basis States (ndngnA)
State* Limit (000) (210) (301) (420) (511) (602) (630)
U(5) 1 0 0 0 0 0 0
0+i 0(6) -.43 -.75 0 -.491 0 0 -.095
SU(3) .134 .463 -.404 .606 -.422 -.078 .233
U(5) 0 1 0 0 0 0 0
0+2 0(6) .685 .079 0 -.673 0 0 -.269
SU(3) .385 .600 -.204 -.175 .456 .146 -.437
U(5) 0 0 1 0 0 0 0
0+3 0(6) 0 0 -.866 0 -.463 0 0
SU(3) -.524 -.181 -.554 .030 -.114 -.068 -.606
*The states are ordered for pedagogical clarity and not necessarily in the order of increasing energy: indeed, the
T= 3 0+ state in 0(6) (here labeled 0+3) is usually the 0+2 state.
Collective Excitations in Even-Even Nuclei 209
Fig. 6.26. Expectation values of n in the ground band in the three limits of the IBA (Casten,
1988a).
slowly with J in SU(3) than in U(5). Third, in U(5), (n^ for a given state is
independent of N, while in SU(3), it is roughly proportional to N. This has
enormously important effects on collective E2 transitions.
The SU(3) energies are given in terms of quantum numbers of the group
chain II of Eq. 6.56
approximation to ignore these mixtures and use the usual notation K.) The
rule that determines the K values that occur in a given (A, //) representation is
K = 0, 2,. . . min(A, ju), K even. For typical values of a2 < 0, the ground state band
has (A, n) = (2N, 0). The next representations are (A, fj) = (2N - 4, 2) with
K = 0, 2 bands, (2N - 8, 4) with K=0, 2, 4 bands; and (2N - 6, 0) with a single
/£ = 0 band. These states are illustrated in Fig. 6.27. The similarities to a
deformed rotor are clear: we see sequences of states resembling a ground state
band, /3 and y vibrational bands, and bands that can be characterized as the
Pfi(K = 0), j8y (K - 2), and 77 (K = 0, 4) two-phonon intrinsic excitations.
However, we note two specific features, exemplified by the (3 and 7 bands, that
distinguish SU(3) from a general deformed rotor and act as characteristic
signatures for the symmetry. They are schematically shown in Fig. 6.28. Since
the P and 7 bands appear in the same representation, states of the same spin of
these two bands must have the same energies. Thus, SU(3) is a special case of
a deformed rotor with degenerate /3 and 7 bands. We stated that a transition
operator consisting of the Casimir operators of a subgroup cannot connect
different representations. Therefore, in the SU(3) limit, the E2 operator with
X -- VT/Z cannot lead to transitions from either the /3 or 7 bands to the ground
band! This is in direct contrast to the usual picture of harmonic collective ft
and 7vibrations in deformed nuclei. Moreover, since these bands are in the
same representation, the E2 operator leads to allowed, collective 7-* j3E2
transitions, again violating the AN h = ±1 selection rule of geometrical models.
Finally, we see from Eq. 6.65 that if we substitute (A, n) = (2N, 0) and
(2N-4, 2) for the ground and y intrinsic excitations, we obtain the energy
difference EJJ) - E (J) <* (2N-1)—the yvibrational energy increases with N
in SU(3) towards midshell.
We note that in Fig. 6.28, the collectivity of /3 -> /transitions is enclosed in
brackets because it persists even with large SU(3) symmetry breaking and, as
such, cannot properly serve as a specific signature of the limiting symmetry. It
does, however, distinguish the IBA from harmonic geometrical models.
A few specific results of SU(3) are useful to cite for practical applications.
The parameters a2 and a, of the eigenvalue expression may be written in terms
of specific level energies by inserting appropriate values of K and //. One
obtains, for example,
212 Collectivity, Phase Transitions, Deformation
Hence,
Note that these go as N2 for large N, in direct contrast with the linearity in N
characteristic of U(5). The reason is obvious and has already been hinted at.
The U(5) B(E2) values scale as N because of the (N - nd) factor in Eq. 6.59.
The nd factor is independent of N because a given pair of U(5) states (e.g., 1*
and Oj+) always have the same pair of nd values (e.g., 1 and 0, respectively)
regardless of N. In SU(3), the d f s and s f d operators in T(E2) give factors
involving N from both operators in each pair, because, as we just saw, both nd
and ns increase with N. So, a dependence on N enters twice, leading to the ~Ni
dependence.
Finally, we note, from direct substitution in Eq. 6.67, an interesting result
that we can illustrate by the ratio
The first factor is the rotational model Alaga. rule. The second factor is
(another example of) an N-dependent finite boson number effect, which
means that even in the strict SU(3) limit, B(E2) ratios deviate from the Alaga
rules. Note that the second factor goes to unity as N -> °°. That its predictions
go over into those of the usual geometrical models for large Wis a characteris-
tic feature of the IBA. Many of the unique aspects of the model (such as
allowed, collective /3 -» y E2 transitions) stem directly from the explicit
incorporation of finite N, which in turn, reflects the model's emphasis on the
valence space.
Since SU(3) is such a specific type of deformed rotor, we already recognize
that it does not characterize most deformed nuclei since such nuclei exhibit
nondegenerate ft and ybands, collective /? -»g and (especially) y-> g transi-
tions, and finite Z . (See Fig. 6.29.) Moreover, in the first half of the deformed
rare earth region E (J)-E (/) actually decreases rather than displaying a pro-
portionality to (2N-1). We will consider shortly how the IBA can treat such
nuclei. First we ask if there are any nuclei that do display the limiting
characteristics of SU(3). The answer is (probably) yes, the rare earth isotopes
of Yb and Hf near neutron number N = 104. The empirical evidence is
displayed in Figs. 6.29 and 6.30, where each of the signatures of SU(3) is
approached in the same general N, Z region (and in no other). At the same
time, it is clear that no single nucleus in the N = 104 region displays all the
Fig. 6.29. Empirical evidence relative to four of the SU(3) signatures near N = 104 (Casten,
1985b).
213
214 Collectivity, Phase Transitions, Deformation
Fig. 6.30. Comparison of empirical and SU(3) values for £ 2 + -£2 + 'n l^e rare earl
h region
(Casten, 1988a). N is ttThe boson number. r
III. 0(6)
The O(6) symmetry is the least familiar geometrically, although it is now
recognized as corresponding to a deformed, axially-asymmetric but y soft
rotor, the Wilets-Jean model. The O(6) Hamiltonian is
These expressions are identical to those of the vibrator, Wilets-Jean, and rotor
models shown in Fig 6.21. Despite the apparent differences in ground band
energies for each of the symmetries, it is important to recall that the curves in
Fig. 6.21 are defined by the characteristic quantum number nd, r, and / for
U(5), O(6), and SU(3), respectively. For U(5) and O(6), however, there is also
a separate / (/ +1) term in the Hamiltonian and, depending on the strength of
its coefficient, the actual ground band energies in these two symmetries can be
made to resemble each other or SU(3) itself.
In terms of the U(5) basis states, the nondiagonal term in //0(6) is P f P, which
has And = 0, ±2 matrix elements. Thus the wave functions are mixtures in a
U(5) basis, but are not as complex as in SU(3). Table 6.12 illustrates this,
showing that in O(6) the finite amplitudes always differ from each other by the
addition of a zero coupled pair of d bosons for a given state. This implies that
the number of unpaired d bosons is constant for a given state. It is zero for the
O/ and 02+ states, and 3 for 03\ But note that this quantity is just the boson
seniority v, and in fact, rand v are identical. The use of different notations has
historical origins only and no physical content. Indeed, v or rarise in U(5) and
O(6) because both chains involve the same subgroup O(5). This has been the
source of some confusion since the common occurrence of this subgroup
means that many predictions of the two symmetries are identical. Differences
between them do exist, but reside principally in transition rates, which depend
on the detailed d boson structure and occur among higher-lying states belong-
Fig. 6.31. Typical 0(6) spectrum for N - 6.
Collective Excitations in Even-Even Nuclei 217
also produces O(6) wave functions and spectra. This Hamiltonian has only
two terms, so it cannot give three independent terms in the eigenvalue equa-
tion. This way of producing O(6) thus leads to a special case of Eq. 6.71 where
A - B. Although this is only one of an infinite set of possible A:B relations, it
turns out to be the one empirically observed in O(6) nuclei, suggesting the use-
fulness of this alternate form for //0(6). We will see that this alternative is in fact
one limiting case of the CQF that offers a simplified approach to many IB A
problems. First, however, we need to delineate a few additional O(6) predic-
tions. The E2 transition selection rules are clear from the form of the E2
operator that is a generator of O(6), namely
The allowed transitions must satisfy Acr= 0 and AT = ±1. The first rule is a
direct consequence of the fact that a generator of a group cannot connect
different representations. It means that there are no allowed transitions from
one a family to another. Eventually, excited levels with a< N must decay, but
only by violations of the strict symmetry. This selection rule provides the most
telling contrast with U(5). (See the following.) The AT = ±1 rule is similar to
the AN h or = And = ±1 rules for the geometrical vibrator and U(5) limits.
Naturally, B(E2) values between yrast states are allowed, and given by
Fig. 6.32. Dependence of ground band B(E2) values on J in the three limits of the IB A.
8 for U(5) and A, B, C for O(6)); indeed, the U(5) scheme is the harmonic
limit. Actually, both symmetries permit a rich variety of level scheme configu-
rations by appropriate choices of parameters. Since both group chains contain
O(5) and O(3) subgroups, the only real structural differences center on the
O(6) and U(5) parent groups. Thus, it turns out that the energies of the entire
lowest representation of O(6), with a- N, can be exactly replicated in U(5).
To do so, the 02+ (two-phonon or nd - 2) U(5) level must be forced up in energy
above the nd = 3 03+ state in order to reproduce the O(6) decay pattern in which
the first excited 0+ state decays to the 22+ level. Such an O(6)-like U(5)
spectrum would be enormously anharmonic but still valid within the context of
this dynamical symmetry.
The real difference between O(6) and U(5) occurs in two other realms:
absolute transition rates and higher-lying levels. Table 6.12 shows that, though
their energies may be identical, the U(5) and O(6) wave functions are com-
pletely different. Thus, B(E2) values will be different; in particular, the
different expectation values of nd imply different finite N effects. We have
seen an example of this in the yrast B (E2) values in Fig. 6.32. In U(5), the yrast
band experiences greater changes in (nd) from state to state (see Fig. 6.26), and
hence the B(E2) values change more rapidly with spin.
High-lying levels in U(5) have high nd and decay to («/-!) levels. There are
many such (nd-V) states and thus many allowed E2 transitions. In 0(6), high-
lying levels often belong to low T states of a < N representations. These
typically have only one allowed deexcitation transition—that permitted by the
AT = ±1 selection rule.
O(6) nuclei are now well known in two regions: the Pt isoptopes, especially
Fig. 6.33. Comparison of empirical energies and E2 branching ratios in 196Pt with the 0(6) limit. The lower numbers on the transition arrows are the predictions,
the upper are the measured values (Cizewski, 1978).
220 Collectivity, Phase Transitions, Deformation
1%
Pt and the Xe-Ba nuclei near A = 130. A comparison of the 196Pt level
scheme with the O(6) limit is shown in Fig. 6.33. The agreement with all the E2
selection rules is impressive. All allowed transitions are observed and domi-
nate the decays of their respective levels; all forbidden transitions are either
weak or unobserved. At the same time, there are at least three important
discrepancies in energies: the 1= 3 0+ state is not below the 3^ level as it should
be, the splitting among the high i states (e.g., the 0+- 2+- 2+, i = 3 - 4 - 5 states)
is much less than predicted, and the energies in the 7 band are less staggered
than predicted.
The data for Xe and Ba is comparable to Pt: only a= N and TV- 2 levels are
known, but many 1 = 4,5,6 states have been assigned and the O(6) character
extends over a large range of nuclei. There are some striking analogies
between Pt and Xe-Ba. Besides their common manifestation of O(6)-like
characteristic, fits of the O(6) eigenvalue equation in each region show nearly
identical ratios A/B ~ 0.9. Interestingly, this common value is very close to the
special case A/B = 1, which corresponds to the CQF Hamiltonian of Eq. 6.73.
Moreover, exactly the same discrepancies with O(6) that occur in 196Pt are
repeated in Xe-Ba. One of these, the weaker 7 band energy staggering,
provides a useful clue to the nature of the responsible symmetry breaking.
Recall from our earlier discussion of asymmetric rotor models that the rigid
triaxial rotor of Davydov is characterized by staggering exactly opposite in
phase to that of the Wilets-Jean 7 unstable model. The O(6) limit corresponds
to the latter, that is, to a completely 7 independent potential and to a nucleus
whose shape fluctuates uniformly over a range of 7 values from 7 = 0 ° - 60°
such that 7ave = 0° but 7ms = 30°. The staggering data suggests that a realistic
potential for Pt and Xe-Ba might contain some small 7 dependence that
would shift the characteristic 7 band energy staggering pattern slightly toward
the opposite couplings of the Davydov model. It turns out that the introduc-
tion of a very small (3-4%) 7 dependence in the potential with a minimum at
30° corrects not only the 7band energy staggering, but the other discrepancies
as well. This idea is in line with our earlier discussion in the context of Fig. 6.22
in which we suggested that the E ^ / E z f d a t a could be accounted for by a
potential that was predominantly 7 independent but with a small minimum
near 7= 30°.
We commented earlier that a strict distinction between O(6) and U(5) can
not rely on the energies or branching ratios of low-lying levels alone, but
requires absolute B(E2) values or branching ratios for a< N states. Both of
these types of data are abundant for Pt and for Xe-Ba. As an example, Fig.
6.34 shows the decay of the most telling a = (N - 2) = 4 level for 196Pt and
compares this with 6(6) and U(5) predictions. The high nd U(5) levels have
multiple allowed decay routes while the O(6) a<N levels often have only one
such route. The data clearly support the O(6) interpretation.
As we have suggested in each of the preceding discussions, it appears that
good examples of all three IBA symmetries exist. This provides support for
the general structure of the IBA model and, moreover, these limiting cases
Collective Excitations in Even-Even Nuclei 221
Fig. 6.34. Comparison of the decay of the 1604 keV level in 196Pt with O(6) and U(5) symmetries
(Cizewski, 1978).
Fig. 6.35. Symmetry triangle of the IB A with the coefficients giving each dynamical symmetry.
Fig. 6.36. Empirical (upper numbers) and calculated relative B(E2) values for the Os 0(6) -> rotor transitional region (Casten, 1978).
Collective Excitations in Even-Even Nuclei 223
Fig. 6.37. Empirical and calculated properties of the transitional Sm nuclei (Scholten, 1978).
constraints of one of these symmetries. Most nuclei do not, yet the IBA is
appealing precisely because it provides a simple way to treat such nuclei. Such
treatments also reveal new aspects of collective behavior not heretofore en-
countered or expected.
To study some of these ideas, we now turn to transitional nuclei between
IBA symmetries. A convenient way to picture the symmetry structure of the
IBA is in terms of the symmetry triangle shown in Fig. 6.35. The three
symmetries mark the vertices. In terms of the Hamiltonian of Eq. 6.51, the
characteristic wave functions of each symmetry are generated by a specific
term whose coefficient is labeled at the appropriate vertex: e for U(5), a2 for
SU(3), and «0 for O(6). We have also seen that for Q to be a generator of SU(3)
one must set# = -V7/2, and that if % = 0, the same Hamiltonian produces O(6)
wave functions. These % labels are also included in Fig. 6.51.
Treating transitional nuclei is trivially simple. Structural evolution from
one limiting symmetry to another merely involves changing the ratio of the
two parameters associated with the two vertices from 0 to °°. For example, a
U(5) -> SU(3), or vibrator -> rotor, transition, such as occurs in the Nd, Sm,
and Gd nuclei near A = 150, is effectuated by varying ela2 and an O(6) -> rotor
transition, as in the Os isotopes, by changing ajar Examples of such structural
evolution are shown for these two cases in Figs. 6.36 and 6.37. The extremely
complex changes in level schemes and transition rates in these phase transi-
tions are rather well reproduced in terms of a variation of a single parameter.
While the SU(3) symmetry represents a deformed nucleus, few deformed
nuclei satisfy its strict rules. Therefore, to properly calculate the bulk of
deformed rare earth and actinide nuclei, one needs to break the SU(3) symme-
try either towards U(5) or O(6). Most deformed nuclei have E^ > E^ We recall
that a high-lying 02+ state is characteristic of O(6). This suggests that devia-
224 Collectivity, Phase Transitions, Deformation
Since J2 is diagonal and has the same effect on any states of the same spin, it
plays no structural role. Thus, if we write
we see that a2 is just a scale factor on energies and has no structural influence.
Thus the wave functions, relative B(E2) values, and relative energies of states
of the same spin are determined solely by x and N. Whereas in the traditional
form of//in Eq. 6.51 the structure and relative energies in an O(6) -> SU(3)
region depend on a0, a2, ay and E2 transitions depend separately on x, now a
single parameter determines all. The only loss of generality is that the O(6)
symmetry approached and obtained in this way is a special case, with A/B = 1.
However, this does not seem to be a deficiency, since it is this special case that
is experimentally observed.
Since predictions of branching ratios and of relative energies (of states of
the same J) depend only on x and N in the CQF, it is possible construct
universal plots for a given N, or universal contour plots against x and N. Some
examples are shown in Figs. 6.38 and 6.39. (In the energy ratio plotted, £2$ is
always subtracted to remove the structurally inconsequential effects of the J2
term.) Figure 6.38 (top) shows the behavior of the /J vibration relative to the 7
band energy. These excitations belong to the representation (N - 4, 2) in
SU(3). As 6(6) is approached, the /J vibration increases rapidly relative to the
7 band. The reason is that the 7 band goes over into a quasi-7 band sequence
starting with the T = 2 2 + level of the a = N family, while the j3 bandhead
Fig. 6.38. Three contour plots in the CQF formalism of the IB A (Warner, 1983).
225
226 Collectivity, Phase Transitions, Deformation
becomes the yrast 0+ level of the higher-lying a= N - 2 family. Figure 6.38 also
shows that for essentially all ^values, and especially for those values typical of
deformed nuclei (cross hatched box), the /? band lies well above the 7 band.
This is in excellent agreement with the data, but not unexpected given our
discussion. What is more impressive is that the B(E2) ratio in the middle panel
shows that the CQF automatically predicts that B(E2:j8 -» g) « B(E2:y^> g).
In Fig. 2.18 we saw that this feature was one of the most characteristic
empirical properties of deformed nuclei. It is particularly interesting that the
IBA predictions for realistic deformed nuclei come out this way since both
P -» g and y-> g transitions are forbidden in SU(3). However, in the transition
towards O(6), the 7 band levels go over into the quasi-7band (in Davydov
language, into the anomalous levels of the ground band) with collective
transitions to the latter, while the fi band goes over into the low-lying states of
the a = N - 2 family and the ACT = 0 selection rule forbids their decay to the
ground band. Thus they remain weak throughout the transition leg while the
y-> g transitions become collective.
The lowest contour plot in Fig. 6.38 shows one of the most remarkable and
surprising predictions of the IBA, that of collective /? o 7 transitions that are
comparable in strength to 7—* g transitions. This is completely contrary to
normal expectations and to the traditional perception of the experimental
situation. When this prediction of the IBA was initially realized, it was thought
to contradict a wide body of empirical evidence showing systematically
collective 7 —> g and weakly collective ft —> g transitions but no /? <-> 7
transitions. However, given the normal closeness in energy of E? and E and
the fact that E2 transition rates T(E2) <* B(E2) Ef, such transitions would be
extremely weak in intensity even if the matrix elements are large. Recently,
though, extremely sensitive experiments, mostly utilizing the (n, 7) reaction
and powerful 7-ray spectometers installed at an Institute Laue-Langevin in
Collective Excitations in Even-Even Nuclei 227
Grenoble (see Chapter 10), have shown that collective j3 ^> 7 transitions do
indeed exist and appear whenever the appropriate experiments have been
carried out.
It is evident in Fig. 6.38 that the y-» gift <-> 7 matrix elements generally
increase with N, and in the large N limit, ft <-» 7 matrix elements become
negligible relative to either 7-* g or g —> g intraband transitions, thus recover-
ing the geometrical picture. Of course, geometrical models can be modified
(perturbed) so as to produce ft <-» 7 matrix elements of collective strength
simply by introducing mixing between f} and 7bands. However, as we pointed
out in the bandmixing discussion earlier in this chapter, the B(E2) values so
produced result not as a correction to an unperturbed value (since the latter
vanishes), but solely from the mixing. Thus, all ft -> 7 B(E2) values are
proportional to the same mixing parameter (usually called Z i n analogy with
Z or zp and, hence, their branching ratios must be independent of the amount
of mixing; /J-7mixing provides no flexibility in the branching ratios predicted
in this way and generally disagrees with the data.
The last contour plot, that of Fig. 6.39, is also interesting. It shows a
branching ratio that vanishes in both SU(3) and O(6) limits but is finite in
between. There is no possible path between these limits that bypasses finite
values. Thus, in a totally parameter-free manner, the IBA automatically
predicts that such a branching ratio (and, indeed, many others) will peak in
transitional regions.
Fig. 6.40. Comparison of the data with predictions of the CQFfor two observables. In each case
a straight-line trajectory in % (see dashed line in Fig. 6.39) was used (Warner, 1983).
228 Collectivity, Phase Transitions, Deformation
Fig. 6.41. Observed and calculated energies in 168Hr. The J = 4 + , 5", 6' levels of the K = 4 band
predicted near 1.6 MeV are not observed below 2 MeV (Warner, 1980).
Collective Excitations in Even-Even Nuclei 229
predictions and their asymptotic approach to the Alaga rule value of 0.7 as N
increases. Thus, the IBA automatically predicts a parabolic behavior for Z
against N, minimizing at midshell, that is nearly identical to the empirical
pattern shown earlier in Fig. 6.15.
Of course, the agreement in energy for the higher bands in Fig. 6.41 is
probably at least partially fortuitous since the empirical excitations seem to
have some two-quasi-particle character. Also, the predicted K = 4 band is not
observed below 2 MeV. (It may be necessary to include g bosons in the IBA to
account for this discrepancy.)
From all these results we hope that the power of the IBA as a simple yet
general phenomological model becomes apparent. In other approaches to
collective nuclear structure, one must invoke different models (vibrator, rotor,
etc.), in an ad hoc way to accommodate different structures. In the IBA, a
single framework, embodied in the simple Hamiltonian of Eq. 6.51, encom-
passes all three limiting symmetries, and most intermediate situations as well.
It does so simply through appropriate relative magnitudes of the coefficients e,
av a0, or %, and by diagonalization of a small set of basis states. The "ad-
230 Collectivity, Phase Transitions, Deformation
Fig. 6.42. Three observables, relating to axial asymmetry, as a function of ^ (for N = 16) in the IBA
(compare Fig. 6.19).
Collective Excitations in Even-Even Nuclei 231
Fig. 6.43. Relation between the geometrical asymmetry y and the IBA parameter % (Casten,
1984). For N - 16 (the dependence on N is weak).
values of each observable, one can assign an effective 7 to each % value. This
7- x correlation is shown in Fig. 6.43. The fact that all three observables show
the same correlation supports the validity of associating each % value with an
asymmetry 7. The picture provided by this correlation is simple: increasing
deviations of % from SU(3) toward O(6) correspond to increasing axial asym-
metry and to larger and larger values of -ymi. Since the IBA never introduces a
minimum in V(j), this increase in ymt can only arise if the potential becomes
increasingly flatter in 7 as % -» 0. Figure 6.44 confirms this by showing the
effective potential V(y) for several % values. This figure shows one other
interesting point. The reader may have noted in Fig. 6.43 that 7*0° for SU(3),
even though this limit is supposed to be that of a symmetric rotor. Figure 6.44
shows the reason. Although the minimum in V(fl occurs at 7= 0°, the potential
is not infinitely steep (for finite N), and zero point motion leads to a finite 7rms.
The body of research relating to the IBA in the last decade is enormous. We
have only summarized a few highlights, emphasizing the symmetries of the
IBA, transition regions, the role of finite boson number, some experimental
tests of the model, and a geometrical understanding of it. We have completely
ignored important topics such as the intrinsic state formalism, which allows
many IBA results to be obtained analytically; the extension of the CQF to the
SU(3) -» U(5) transition region by adding an end to Eq. 6.76; the use of
effective boson numbers, especially in regions where important subshell gaps
occur that might alter the proper counting of valence nucleons; the N Nn par-
ameterization of the IBA; and numerous extensions to the model.
232 Collectivity, Phase Transitions, Deformation
Fig. 6.44. Dependence of the IBA potential on 7 for several % values (Casten, 1984, Ginocchio,
1980).
The two most important of these latter are the IB A-2, in which protons and
neutrons are distinguished and treated separately, and the IBFM, which incor-
porates a single fermion coupled to the boson core so that odd A nuclei can be
calculated. The IBA-2 has its own symmetries and has led to the recognition of
an important new collective mode, the so-called Ml scissors mode, in which
protons and neutrons oscillate (in angle) relative to each other (see Fig. 6.1).
This mode, recently discussed by Richter and collaborators in the rare earth
region, is now known to be widespread near 3 MeV excitation energy in
deformed nuclei. Its existence and properties are closely connected with a new
quantum number that arises in the IBA-2, F-spin, which describes the degree
of proton-neutron symmetry. For many IBA-2 Hamiltonians, F-spin is a good
quantum number. This classification leads to the concept of an F-spin multi-
plet—a group of nuclei, widely dispersed throughout a mass region but dis-
playing similar level schemes. F-spin also facilitates the projection of complex
IBA-2 calculations into simpler, equivalent IBA-1 cases. The development of
boson models for both even and odd mass nuclei raises the possibility of
treating both in a single unified framework. This leads to the concept of bosc-
fermi symmetries or even supersymmetries (often called SUSY's).
These are all extensive topics that merit their own discussion but that arc
well outside the scope of this book. Also beyond that ken are important
Collective Excitations in Even-Even Nuclei 233
studies extending the basic model to incorporate g bosons. These efforts aim
both to reduce the impact of "boson cutoffs" that lead to reduced B(E2) values
between high-spin states and to account for classes of excitations (e.g., K - 3
and K = 4 bands) that regularly appear among the low-lying intrinsic excita-
tions of deformed nuclei. The incorporation of g bosons has followed two
distinct paths, one in which a single g boson is introduced numerically and
allowed to interact with the s - d bosons, and another in which a full comple-
ment of g bosons is incorporated group theoretically by expanding the parent
group U(6) to U(15).
Other extensions to the IBA involve the inclusion of two-quasi-particle
excitations so that "backbending" phenomena (see Chapter 8) may be treated,
the introduction of higher-order terms such as those cubic ins or d bosons (e.g.,
(d fd f d f ) y (dddy) that incorporate a y dependence in the IBA potential, and
the expansion of the model basis withp and/bosons so that negative parity
states appear.
Finally, a most significant recent facet of the whole arena of algebraic
modelling in nuclear structure is the development of sophisticated approaches
that invoke symmetries in the Fermions directly. Some of these, like the IBA,
emphasize fermion dynamical symmetries in the valence space. Others are
even more general. They use techniques founded in symplectic group theory
to incorporate all oscillator shells simultaneously, thereby affecting important
renormalizations and offering the possibility to describe giant E2 resonances
Fig. 6.4S. An imaginary, but typical, sequence of level schemes and nuclear structure types from a
near closed shell region to a well deformed midshell nucleus.
234 Collectivity, Phase Transitions, Deformation
Fig.6.46. Normal (left) and N A^ (right) plots of £ 4 +/£ 2 + for the/1 = ISOrcgion (Casten, 1985a).
Collective Excitations in Even-Even Nuclei 235
Fig. 6.47. N Nn plot of E2< for the A = 130 region (compare Fig. 2.14) (Casten, 1985a).
236 Collectivity, Phase Transitions, Deformation
Fig.6.48. Normal (left) and N Nn (right) plots of £ 4 +/£ 2 + for the A = 130 region (Casten, 1985a).
not the total p-n interaction, but the collectivity-and deformation-driving part
of it (primarily the T = Q component). The T- I component must be identical
to the p-p and n-n T = I forces, and as we have emphasized repeatedly, aside
from the pairing interaction, these are repulsive on average and do not lead to
configuration mixing, collectivity, and deformation. (The Sn nuclei in Fig. 2.6
are a classic example of this.)
To illustrate the NpNn scheme, the data shown for the A = 130 region in Fig.
2.14 are replotted, in Fig. 6.47, in terms of N Nn instead of N or Z as is normally
done. Figure 6.48 compares normal and N Nn plots for E^\IE^\ in the same
region.
It is evident that the N Nn plot substantially simplifies the systematics: the
data that fell on several distinct curves before now coalesce so that they can be
described by a single curve for a given mass region. If such behavior is general,
the N Nn scheme offers a phenomenological tool to simplify and unify the
treatment of nuclear systematics, and has a simple underlying microscopic
basis.
However, applying the N Nn scheme to some other regions is less straight-
forward. If there are substantial subshell gaps in the single-particle level
energies, the counting of Np and Nn may be ambiguous. Moreover, the gaps
themselves may also evolve, partly as a consequence of the p-n interaction
itself. If the p-n interaction is expanded in multipoles, the monopole and
quadrupole components generally dominate. The monopole component
P0(cos S) is obviously constant as a function of the angles between the proton
and neutron orbits, and is therefore independent of the total angular
momentum J to which a pair of protons and neutrons are coupled. The
monopole p-n interaction does depend on the relative radial behavior of the
proton and neutron orbits, and therefore on which orbits are tilling in a given
mass region. Its effect is to shift the effective proton and neutron single-
particle energies. As one proceeds through a pair of proton and neutron major
Collective Excitations in Even-Even Nuclei 237
Fig. 6.49. Illustration of the eradication of a proton subshell gap at Z = 64 as a function of neutron
number due to the monopole p—n interaction.
shells, the specific behavior of the integrated monopole p-n interaction in-
duces distinct patterns of shifts in the single-particle energies that can and do
cause the appearance and disappearance of subshell gaps. Since the radial
integral is largest for orbits with similar shell model quantum numbers (see
Fig. 3.5), the largest effects in heavy nuclei generally occur when spin-orbit
partner orbits, such as g9/2p and g7/2n, are filling. In such a situation, the single-
particle energies of both these orbits are significantly lowered relative to their
neighbors. The consequences can be rather dramatic and are now thought to
be the underlying reason for the very sudden onsets of deformation near
A = 100 and A = 150. The idea is illustrated in Fig. 6.49 for the latter region. As
neutrons begin to fill the h9/2 orbit near N = 90, the strong Ih - In
interaction causes the effective single-particle energy of the lhu/2 proton orbit
to decrease and to obliterate the gap at Z = 64. Thus, for N < 90, the effective
proton shell in this region is Z - 50 - 64, while for N > 90, the effective shell is
the normal major shell Z = 50 - 82.
There is some very simple and beautiful empirical evidence for this concept
from the energies of the first 2+ states in the N = 88 and N = 90 isotones. These
data are shown in Fig. 6.50. If the effective proton shell were the normal one
from Z = 50 - 82, one would expect £2; to decrease as the number of protons
were increased past Z = 50 until the near midshell region at Z = 66, after which
£2 f should increase once again. This is precisely what happens for N - 90.
Exactly the opposite behavior, however, characterizes the N - 88 isotones.
(The N = 84, 86 isotones are similar to N = 88.) Without the concept of a
significant subshell gap at Z = 64 for these neutron numbers, such behavior
would be completely incomprehensible. But if one assumes an effective
proton shell Z = 50 - 64 for N = 88, the midshell point is Z = 57, and one would
now expect E^\ to increase for Z between 57 and 64. This is exactly the
behavior observed.
Fig. 6.51. Effective numbers of valence protons and neutrons in the A = 150 transitional region
empirically extracted from g(2 T ] ) factors. The solid lines give the normal dependence of N
against N and the changes in N against N, assuming the Z = 64 gap disappears suddenly between
N = 88 and N = 90 (Wolf, 198?£
Collective Excitations in Even-Even Nuclei 239
Fig. 6.52. N Nn plot of E2+ f°r the A = 100 region (compare Fig. 2.13) (Casten, 1985a).
The monopole p-n interaction effectively modifies the valence space for the
protons and neutrons. This in turn affects the numbers of protons and
neutrons on which the quadrupole p-n interaction may act. All this suggests
that if one defines Np and Nn based on the known evolution of these subshell
gaps, then NpNn may again be a realistic first-order estimate of the integrated
quadrupole p-n interaction strength. Figure 6.52 demonstrates this for E2{ in
the A = 100 region. Compare Fig. 2.13, which shows the same data in a normal
plot. On the right of Fig. 6.46, the NpNn plot of £4 {/ £ 2 j for the A - 150 region
was included. In these figures the protons shells are taken to be Z = 38-50
(50-64) for N < 60 (N < 90) and Z = 28-50 (50-82) for N > 60 (N > 90). The
effects are dramatic: the extremely complex systematics seen in normal plots
against Nor Z are instantly simplified and the data coalesce into single smooth
curves in NpNn plots. The existence of two curves on the right in Fig. 6.46 is not
surprising; they correspond to the two halves of the proton shell. The circled
points at N - 90, which deviate from the smooth curve, demonstrate the
imperfectness of the assumption of an instantaneous change in proton shell
structure at N = 90. Thus, the N Nn scheme is generalized to all transition
regions in medium and heavy nuclei and provides a simple, yet powerful, guide
to understanding and predicting the systematic behavior of nuclear properties.
It can also be used as a more reliable way of estimating the properties of
unknown nuclei. The reason is that, whereas in traditional plots against N or
Z such nuclei constitute extrapolations beyond the known ranges, the NiNn
values of many unknown nuclei far off stability are actually smaller than those
for known nuclei in the same region. The N Nn scheme converts the normal
Collective Excitations in Even-Even Nuclei 241
Fig. 6.53. Normal (left) and N Nn (right) plots of g 2+ for the A = 150 region. In the N Nn plot,
only the Z < 64 points are shown. The arrows point to the N Nn values for the extremely neutron
rich nuclei indicated. Predictions for £ 2 + (or other observables) for these nuclei are obtainable
simply by reading off the appropriate curve at these N N values (Casten, 1986).
Fig. 6.54. NfNn (left) and P-factor (right) plots summarizing the systematics of £ 4 +/£ 2 + in six
regions of medium and heavy nuclei (based on Casten, 1987).
noted its simplifying power. Further advantages of this approach are evident
when one compares different regions. As we saw in Chapter 2, different
regions appear to behave in entirely different fashion vis a vis the development
of collectivity and deformation: one needs only to glance at Figs. 2.13 and 2.14
or Figs. 6.46 and 6.48 for a dramatic illustration of this. In contrast, N Nn plots
are similar for essentially all regions of medium and heavy mass nuclei. Figure
6.54 (left) collects the smooth curves, drawn through N Nn plots, of the ratio
E^/E2i for six mass regions in heavy nuclei. What appeared earlier to be
radically different behavior from one region to another becomes a set of nearly
parallel curves in N Nn. The similarity in structure of these curves gives
confidence that the N Nn scheme is a meaningful indicator. It also provides
some confidence in extrapolating the curves for a new region once a few nuclei
in that region are studied.
The other feature in Fig. 6.54 is that the curves for different regions are
rather widely displaced: changes in structure are correlated with changes in
NpNn throughout medium and heavy nuclei but the absolute value of N Nn
provides little information. However, a slight modification of the N Nn scheme
puts it on an absolute scale and provides further physical insight. To see this we
consider the parameter
Fig. 6.5S. Calculated | S | values for the deformed rare earth nuclei, illustrating the saturation in
quadrupole collectivity. Compare the behavior of empirical B(K2) values in Fig. 5.3 (Casten,
1988b). The legend gives the neutron number corresponding to each symbol.
244 Collectivity, Phase Transitions, Deformation
to Z = 82, are indeed only apparent since the reason deformation sets in these
nuclei is another example of the movement of single-particle energies of one
type of nucleon as a function of the number of the other: as N decreases, the
orbit from above Z = 82 descends across the Z = 82 gap and enters the shell
below. The counting of N should therefore be based on some effective Z
value between 82 and 92.
The value of Pcrit - 4 to 5 is interesting in itself. We know from the energy
gap, 2A, in even-even nuclei that typical like-nucleon pairing interactions
have strengths of V ait ~ 1 MeV. Similarly, p-n interactions are on the order of
200-300 keV (see Chapter 4). Thus, P, which gives the ratio of the number of
p-n interactions to pairing interactions, equals 4-5 at precisely the point at
which the integrated p-n interaction strength begins to dominate the pairing
strength. This provides an appealing physical picture that simply states that
softness to deformation and the phase transition to deformed shapes occur just
when the deformation-driving p-n interaction begins to dominate the spheri-
cal-driving like-nucleon pairing interaction.
The N Nn scheme is based on the rather crude assumption that the p-n
interaction is orbit independent. Since it is not, one might expect situations in
which N Nn is not the best scaling parameter. In fact, we have already seen
evidence for this: most observables are smooth against N Nnbui they are not
linear in N p Nn . The valence *p-n interaction increases with jV N ', but is not
p n
necessarily proportional to it. The most dramatic evidence for this occurs in
the empirical behavior of B(E2:0,+ -> 2t+) values in deformed nuclei. These
increase from the beginning of a shell through the transition region. But,
instead of continuing this increase unabated until midshell, they saturate. This
was illustrated earlier for rare earth nuclei in Figs. 2.16 and 5.3. The reader will
recall from our discussion of the IB A that this B(E2:014 -» 2/) ~ N2 both in
SU(3) and in realistic symmetry breaking calculations for deformed nuclei.
Thus, assuming constant boson effective charges es, such IBA calculations
must disagree significantly with the data near midshell. These and other data
suggest that, although some simplicity will be lost, it might be useful to have a
more refined estimate of p-n strength than N Nn. This can be obtained very
easily by a simple, explicit calculation of the integrated quadrupole p-n
interaction among the valence nucleons. The result, called \Spn\ , is shown in
Fig. 6.55. Clearly, instead otNfNn, one can use these calculated \Spn values to
define effective (N Nn)eff products and therefore effective values of Np and Nn
themselves. If these are then used to recalculate quantities such as
B(E2:Oj+ -* 2j+) values, the observed saturation is excellently reproduced.
If we anticipate our discussion of the Nilsson model in the next chapter, we
can easily explain the behavior of LSJ. (Readers unfamiliar with the Nilsson
model, please forgive this short digression: better yet, return to it after reading
the next chapter.) \S J is obtained simply by integrating (summing) the
products of the individual proton and neutron quadrupole moments for each
Nilsson orbit over all filled orbits up to a given N, Z. Downward sloping lines
in the Nilsson diagram correspond to equatorial orbits, while flat or upward
sloping lines represent more polar orbits. Thus, at the beginning of a shell,
Collective Excitations in Even-Even Nuclei 245
when both protons and neutrons are filling the equatorial downward sloping
orbits, there will be high overlap and significant contributions to the integrated
quadrupole p-n interaction. Near midshell, however, some neutrons will enter
flat or upward sloping orbits that will have lower overlap with the downward
sloping proton orbits, and vice versa. These, and proton-neutron pairs in
which both particles are in flat Nilsson orbits with zero quadrupole moments,
will contribute little or nothing to further increases in the integrated quad-
rupole p-n interaction strength. This strength should therefore increase line-
arly with TV Nn at the beginning of a shell and then saturate toward midshell.
One last point is relevant before ending this discussion—the concept of a
p-n interaction scaling approximately with NfNn at the beginning of a shell and
then saturating near midshell was originally proposed as a convenient ansatz in
order to provide a simple phenomenological approach to the systematics of
certain collective observables. Very recently, however, Zhang and co-workers
have empirically extracted actual p-n interaction energies for the last proton
and last neutron in a given nucleus and, by carrying out appropriate sums,
obtained the total valence p-n interaction energies for a given nucleus. The
results show exactly the initial linearity with AyV_ and subsequent slower
growth that we have been discussing. That they do not completely saturate
highlights the fact that S n is just the quadrupole interaction. The total p-n
interaction includes the always attractive monopole component whose strength
is monotonic throughout a shell and so provides a continuously increasing
"base" to the total T= 0 p-n strength.
We have seen that the NpNn scheme, and its siblings P and ISJ plots, provide
a simple yet powerful way of correlating a vast amount of systematic data on
the development of collectivity, phase transitions, and deformation in medium
and heavy nuclei. These concepts have an appealing microscopic foundation
in the deformation-driving T~ 0 component of the valence p-n interaction and
its competition with its opposite number, the spherical-driving like-nucleon
pairing interaction. In this way, these ideas, albeit phenomenological, bring
together a number of threads running throughout this book. Through this
phenomenology, we are beginning to develop a unified, coherent view of the
evolution of nuclear structure. This view emphasizes the importance of the
p-n interaction, and its role both in modifying the underlying shell structure
and in inducing correlations, configuration mixing and deformation. What is
needed now is to graduate from phenomenology to a real microscopic theory
of nuclear structure and its evolution that embodies these ideas.
Another key issue is to understand the relation between the obvious
centrality of the p-n interaction and the equally obvious successes of collective
models that make no explicit mention of this interaction. The resolution of this
seeming paradox appears to be twofold. First, the p-n interaction determines
the "mean field," that is, the correlations and deformation of the ground state
(or the base state of a family of states such as the intruders), upon which the
collective excitations are then constructed. Secondly, the p-n interaction,
through its monopole component, affects the detailed distribution and ener-
gies of the underlying single-particle states. As we have seen, this is critical to
246 Collectivity, Phase Transitions, Deformation
the evolution of subshell gaps and therefore to the content of the valence space
on which the quadrupole component acts. Moreover, these single-particle
states are the fodder with which (see Chapter 9) the detailed microscopic
structure (e.g., energies, collectivity) of the vibrational excitations is con-
structed. Thus, the p-n interaction implicitly enters collective models both in
the equilibrium shapes they present and in the single-particle energies used in
obtaining their predictions.
7
THE D E F O R M E D SHELL M O D E L OR
NILSSON MODEL
Fig. 7.1. Illustration of two single-particle orbits at different inclinations to a prolate deformed
nucleus.
247
248 Collectivity, Phase Transitions, Deformation
shape the orbit of the single nucleon takes then depends on the extent to which
its motion is coupled to that of the core, that is, it depends on the separation of
rotational and single-particle degrees of freedom. A rigorous separation is, in
general, not possible. An approximate separation can be made, however, if the
frequency of the nucleonic motion is much larger than the frequency of the
nuclear rotation, in which case the individual nucleon executes many orbits
during a single nuclear rotation, or, alternately phrased, the nucleus is essen-
tially stationary during a single orbit of that nucleon.
This discussion is not only of formal interest, but alerts us to the possibility
that the separability of these motions may be rather poor for extremely high
rotational velocities. Modern experimental techniques have approached the
limit where the characteristic frequencies for rotational and single nucleon
motion are not distinct. It is then necessary to explicitly incorporate the effects
of the rotation on the single-particle motion. Coriolis mixing is one such effect
that we shall discuss at length, but there are others (relating to the underlying
core shape, the effective single-particle energies, and so on) that are beyond
the scope of this book. Mathematically, the separation of single-particle and
rotational motion greatly simplifies calculations and is the principle reason
why in a body-fixed frame of reference, one evaluates the single nucleon
motion first and later superimposes the rotational motion. (Incidentally, the
same basic type of problem applies to the spherical shell model, in which the
"global" motion is linear motion of the center of mass, which is more easily
distinguished.)
Fig. 7.2. Diagram defining the quantities/, K, and 9 in the discussion of the Nilsson model.
Pig. 7.3. Variation of single-particle energies of i^ orbits with different projections K (orientations
6) as a function of deformation (/} > 0, prolate, to the right).
Fig. 7.4. (a) Nilsson diagram for the Z = 50-82 regions. The abscissa is the deformation parameter
£, which is nearly the same as /3. (Gustafson, 1967).
252 Collectivity, Phase Transitions, Deformation
Fig. 7.4. (b) Nilsson diagram for the N - 82-126 regions. The abscissa is the deformation
parameter e, which is nearly the same as ft. (Gustafson, 1967).
small K). Small angles can occur either because K is low, or for given K,
because; is high. Thus, the energies of the K -1/2,3/2, and 5/2 orbits from the
h9/2 shell decrease in energy faster with deformation than those from the f7/2
orbit. This difference in rate of decrease of the Nilsson energies with deforma-
tion can overcome the small spherical f7/2-h9/2 energy separation. The low
Ktia and h9/2 orbits therefore approach each other, mixing more and more.
However, the two orbits cannot cross and so repel each other, leading to an
inflection point at the value of j8 where they would have crossed. This effect is
very clear for the K = 5/2 and K = 7/2, f^ and h9/2 orbits in Fig. 7.4.
An interesting feature of the Nilsson diagram is apparent if one looks at the
energies past the "pseudo crossing." Starting at large deformations and
tracing back toward /? = 0 the energy of the lowest K = 5/2 orbit is drawn as if
it stems from the f7/2 shell. However, one sees that it actually points directly
back to the h9/2 spherical energy. This reflects the fact that this orbit, for large
deformations, is actually the continuation of the hW2 shell. In effect, while the
The Deformed Shell Model or Nilsson Model 253
The first quantum numbers give the A" value and parity. Inside the brackets
the three quantum numbers are N, the principle quantum number of the major
shell; nz, the number of nodes in the wave function in the z direction (nz is
particularly critical for understanding the structure of the wave function); and
A, the component of the orbital angular momentum along the z, or symmetry,
axis. By definition, K = A + £ = A ± 1/2, where £ is the projection of the intrinsic
nucleon spin on the symmetry axis. Hence, one sometimes sees an alternate
notation [MizAt] or [Nn^A-l], where the arrow replaces K and indicates
whether the spin angular momentum aligns (T) or antialigns (1) with the
oribital angular momentum. The two notations are equivalent, but using K is
somewhat more convenient and common.
For the most common case of a prolate nucleus, equatorial orbits are
nearest the nuclear matter and lie lowest. It is clear from Fig. 7.1 that these
orbits are also those in which the nucleon wave function is most extended in
254 Collectivity, Phase Transitions, Deformation
the z direction. Their wave functions have the largest number of nodes in the
z direction and hence the largest values of n^. We note also that there is a
relation between the permissible values of nz and A such that their sum must be
even if N is even (positive parity) and odd if AT is odd (negative parity).
We are now in a position to label all the Nilsson orbits in a given shell. There
are two ways this can be approached, one by labeling the sequence of orbits
from each spherical; shell and another in which all orbits of a given K value are
labeled according to increasing energy. To make the labeling absolutely clear,
we will illustrate both approaches (for a prolate nucleus). We start from the
fact that the lowest-lying orbit has the highest possible nz. Clearly the maxi-
mum value of nz for a given principle quantum number N is nz = N. The lowest
orbit for a given N has K = 1/2, since this is the most equatorial orbit. Taking
N = 5 for illustration, noting that if K - 1/2, A an only be 0 or 1, and that nz + A
must be odd since N is odd, the Nilsson quantum numbers for the lowest N=5
orbit must be
This Nilsson labeling describes the K = 1/2 orbit stemming from the hn/2; shell.
Although most of the N = 5 orbits occur in the 82-126 major shell, the hllfl orbit
is the one that is pushed down by the spin orbit interaction into the lower, pre-
dominantly N = 4, 50-82 shell. Continuing for the other orbits from the hn/2
shell, the next has K= 3/2. Since its orbital orientation is slightly more inclined
away from the equatorial plane, it is less extended in the z direction and has the
next lower «z value. Its Nilsson quantum numbers are 3/2 "[541]. Again, the A
value is fixed by the requirements that nt + A is odd and A = K ± 1/2. The rest
of the hm-based orbits are then 5/2~[532], 7/2-[523], 9/2-[514], ll/2-[505].
The next N = 5 orbits are in the 82-126 shell proper, and stem from the f7/2
parent (although their actual wave functions contain large h9/2 amplitudes).
The K = 1/2 orbit, having higher energy than the K = 1/2 orbit from the h11/2
shell, must be less extended in the z direction and must have a lower nt. Its
Nilsson quantum numbers are then trivially, 1/2"[541]. The K = 3/2 orbit is
3/2"[532]. One can continue filling the entire shell in this way, and it is easy to
reproduce the labels shown in Fig. 7.4.
An alternative way is to proceed not by energy for a given parent; shell but
by lvalue. For example, the sequence of K = 1/2 orbits, starting with that from
the hn/2 orbit, will be (recall that A = 0 or 1)
l/2-[550], l/2-[541], l/2-[530], l/2-[521], l/2-[510], and l/2-[501]
The unique parity orbit in the 82-126 shell is the lowest N = 6 orbit, and there-
fore its Nilsson quantum numbers must be l/2+[660)], 3/2+[651] 13/2+[606].
Let us now consider the wave functions in more detail. These wave
functions can be written in many forms. Because they involve single-;' configu-
ration mixing, they can be expanded in a spherical basis. For many purposes
the easiest and most physically transparent form is one that the previous
discussion anticipated, an expansion in shell model orbits specified by their;
values. Thus, we write
The Deformed Shell Model or Nilsson Model 255
where the 0. are solutions to the spherical independent particle model and the
C. are expansion coefficients. Using this language, it is easy to make at least
crude estimates of the actual Nilsson wave functions. For example, for ft = 0.23
(e<= 0.95 /?= 0.22) the 5/2-[523] orbit will have a wave function that is predomi-
nantly h9/2, with the next largest component f7/2. Crudely, we can estimate that
Y (5/2~[523j) = 0.507/2 + 0.809/2 + .... We emphasize that this is at best a guess as
to the Nilsson wave functions and that the phases are arbitrary. We engaged in
this exercise simply because it is useful to have at least some, albeit crude, a
priori feeling for the structure of these wave functions. It will also remove
some of the mystery from the actual Nilsson wave functions. To take another
example, consider that the 3/2~[532] orbit for a deformation of /3 = 0.23 is just
past the inflection point with the 3/2'[521] orbit. Its wave function should be
roughly equal admixtures of f7/2 and h^ components. In contrast, the wave
function for the same orbit for a deformation of P~ 0.05 would be largely f7;2.
The wave functions near the top of the 82-126 shell are particularly simple:
V (1/2~[501]) is dominated by a p1/2 component. Near midshell, the situation is
somewhat more complicated since a given wave function will contain compo-
nents from/shells both above and below it. A particularly nice example is the
1/2~[521] orbit. Careful inspection of the Nilsson diagram shows that although
it starts out from the p3/2 shell, it soon mixes with the K = 1/2 f5/2 orbit,
undergoing a virtual crossing before /?= 0.1. At this point its wave function is
a very thorough mixture of 0M and 05/2 components. Continuing to larger
deformations, at ft ~ 0.2 there is another inflection point due to interaction with
a combination of the K = 1/2 orbits from the f^ and hM shells. Therefore, we
might expect at /? <= 0.25 that iff (l/2-[521j = Cm
with all four C. values substantial in magnitude.
As we anticipated, the unique parity orbits are extremely pure, and increas-
ingly so as K increases. For example, the 13/2+[606] orbit must (assuming no N
mixing) be pure ; = 13/2, while v^(l/2[660]) » 0.95 0m + 0.3 09/2.... (These
numbers are only rough estimates but do embody the physics and reflect the
actual amplitudes from real calculations.) Note that the only other significant
contribution besides i13/2 comes from the g9/2 orbit, which differs by two units of
/
^N«2A] 1/2 3/2 5/2 7/2 9/2 11/2
3/2-[532] 0.234 0.369 -0.560 -0.651 0.268
5/2~[523] 0.237 -0.472 -0.826 -0.196
7/2-[514] 0.323 0.938 0.128
1/2 [521] -0.510 0.345 0.473 0.431 0.444 0.120
5/2'[512] -0.023 0.836 -0.515 0.157
1/2 [510] 0.021 -0.676 0.586 -0.343 0.277 0.067
3/2'[5l2] 0.379 0.815 0.283 0.327 0.063
7/2-[503] 0.937 -0.336 0.099
9/2-1505] 0.998 0.071
m-[SQi] -0.821 -0.361 -0.411 -0.122 -0.104 -0.019
angular momentum and has its spin and orbital components aligned in the
same way as the i1M.
Table 7.2 gives examples of some actual Nilsson wave functions for a typical
Nilsson potential (8= 0.22, j3=0.23). Inspection of the table shows that all our
guesses as to the structure are semiquantitatively correct. Of course, fine
details are beyond this discussion and some amplitudes are more difficult to
intuit a priori.
It is worth pausing here to reflect on and to re-emphasize what we have
done. Without any detailed calculation whatsoever, using only simple consid-
erations of the attractive nature of the nuclear force and the nuclear shapes
involved, we have essentially "derived" the entire Nilsson diagram, the Nilsson
energies, and the asymptotic Nilsson quantum numbers. We have also dis-
cussed the basic structure of the wave functions.
At this point, however, we can obtain a deeper understanding of the Nilsson
model and diagram and of the role of the quantum number, n , by a slightly
more formal approach.
To begin, we consider the Nilsson Hamiltonian for a single-particle orbiting
in a deformed potential and inspect two instructive limits, corresponding to
small and large deformation. Actually there are many Nilsson-type Hamil-
tonians incorporating many variants of the single-particle deformed potential.
Various authors have used deformed harmonic oscillator or modified har-
monic oscillator potentials, Wood-Saxon potentials, and others. The differ-
ences reside primarily in details that do not concern us here so we will content
ourselves with the modified harmonic oscillator originally used by Nilsson.
The Nilsson model is a shell model for a deformed nucleus. It provides a
description of single particle motion in a nonspherical potential,
V - V0(r) + V2(r) P2(cos#). The original and basic form incorporated only
quadrupole deformed axially symmetric shapes. An appropriate single-par-
ticle Hamiltonian for a nucleus with symmetry axis z is:
where (ox, (Oy, and a>^ are one-dimensional oscillator frequencies in the x, y, and
z directions. This Hamiltonian satisfies the eigenvalue equation //y/. = £.y/-.,
where y/. is a Nilsson wave function written in the form y.= £.C.'0.. The / 2 and
/ • s terms ensure the proper order and energies of the single-particle levels in
the spherical limit (/3 = 0).
Although the form of the Hamiltonian in Eq. 7.2 is useful, it is also
convenient to introduce an alternate version written directly in terms of a
nuclear deformation parameter 8~ 3/2 ^ 5/4 K p~ 0.95 p. To do this, one writes
the frequencies as
The Deformed Shell Model or Nilsson Model 257
where co0 is the oscillator frequency (ftfi>0 = 41A~113) in the spherical potential
with 5=0. It is assumed that the nuclear volume remains constant as a function
of t»0. Therefore, one has the condition co^co cot constant or
The two equivalent versions of the Nilsson Hamiltonian in Eqs. 7.2 and 7.5
allow us to understand the structure of the model in the limits of large and
small deformations, respectively. Note that, in the literature the 1 • s and I 2
terms are usually expressed in terms of parameters K= C/2ft6)0 and /LL = 2D/C.
K typically takes on values around 0.06 and ju varies from 0 to =0.7.
For small deformation,;' is approximately a good quantum number. Equa-
tion 7.5 consists of a Hamiltonian for an isotropic oscillator with 1 2 and I • s
terms plus a perturbation proportional to ^Y^. The former part gives the
spherical shell model energies and is spherically symmetric. The eigenstates of
this Hamiltonian can be labeled by the quantum numbers Nlj and m of the
spherical single-particle states. Treating the Y20 term as a perturbation, the
shift in energies relative to 8 = 0 is
We can evaluate this by separating the radial and angular parts and using the
relation for a harmonic oscillator potential that
Evaluating the matrix element of the spherical harmonic Y20 gives the final
result for small 8
where we have replaced the projection m with K, the projection of the total
angular momentum on the z axis.
This simple result has three facets that account for the structure of the
Nilsson diagram for small deformations:
258 Collectivity, Phase Transitions, Deformation
and upward sloping for K > 0.65;. For example, for ;' = 13/2, orbits with
K = 1/2, 3/2,5/2, 7/2 should be downward sloping and K = 9/2,11/2, and 13/2
upward sloping. This feature is indeed displayed by the exact numerical diag-
onalizations depicted in the Nilsson diagram of Fig. 7.4. Note the interesting
physical correlation here. The angular orientation of an orbit to the symmetry
axis is approximately given by sin0= Klj and K/j-Q.65 corresponds to 0= 40°.
Inclinations greater than these are unfavored energetically by a prolate quad-
rupole deformation.
The dependence on N implies that the slopes of the energy levels in a
Nilsson diagram are steeper for larger N. Thus, heavier nuclei are easier to
deform than lighter ones. We commented implicitly on this N effect earlier
and can now explain its physical origin a bit more precisely. A nucleon in a
high oscillator shell will have a larger average radius [indeed, we just utilized
the fact that the expectation value <r2) <* (N + 3/2)]. Therefore, as the nucleus
deforms, the nuclear matter approaches this outer orbit. Since the nuclear
force is attractive, the energy of a particle in this orbit decreases. The effect is
obviously less for a particle in a lower oscillator shell that is already closer to
(or inside) the bulk of the nucleus when it is spherical.
In the opposite limit of large deformation, the 1 • s and I 2 terms in Eqs. 7.2
and 7.5 are negligible and the Hamiltonian simply reduces to an anisotropic
harmonic oscillator whose form shows that the motion clearly separates into
independent oscillations in the z direction and in the xy plane. Therefore the
number of quanta in these directions, nz and (n^ + ny), separately become good
quantum numbers. The eigenvalues of the one dimensional harmonic oscilla-
tor with quanta n. are simplyft<».(«i+1/2). This gives the familiar result for an
isotropic three-dimensional harmonic oscillator that E = nco(N + 3/2) where
N - «j + «2 + ny Thus, in the present case of large S, the eigenvalues of the
anisotropic harmonic oscillator of Eq. 7.2 go asymptotically to
mately given by Eq. 7.8. They are linear in S, j remains an approximately good
quantum number (the configuration mixing is still small), and the orbits are
separated principally by their K quantum numbers. For large deformations,
the energies (Eq. 7.10) are again linear in 5 (recall that the ft), are linear in S).
The slopes, however, now depend on n and the energies separate according to
the distribution of motion along and perpendicular to the z axis. For interme-
diate deformations, a transition between these two coupling schemes takes
place.
7.2 Examples
Having discussed the Nilsson model both physically and formally, we can now
turn to its application to odd mass deformed nuclei. Actually, this works in
much the same way as the shell model for the single-particle excitations of
spherical nuclei. The principle difference lies in the degeneracy of the orbits.
In the shell model, an orbit; can contain 2j +1 nucleons. In the Nilsson model,
the degeneracy is broken according to the orbit orientation, or K value, and
each Nilsson orbit can contain only two nucleons, corresponding to the two
ways (±K) in which the nucleon can orbit the nucleus (clockwise or counter-
clockwise). Neglecting pairing for a moment, in a deformed region the Nilsson
orbits are sequentially filled, two protons and neutrons to each, until the last
odd nucleon is placed. This defines the ground state. Excited single-particle
excitations can be obtained two ways, either by raising the last odd nucleon to
a higher orbit, thereby changing its Nilsson quantum numbers, or by lifting a
nucleon from one of the filled orbits to the last orbit, completing a pair of
nucleons in the latter and leaving a hole below the Fermi surface. One
therefore expects to have a sequence of intrinsic excitations whose energies
and quantum numbers can be simply read off from the Nilsson diagram once
the deformation is specified.
Here, in effect, is the major difference between the spherical shell model
and the Nilsson model: N = 105 corresponds to 21 holes relative to the magic
number 126. A typical shell model calculation would diagonalize some resid-
ual interaction among 21 neutrons in the 82-126 shell and the complexity
would be enormous. By switching to a deformed basis, the Nilsson model
regains a "single-particle" picture, but with deformed single-particles orbits,
each a relatively simple mixture of spherical / orbits. Multiparticle (or quasi-
particle) excitations (the equivalent of seniority v > 3 in the spherical shell
model) only begin to appear near the pairing gap at 1.5-2.0 MeV. The
deformed ansatz gives a remarkable simplification.
We recall and emphasize here that the Nilsson wave function is only a
specification of the orbital motion of the nucleons in a body fixed coordinate
system: the full specification of the wave function requires a consideration of
the rotational behavior. This is absolutely crucial for an understanding of the
structure of odd mass deformed nuclei and, indeed, for a deeper understand-
ing of the Nilsson model itself, as well as its testing and application to real
nuclei. We shall turn to the rotational motion shortly.
The Deformed Shell Model or Nilsson Model 261
It is nevertheless useful at this point to indicate how well and simply the
Nilsson model can be applied to deformed nuclei by way of a brief example or
two. Consider the nucleus 177Hf with 72 protons and 105 neutrons. All the
protons will be paired off to total angular momentum zero and, at least for the
low-lying single-particle excitations, can be ignored. The same applies to the
first 104 neutrons. Simple counting in the Nilsson scheme for e~ 0.3 shows that
the 105th neutron will enter the 7/2~[514] orbit. We therefore expect that the
ground state of 177Hf will be 7/2". (Actually, this is not so trivial: we have
implicitly assumed that the lowest angular momentum will be equal to the K
value for a given orbit. While this is generally true, it is not always the case,
especially when strong Coriolis effects are present. This is a question that must
be dealt with when we consider the rotational motion of an odd nucleus in
more detail. For the moment we accept this assumption.) A low-lying excited
state can clearly be formed by lifting the last neutron to the 9/2+[624] orbit,
giving a 9/2+ state. Similarly, one of the two nucleons in the 5/2~[512] orbit may
be raised into the 7/2"[514] orbit leaving a hole with spin 5/2~. Other low-lying
excitations should correspond to the l/2-[521], 7/2-[514], and 7/2~[503] orbits at
appropriate energies. A partial empirical level scheme for 177Hf is given in Fig.
7.5, showing the bandhead levels corresponding to each intrinsic Nilsson
excitation. It corresponds rather well with our predictions. If we now go to
179
Hf, we would expect the ground state to be 9/2^624] with the 7/2-[514] an
excited (hole) state. Moreover, all the excitations that were above the Fermi
surface in 177Hf will now decrease in energy while those that were below the
Fermi surface will increase in excitation energy. Comparisons of the two-level
schemes in Fig. 7.5, which uses the convention that particle excitations are
shown on the right and hole excitations on left, reveals exactly this behavior.
In general, as one sequences through a series of isotopes (or isotones, if one is
Fig. 7.5. limpirical bandheads of intrinsic Nilsson excitations in 177 Hf, 17'JHf. Particle (hole) states
arc on the right (left).
262 Collectivity, Phase Transitions, Deformation
Fig. 7.6. Systematics of some Nilsson orbit excitation energies in the rare earth region. Hole states
are given negative energies (extracted from Bunker, 1971).
dealing with odd proton nuclei), the energy of a given Nilsson orbit should
descend along the right-hand side of the "V." At some point it should become
the ground state, or at least occur very low in the spectrum, then increase in
energy along the left arm of the "V." At least approximately, the behavior
exemplified by the systematics in Fig. 7.6 (here hole energies are shown as
negative values) is typically observed. Deviations from it can be due to
changes in deformation across such a sequence (we have implicitly assumed a
constant deformation), or to shifts in the relative positions of the Nilsson orbits
from effects such as higher order deformation components (hexadecapole
deformations), or to Coriolis effects.
A nearby nucleus that shows one such case is 183W, whose level scheme will
be discussed at great length in the next chapter and is illustrated in Fig. 8.1.
Simple counting would suggest that the ground state is 7/2-[503], with low-ly-
ing l/2-[510] and 3/2~[512] particle excitations. Yet, the empirical level scheme
shows that the latter two orbits are near the ground state and the 7/2" [503]
occurs at a few hundred keV excitation energy. An explanation of this will be
given in Chapter 8.
We have seen that it is as easy in the Nilsson model as in the shell model to
determine the expected order of single-particle excitations and their energies
and to deduce, virtually by inspection, an anticipated level scheme. Though
this seems a trivial exercise, one should not lose sight of the fact that by
considering a deformed shell model potential, one is able to account instantly
The Deformed Shell Model or Nilsson Model 263
for the low-lying levels of literally hundreds of deformed odd mass nuclei,
ranging from A ~ 20 to the actinides. There have been innumerable tests of this
model over the last three decades and it has proved capable of correlating a
vast amount of data, particularly when some rather simple refinements (pri-
marily Coriolis mixing and hexadecapole deformations) are incorporated.
This is almost, but not quite, the desired explanation. One other ingredient
is necessary. To see this, consider the sequence of spherical / values (see Fig.
3.2, or Fig. 7.4 for ft = 0). Suppose that, instead of this order, the lowest orbit
had been p1/2, pM, or even f^ instead of a higher j orbit. The preceding
argument would then have had little weight. Upsloping orbits would soon be
encountered for both prolate and oblate shapes. It is because the lowest orbits
after a shell closure have relatively high j, with many K values, that a distinc-
tion between the oblate and prolate behavior can be made and a preference for
prolate deformations can develop. Thus, the second key feature is the modifi-
cation of the shell model potential to include components that favor lower
energy for higher / and; orbits. The nearly universal preference for prolate
shapes in nuclei stems from the specific radial shape of the shell model central
potential that is intermediate between harmonic oscillator and square well,
which favors large I values, and from the properties of the sine (the orbit
inclinations as a function of K). Once again, we have an example of how a very
simple but physically intuitive appreciation of the structure of a given model
leads to important results even without detailed calculation.
Fig. 7.7. Angular momentum diagram for an odd mass deformed nucleus. Note: this figure can be
very misleading—see text and Fig. 7.9.
sion in even-even nuclei can also be used for odd mass nuclei. If the odd
nucleon, considered for a moment to be in a single; orbit, does not polarize the
even-even core, then the total angular momentum results from the vector sum
of the core rotation and the odd particle angular momentum. This is illustrated
in Fig. 7.7. We now start with the same rotational Hamiltonian as before and
obtain, using the notation for the different angular momenta given in Fig. 7.7,
We can convert this to a more useful form by defining the familiar raising and
lowering operators
and hence
Replacing these operators with their eigenvalues where possible and using the
fact that, for low-lying states, both J3 and/, have the same projection K, gives
266 Collectivity, Phase Transitions, Deformation
or
where
The first term on the right in Eq. 7.14 is identical to the rotational energy
expression for an even-even nucleus. It is clear from Fig. 7.7 that this must be
the case, since this figure would collapse to that of an even mass rotor if there
were no single-particle angular momentum;. The other terms in Eq. 7.14 arise
specifically from the presence of the odd particle and are intimately connected
with the coupling between rotational and particle degrees of freedom. Even
the resemblance of the first term to the symmetric top formula is fundamen-
tally misleading. We shall see momentarily that the simple picture illustrated
in Fig. 7.7 conceals some important physical effects, and that the rotational
motion is not as simple as commonly believed, but an alternate picture that is
nearly as simple will allow us to retrieve Eq. 7.14 in a transparent, elegant way
that will disclose a much different understanding of rotational motion in odd
mass nuclei.
The third and fourth terms in Eq. 7.14 involve the / structure of the de-
formed single-particle wave function (Nilsson wave function). Whatever
value (f) takes, both it and K should be constant within a rotational band
(neglecting band mixing). The last term is called the Coriolis interaction
because its effects are very similar to the classical Coriolis force acting on any
rotating macroscopic body. The Coriolis interaction has important conse-
quences in both even and odd deformed nuclei and will be extensively dis-
cussed later. We will show then that, in first order, its effects on energies simply
correspond to a change in the WLI.
To this order then, the energy levels of a given rotational band in an odd
mass nucleus should behave as ft2I2IJ(J + 1). In Fig. 7.8, we illustrate a number
of examples of rotational bands in heavy nuclei and show that this simple
formula works remarkably well. Also included in the figure are two examples
where it clearly fails to provide even a reasonable first-order estimate. One of
these involves a K - 1/2 band (1/2~[521] in 169Er) that we shall later see incor-
porates a special (diagonal) Coriolis interaction. The other is a "band" ("h11/2"
in 133La) that appears to be partly "upside down" (e.g., E(ll/2-) « E(5/2-),
E(7/2~)) and unrelated to the kind of structure that we have been examining. It
too involves especially strong Coriolis effects and will be discussed later. Here,
we wish to raise, and resolve, an apparent paradox that arises from Fig. 7.7
(and that is intimately related to this type of inverted structure).
Equation 7.11 states that the total angular momentum results from the
vector combination of the rotational angular momentum R and the particle
angular momentum/, that is J = R + j. The situation was sketched in Fig. 7.7,
which is a simplification since the Nilsson wave function contains, in general, a
linear combination of functions of different/. That is not the point. Consider,
The Deformed Shell Model or Nilsson Model 267
Fig. 7.8. Rotational bands in some deformed odd-mass nuclei. The dots give the rotational energy
predictions from Eq. 7.14 after normalization to the first two levels. Dots for the energies of the
K = 1/2,1/2~[521] band in 169Er are omitted since they require a "decoupling parameter" term (to
be discussed in Chapter 8). As discussed later in the text, the rightmost "band" is so Coriolis mixed
that no single Nilsson label is possible.
for simplicity, the one case in which a single / value does nearly characterize the
Nilsson wave functions—the unique parity orbits. To be specific, let us take a
Nilsson wave function such as that for the l/2+[660] orbit from the i13/2 neutron
shell in the rare earth region. (For convenience we neglect Coriolis mixing.)
Now, we have seen examples of rotational bands with spins J = K, K + 1,
K + 2,... whose energies vary approximately as J(J + 1), or in this case, a
sequence with J = 1/2,3/2,5/2,7/2,.... The common view (Fig. 7.7) of this band
as consisting of a single particle in the l/2+[660] orbit coupled to a sequence of
successively faster core rotations is seriously in error. To see this, recall that we
have taken a simple case where the Nilsson wave function consists of only one
; value,; = 13/2. This is therefore the only single-particle angular momentum
in the system. This in turn implies that any total angular momentum other than
/ = 13/2 must incorporate angular momentum from another source. That
source can only be the rotational motion. Thus, as in Fig. 7.7, one can imagine
a J = 17/2 state obtained by coupling a; = 13/2 single-particle angular momen-
tum to a core rotational angular momentum R = 2. Similarly, J = 9/2,5/2 states
could be formed by the antiparallel coupling of / = 13/2 and R = 2, R - 4,
respectively. The only energies in the system are the Nilsson energy, which is
constant (independent of J), and the rotational energy, fi.2/2IR2. Hence, the
J = 9/2 state (with R = 2) should have higher energy than the/ = 13/2 state (with
R = 0). The J = 5/2 and 1/2 states should be expected still higher. Moreover, the
energy difference £(9/2) - £(13/2) should equal £ R _ 2 - £R=0, or in other words,
£2\ of the neighboring even-even core nucleus; £(5/2) - £(13/2) should equal
£4^ , and so on. This picture leads to "upside down" rotational bands with the
268 Collectivity, Phase Transitions, Deformation
lower spin states lying higher than the state with J=/. Such band structures do
indeed exist (as we noted in discussing the right-most band in Fig. 7.8). They
have recently become highly interesting as a particular manifestation of im-
portant Coriolis effects in high-spin states. However, they are not the normally
observed situation and are certainly not consistent with most of the empirical
rotational bands shown in Fig. 7.8. Clearly, there is something wrong with this
picture.
A clue to a more accurate understanding of the rotational motion begins by
recalling that we are dealing with an axially symmetric deformed nucleus. This
means that any orientation of the angular momentum vector j with respect to
the symmetry axis z that maintains a projection K is indistinguishable from any
other orientation and therefore is equally likely: the angular momentum
vector j is free to precess around the z axis. Figure 7.7 showed only one
particular orientation of this angular momentum vector—that corresponding
to the smallest possible value of \R\, since j and R are aligned. In contrast,
imagine that the angular momentum j were rotated 180° to that shown in the
figure so that it lay in the plane of the page but pointed downward, below the
z axis. This situation is depicted on the left in Fig. 7.9. The amount of core
rotation required to produce a final total angular momentum J would clearly
be much larger. If we extend this idea to other angles of the angular momen-
tum vector j, then, as j processes, R will point in a continually varying direction
and I R \ will take on a constantly changing series of values.
Fig. 7.9. (Left) Angular momentum diagram for an odd-mass deformed nucleus. This is a more
rigorous version of Fig. 7.7 that incorporates important refinements. (Right) Path (circle)
followed by the tip of the j vector with time as it precesses about the z axis (point C). This is an
"end view" of the time dependence of the diagram on the left. I am grateful to D. D. Warner for
this figure.
The Deformed Shell Model or Nilsson Model 269
but
and
Thus
This is exactly the same as Eq. 7.14 except for the Coriolis term that has not
been included since we have assumed that K is constant and, as we shall see,
the Coriolis interaction inherently mixes different K values. (Since we as-
sumed a single-/ wave function, (j2) has become /(/ + 1).) Note that the
assumption of constant K is equivalent to the assumption that the particle
angular momentum vector j precesses exactly about the z axis. We shall
shortly encounter a special, though not uncommon, situation in which preces-
sion is about an axis perpendicular to the z axis: clearly in such a situation it will
be K, not R, that changes continuously. This then will bring us back to the
picture in Fig. 7.7, which leads to "inverted" rotational sequences.
We therefore see that the derivation of Eq. 7.14 for the eigenvalues in a
rotational band built on a given Nilsson orbit was correct, even though the
270 Collectivity, Phase Transitions, Deformation
simple picture that is commonly used to illustrate this situation, Fig. 7.7, is too
simplistic. The more accurate view shown in Fig. 7.9 gives the same formula in
a trivial manner. It has been a constant emphasis in this book that most results
in nuclear structure physics can be derived, at least semi-quantitatively, by
very simple, often intuitive, analyses. This example warns us that such an
approach cannot be careless handwaving, but must accurately reflect the
correct underlying physics.
The most important conclusion from the present analysis is the recognition
that the rotational motion in a deformed odd mass nucleus is far from simple.
Not only does the magnitude of the rotational angular momentum I R \ vary
with time but its direction in space also changes. In the / = 1/2 state, for
example, the nucleus at times rotates clockwise about the y axis, at other times
counter clockwise about this axis, and at still other times about a continuously
varying axis in the xy plane. The rate of rotation is relatively constant
corresponding to \R\ =6, In the 7=13/2 state, on the other hand, the nucleus
is at times stationary (R = 0) while at others it rotates at frequencies varying
from values corresponding to R = 0 to those for | R \ = 12! We can see that the
rotational motion is really a complex, time dependent, variation that includes
not only a true rotational component but also a kind of tumbling motion. The
fact that the principle term in Eq. 7.14 (i.e., the term that depends on /, the
others being constant) has the same form as in the symmetric top or in the
rotational energy expression for an even-even nucleus, is really almost an
accidental result of the particular combination of the processing single-particle
motion and the varying core rotational motion needed to produce a constant
total momentum J*.
For typical (nonunique parity) Nilsson wave functions, several / values
commonly appear and the "rotational" motion is even more complex. From
the formal standpoint, one need not worry about this since, as we saw in
deriving Eq. 7.14, as long as the proper vector character of the angular
momenta are taken into account the correct results always emerge.
There is one other important point implicit to this discussion. It should be
clear by now that while many values of I R \ contribute to the wave function for
a state of spin / in a rotational band built on a given Nilsson orbit, the special
value R = 0 can only occur if the Nilsson wave function y/Mfa contains an
amplitude 0,—that is, only if it contains an amplitude for the single nucleon in
the orbit with j = J. For any other / value, the state with total angular
momentum J can be constructed only by incorporating some rotational mo-
tion. As we shall see, this has important consequences for single nucleon
transfer reactions. We recall that for a given oscillator shell N, ;mai = N +1/2
(e.g.,;mai = 11/2 for N = 5). It is thus clear that all states with J>N + l/2 must
have R * 0.
Having discussed the basic Nilsson model and the basic rotational motion in
odd mass deformed nuclei, we turn in the next chapter to a detailed discussion
of specific tests of and refinements to the model, with emphasis on the crucial
and pervasive effects of Coriolis coupling.
'The author is grateful to D. D. Warner, with whom this analysis was worked out.
8
NILSSON MODEL:
APPLICATIONS AND R E F I N E M E N T S
271
272 Collectivity, Phase Transitions, Deformation
and emerge from the collision as a triton, leaving behind a residual nucleus
with A' = A-l. This is a (d, t) reaction. Or, a neutron could be stripped off and
enter an orbit around the target nucleus, producing a final nucleus with mass
A' = A +1, a (d, p) reaction. These reactions can be experimentally selected by
detecting the outgoing particle and identifying it. This can be done with a
number of different techniques we will not discuss here. Some utilize the
different magnetic rigidities of the outgoing particles. Others exploit the
dependence on the mass and charge of the ratio of the energy loss AE in a thin
detector to the total energy.
In any case, we assume we have identified an outgoing proton, thereby
"tagging" a (d, p) reaction event. This reaction is not necessarily a single step
process. It could be accompanied by inelastic scattering or Coulomb excita-
tion. Or it could be the result of a compound nuclear reaction in which the
deuteron was first fully absorbed. To select appropriate experimental condi-
tions to favor a simple single step process, we consider some of the parameters
describing the reaction process of Fig. 3.3. If the closest distance of the
incident projectile from the target nucleus is large, the interaction is weak, and
stripping occurs with low probability. When scattering occurs it will not be at
large angles. Two-step processes involve a product of such single-step ampli-
tudes, and will be negligible. For a close collision (distance of closest ap-
proach), the nuclear interaction and hence the scattering angle are much
greater. The probability of a single event occurring is much larger but so is that
for multistep processes. The optimum situation of large, direct, reaction cross
sections but small multistep amplitudes occurs for an intermediate angle
corresponding to a "grazing" collision. This discussion is highly qualitative
since quantum mechanical interference effects lead to significant oscillations
of o(d, p) with 9. Nonetheless, for targets with A > 100 and typical incident
deuteron energies Ed~ 10-15 MeV, most (d, p) experiments emphasize detec-
tion at angles 40° < 6< 125°. Similar considerations apply for heavier projec-
tiles, except that, for a given bombarding energy per nucleon, the heavier
projectile brings in more angular momentum and is more likely to excite final
states involving larger angular momentum transfer.
With this brief digression finished, we turn to consider the nuclear structure
information obtainable from single-step reactions such as (d, p). We consider
the population of members of a rotational band built on some Nilsson orbit in
a deformed odd mass final nucleus. Since the reaction is single step, sequential
processes, such as transfer followed by inelastic scattering, are eliminated by
the choice of experimental conditions. Indeed, by definition—the only thing
that can occur is that a single nucleon can be transferred to a given, quantized,
empty valence orbit (Nilsson wave function t//Nils). In particular, the process
cannot induce any rotation of the target nucleus. Since no rotational notion
can be imparted, it follows that the probability of populating a state with a
given total angular momentum J must be proportional to the probability, C2._,,
for a shell model single-particle wave function 0 ; in the Nilsson wave
function. Thus, even though the intrinsic wave function for each state in the
rotational band is identical, the (d, p) cross section for populating each
Nilsson Model: Applications and Refinements 273
where we have included a pairing factor denoted by P2, which we shall discuss
in a moment. (The quantity (C P)2 is analogous to the spectroscopic factor
(see chapters 2 and 3) for spherical nuclei.) This formula is extremely well
known and has been extensively used for two decades in probing the structure
of Nilsson wave functions; strangely, its simple origins described earlier are
often only vaguely understood, and the power of (d, p), (d, t), and other single-
particle reactions in elucidating Nilsson structure often seems almost magical.
Before discussing this equation in relation to the empirically deduced
structure of various Nilsson orbits in typical deformed nuclei, we must add a
few more comments on the other factors appearing in it. In a formal derivation
of Eq. 8.1, there must occur an integral linking the initial and final states
even-even nucleus, the final angular momentum J = j = I ± 1/2 and the final
state parity is n= (-1)'. In principle, a measurement of the angular distribution
of the outgoing particles can provide information on the J* values of various
final states. In practice, this information is somewhat unreliable in deformed
nuclei, and measurements are typically made at only two or three angles: the
ratio of the cross sections at these angles provides at least a qualitative guide to
the transferred orbital angular momentum /. On account of the centrifugal
barrier, it should not be surprising that, for low-energy, light projectiles, the
population of higher / values is inhibited: the cross sections decrease with
increasing /(and therefore J). Generally, it turns out that the angular distribu-
tions for small / values are somewhat forward peaked, while those for large /
transfers are backward peaked. Therefore, a ratio such as a(125°)/o"(600)
increases with transferred orbital angular momentum /. The cross sections
<t>t(9) also have a dependence on the reaction Q value (the difference in
incoming and outgoing projectile energies). The Q value is easily deduced
from the known nucleon separation energies. For example, for a (d, p)
reaction
where B.E.(d) is the deuteron binding energy 2.23 MeV. Since S(n) = 5-8 MeV
in heavy nuclei, Q(d, p) is typically positive for low Ex and decreases as Ex
increases.
As noted, the use of reactions that carry more momentum into the system
such as (3He, a), favors high / transfers. Therefore, the ratio of populations of
a given state in (a, 3He) and (d, p), a(a, 3He)/o(d, p), can also serve as a
"meter" for the transferred angular momentum /. Indeed, at back angles, it
singles out the highest// values accessible (e.g.,/ = 13/2 in the odd neutron rare
earth nuclei: see the following discussion (Fig. 8.9)).
The factor P2 is U2 for a stripping reaction such as (d, p), and V2 for a pickup
reaction such as (d, t). It represents the probability that the single nucleon
orbit involved is initially either empty or filled, respectively. It is reasonable
that this factor is present. In a (d, p) or (d, t) reaction, a given orbit can be
populated only to the extent that it is initially empty or full, respectively. Thus
(d, p) tends to populate orbits above the Fermi surface, while (d, t) populates
orbits below the Fermi surface most intensely.
Now we can turn to the extraction of specific nuclear structure information
from these reactions. It is easiest to show this using a specific example.
Consider the final nucleus 183W whose level scheme is shown in Fig. 8.1 with the
states arranged according to Nilsson assignment and rotational band in the
same format as in Fig. 7.5 for 177>179Hf. Figure 8.2 shows (d, p) and (d, t) spectra
leading to 183W, while Table 8.1 summarizes the measured cross sections (at
90°) for those negative parity states that were assigned to specific Nilsson
states. Table 8.2 gives similar information for 185W, We assume that the
"kinematic" factors Afy, (Q) are known so that the cross sections may be used to
extract empirical values of OP for the ith band.
Look at Fig. 8.2, bearing in mind the strong / dependence of the DWBA
276 Collectivity, Phase Transitions, Deformation
Fig. 8.Z (d, p) and (d, t) spectra leading to 183W. The peaks are labeled by the Nilsson assignments
(Casten,1972).
Table 8.1 Comparison of unperturbed, Coriolis coupled and experimental cross sections in 183W
(6 =90°)*
C^d.p)^/^ o(d,t) pblsr
State Unper- Per- Experi- Unper- Per- Experi-
turbed turbed mental turbed turbed mental
9/29/2~[505] 64 65 25 1.8 1.2
7/27/21503] 357 364 284 38 48 72
9/2 7.0 8.0 =3 0.6 1.3 =1
3/23/2~[512] 138 208 131 29 60 58
5/2 249 184 96 48 34 35
7/2 30 48 87 5.8 17 44
9/2 6.0 3.9 4 1.0 1.0 <1
11/2 0.2 0.4 4 0.0 0.1 1
1/2 1/2~[510] 0.3 0.4 8 0.2 0.6 5
3/2 334 269 264 195 179 150
5/2 98 163 202 53 65 103
7/2 33 18 11 18 15 14
9/2 3.3 5.5 11 1.6 1.6 7
5/25/2-[512] 0.0 0.0 6 0.2 0.2 19
7/2 17 7.9 29 269 217 237
9/2 1.2 1.2 3 14 25 18
11/2 0.1 0.0 1.3 0.8 4
l/21/2-[521] 25 15 15 280 280 364
3/2 7.0 2.8 3 128 114 54
5/2 5.1 4.3 86 87 66
7/2 4.3 1.8 71 64 69
9/2 0.7 0.6 10 10 8
7/27/2-[514] 2.0 3.8 16 40 81 91
912 2.6 3.0 3 45 35 29
11/2 0.0 0.1 <4 0.8 1.3 15
•Casten, 1972.
is to assume a Fermi energy, and the simplest assumption here is to assume that
it coincides with the energy of the Nilsson orbit that forms the ground state in
the odd mass nucleus. For a reasonable choice of the gap parameter A (typi-
cally 0.75-1 MeV), it is easy to solve the quasi-particle Eq . 5. 23, to obtain U
and Vas a function of excitation energy. However, the Fermi energy need not
coincide exactly with any specific Nilsson orbit. For example, in 183W the
ground state is the l/2^[ 510] orbit (see Fig. 7.4) and the 3/2~[512] orbit occurs
at an excitation energy of approximately 200 keV, while in 185W the order is
reversed but the two bands occur within =20 keV each other. If the deforma-
tion has not changed, one cannot account for this asymmetric situation by
placing the Fermi surface at, say, the position of the 1/2^[510] orbit in 183W and
at the 3/2~[512] orbit in 185W. Given the separation of these two orbits in the
Nilsson diagram, <= 150 keV, it is clear that in 185W, the Fermi surface must be
approximately centered between these two orbits, producing (see Fig. 5.7) low
excitation energies for both. Since the U and V factors change rapidly near the
Fermi surface, such fine details of the Fermi surface location can have signifi-
cant effects on single nucleon transfer cross sections for low lying orbits.
278 Collectivity, Phase Transitions, Deformation
Table 8.2. Comparison of unperturbed, Coriolis coupled, and experimental cross sections in ^W
(0 = 90°)*
d(d,p) fjb/sr a(d,t) i^b/sr
State Unper- Per- Experi- Unper- Per- Experi-
turbed turbed mental turbed turbed mental
9/29/2-[505] 66 62 25 3.3 3.9 10
7/27/r[503] 334 341 316 58 72 154
9/2 6.4 10 1.0 2.1 =3
11/2 0.5 0.5 -6 0.1 0.2 1
3/23/2-[512] 107 30 5 59 1.7 1
5/2 195 247 301 97 173 207
7/2 23 7.2 11 12 1.0 1
9/2 4.7 6.5 11 2.1 4.5 11
11/2 0.2 0.1 0.1 0.0 =0.4
1/2 1/21510] 0.2 0.2 4 0.3 0.9 3
3/3 180 262 357 343 423 308
5/2 53 0.3 =4 92 14 11
7/2 18 37 104 32 55 99
9/2 1.8 0.0 2.7 0.1 =1
11/2 0.1 0.2 5 0.2 0.4 4
5/25/2^512] 0.0 0.0 <3 0.2 0.2 14
7/2 17 7.7 32 268 208 206
9/2 1.2 1.1 14.1 24 =20
1/2 1/2~[521] 14 14 <62 281 280 =266
3/2 6.4 2.8 129 107 =43
5/2 4.7 3.8 86 89 =20
7/2 3.9 1.6 72 60 =25
9/2 0.5 0.6 10 10 11
7/27/2"[514] 2.1 4.1 15 40 86 80
9/2 2.7 1.2 45 34 19
*Casten, 1972.
A more empirical way to extract U and V factors is from the ratio of (d, p)
and (d, t) cross sections to the same state in an odd mass nucleus of mass A.
This procedure is slightly inconsistent, since the U2 factors refer to the empti-
ness of Nilsson orbits in the target nucleus A in the (d, p) reaction while the V2
factors relevant to (d, t) refer to the orbit occupancies in its A + 2 target
nucleus. Nevertheless, this technique is widely used and is adequate for
essentially all cases of practical importance.
We could now extract empirical C values from the measured (d, p) and (d, t)
cross sections. For technical reasons, it is easiest and most common to use
theoretical sets of C coefficients to calculate theoretical cross sections and to
compare these with the measurements.
By identifying appropriate sequences of cross sections, the states in an odd
mass nucleus may be sorted into rotational hands whose Nilsson wave func-
tions can be identified. Such assignments are made in Tables 8.1 and 8.2 for
183, iss^ an(j [he theoretical and experimental cross sections are compared. In
Nilsson Model: Applications and Refinements 279
Fig. 8.2, the deduced J values and Nilsson quantum numbers are indicated
above the corresponding peaks.
Careful inspection shows very different patterns for the different rotational
bands, justifying the term fingerprint patterns. For example, the 1/2~[510] band
has a very small cross section to the l/2~ state, large cross sections to the 3/2-
and 5/2- levels, and smaller cross sections thereafter. The 7/2-[503] band has a
large cross section only for the 7/2" state. For the positive parity levels,
essentially only the 13/2+ states are populated. (Note that we have not
attached specific Nilsson quantum numbers to some of the latter levels. The
reason will be clear after we have discussed the strong Coriolis mixing between
these bands.) The U1 and V2 dependence in Eq. 8.1 is also evident in Fig. 8.2.
For the low-lying bands where V1 and V2 are roughly comparable, the same
states were populated in both (d, p) and (d, t), whereas at higher energies, the
levels separate according to whether they are populated in stripping (LP) or
pick up (V2). For example, the 1/2 1/2" [521] and 7/2 5/2- [512] hole states are
stronger than the ground band states in (d, t), but much weaker in (d, p).
Tables 8.1 and 8.2 are well worth careful inspection. Although there are
small differences in detail, especially for weaker states, the characteristic
fingerprint patterns are often observed experimentally. Examples are the
5/2-[512] and l/2-[510] bands in 183W and the 7/2-[503] band in 185W. Indeed,
these fingerprint patterns are often the technique used to identify the specific
Nilsson orbits in the first place. The reader should not minimize the impressive
successes of such a simple model, many of whose predictions can be antici-
pated without calculation despite the presence of perhaps dozens of valence
nucleons.
Nevertheless, while qualitative patterns emulate the data, the detailed
predictions often disagree substantially with the experimental results. Ex-
amples are the 5/2 3/2-[512], 5/21/2-[510], 3/21/2-[521], and 7/2 7/2-[514] states
in both nuclei. There are differences of nearly an order of magnitude in the
cross sections for populating certain corresponding states in the two nuclei.
For example, in (d, p) in 183W, the 5/21/2~[510] state is strongly populated while
the 5/2 3/2-[512] state is weak; in 185W, it is just the opposite.
Both of these phenomena are striking manifestations of the importance of
the Coriolis interaction in odd mass nuclei. As we shall see, we can greatly
improve the predicted cross sections if we take this residual interaction into
account. The importance of the Coriolis interaction goes far beyond the
question of sorting out difficulties with single nucleon transfer reactions. It is
especially important for high-spin states in both odd and even mass nuclei, and
has been shown to lead to a new coupling scheme—the so-called rotation
aligned coupling scheme characterized by "decoupled" bands in many nuclei,
and by the backbending phenomenon. It is in fact difficult to overestimate its
significance in understanding odd mass deformed nuclei. We turn now to a
systematic treatment of the Coriolis interaction with emphasis on its physical
origin, its principal effects, and a simplified discussion of some easy ways to
estimate its effects by inspection.
280 Collectivity, Phase Transitions, Deformation
Kg. 8.3. Coriolis effect in a rotating system resulting from the dependence of rotational velocities
on "latitude."
Nilsson Model: Applications and Refinements 281
projectile that was fired directly northward will appear to be deflected toward
the east. This is the Coriolis effect, and while it is sometimes called a fictitious
or apparent force, it has real physical effects. (It accounts for the fact that river
banks tend to be eroded more on the right (facing downstream) side in the
northern hemisphere.) We see from this illustration that the Coriolis interac-
tion effectively tilts the orbit relative to the equator. If we now picture the
orientation of the angular momentum vector perpendicular to the orbit in-
stead of the orbit itself, the Coriolis effect is equivalent to a change in its
projection onto the equator. Thus it is understandable that, in the nuclear case,
the Coriolis interaction alters the projection of the angular momentum K on
the symmetry axis, admixing different K values. Another way of looking at
this that will be useful later is to recall that K is only a good quantum number
if the nuclear potential is axially symmetric. Therefore, the Coriolis interac-
tion effectively introduces small amounts of axial asymmetry as it mixes K
values.
We now evaluate the Coriolis matrix element explicitly. We consider two
intrinsic Nilsson states characterized by K and K +1. Since J is a good quantum
number, we can replace J± with its eigenvalue. This cannot be done with j
because of the configuration (j) mixing in Nilsson wave functions, although it
can be done approximately for the unique parity orbits where one / value
dominates. Noting that only one of the two terms in J + j + J j+ gives a
nonvanishing result, we get
where the symbols K, K +1 in the j matrix element are a shorthand for the two
Nilsson wave functions and the effects of pairing are included in the factor
(C/jUJj + V^V^). The pairing factor has the general effect of reducing the
Coriolis matrix elements since its maximum value is unity. The reduction is
least for orbits in similar positions relative to the Fermi surface (then (7, = U2
and Vl = V2, hence (£/,£/2 + VjV2) = U,2 + V,2 = 1). It is least for orbits laying on
far opposite sides (then £/; - V2 and U2» Vv so (U^U2 + V^V2) - (U^ + U2V2~)
and one factor in each term is small). For diagonal Coriolis matrix elements,
the factor is obviously unity.
For a single j shell, the j matrix element is
This is approximately correct for unique parity orbits. For an arbitrary Nilsson
wave function, terms like this occur for each/, so that, in general,
where the CK and C *+1 coefficients are those describing the specific Nilsson
wave functions. An interesting limiting case occurs for /, / » K, if / is
approximately a good quantum number. This applies for low # orbits from the
high-spin unique parity states or for high spin, low K states generally if ft is
small (little configuration mixing). Then Eqs. 8.4 and 8.6 give
282 Collectivity, Phase Transitions, Deformation
This is a general upper limit on the strength of the Coriolis interaction. For
typical inertial parameters (say 15 keV for rare earth nuclei) and say,
J,;' =11/2, this attains = 400 keV! Typical spacings between Nilsson orbits are
« 150 keV. Coriolis mixing is not necessarily a minor perturbation!
Since the Coriolis matrix elements change K by AK = ±1, it is generally a
nondiagonal interaction. However, it has an important diagonal matrix ele-
ment that contributes to certain rotational energies. For K = 1/2, the symmet-
rization of the wave function gives rise to terms with K = ±1/2, allowing a
diagonal AK= 1 contribution to the energies from the cross terms. Substituting
Eqs 8.4 and 8.6 in Eq. 7.14, we get for the rotational energies for K -1/2 bands
including Coriolis mixing
The two phase factors in Eqs. 8.8 and 8.9 come from the symmetrization of the
wave functions (e.g., Eq. 6.10). The phase factor in Eq. 8.8, (-1)J+1/2, means that
the contribution to the rotational energies from the Coriolis interaction alter-
nates in sign with /.
It is clear from Eq. 8.9 that a can be either positive or negative. The
behavior of E(J) in Eq. 8.8 as a function of a is shown in Fig. 8.4. If a < 0, states
with spins 3/2,7/2,11/2... are lowered in energy while the alternate spin states
are raised. For a > 0, the opposite situation occurs. On account of the factor
(/ +1/2) the effect grows with spin. On account of the factor (/ +1/2) the effect,
on average, increases for heavy nuclei (jm <* W). For I a I =1, Eq. 8.8 shows
that the levels occur in degenerate pairs. If a = -1, the J - 3/2 state coincides
with the J = 1/2 level. Similarly, the (5/2,7/2) and (9/2,11/2) pairs are degener-
ate. Ifa = +l, the degenerate pairs are (3/2,5/2),(7/2,9/2),.... For |a| >l,the
level order within a rotational band is no longer monotonic in spin. Clearly, the
typical rotational spacings can be so severely perturbed as to obscure the
normal J(J +1) spacings and even the ordering of different spin states. As we
shall see, these effects propagate via nondiagonal Coriolis mixing, and affect
many bands with K * 1/2. We can now understand one of the anomalous
rotational spacings and sequences in Fig. 7.8, specifically that for the l/2~ [521]
band. This is a K = 1/2 band with its decoupling parameter close to unity.
For an arbitrary Nilsson wave function, many terms can appear in Eq. 8.9
for a. Frequently these terms (each carrying a phase) largely cancel and the
resultant a values are rather small, typically less than unity. However, very
large a values can be obtained if the wave function is dominated by few terms
with high; values. The classic example of this is the unique parity orbits for
Nilsson Model: Applications and Refinements 283
Fig. 8.4. Dependence of rotational energies (7) on the decouph'ng parameter a. The dashed lines
are for | a I = 1 (based on Preston, 1975).
which(j}~N+1/2 (e.g.,; = 13/2 in theN = 6shell), and the wave functions are
nearly pure in/. For these special orbits, Eq. 8.9 gives a = (-1)N(N + 1). For
example, a ~ 1 for the l/2+[660] orbit. This enormous value so perturbs the
normal spacing that the 13/2+ level is among the lowest-lying levels in the
rotational band. For h2/2I = 15 keV, the decoupling term is ~ -750 keV! In
general, the sign of a is always such that the J = N +1/2 level (e.g., J = 13/2 for
the ijM shell) is lowered.
The reader may recall an apparent paradox in the order and spacing of
rotational energies in odd mass nuclei that was discussed in Chapter 7. The
simplest view led to the notion that rotational bands should be "upside down."
We showed that the "normal" order was regained when the precession of/
around the symmetry axis was considered. We also pointed out that in some
cases an "upside down" pattern does in fact occur. We have just encountered
that case where large Coriolis effects in unique parity orbits upset the mono-
tonic order of rotational energies with /. Having gone to great lengths to
explain away this paradox in Chapter 7, why does it now appear in the data? In
284 Collectivity, Phase Transitions, Deformation
other words, what happened to the precession argument? As we shall see later,
the physical difference here is that K is no longer a good quantum number
(±1/2 values are admixed), so the precession need not be about the symmetry
axis: the Coriolis force, by mixing K values, forces the angular momentum
vector to switch back and forth (precess) about the rotation axis instead. Thus,
j and R are nearly parallel or antiparallel, so now 7=13/2 does correspond to
| / ? | = 0 , / = 9to R | = 2, and so on.
One often reads that a is called the decoupling parameter because it repre-
sents a decoupling of the rotational and single-particle motion. It is now easy
to see why this name is appropriate. For a = +1, the energy differences of the
1/2, 5/2, 9/2,... states are exactly the same as those between the 0+, 2+, 4+,...
states of the even-even core: the odd particle acts like a spectator to the
rotation. Moreover, it is clear from the comment just made that changes in
core rotation have little effect on the orientation of J = (R +j) when a is large
(i.e., when | ;' | is large, Klj is small, so sin0= Klj is small, and J is nearly aligned
along the rotation axis). Extrapolating to very large a values in Fig. 8.4 shows
that the J = 13/2 state will lie lowest, followed by the 9/2 level (13/2 - 2), and
then by the 5/2 level (13/2-4). The alternate spin levels are much higher. Thus
the rotational energies of alternate / values (with \J-j\ even) are nearly para-
bolic in | J-j\ where /is the dominant; of the unique parity orbit. Sequential
states differ mainly in R: the rotational motion is effectively decoupled from
that of the odd particle. We shall encounter a related but even more dramatic
effect later as a consequence of the nondiagonal Coriolis interaction.
Turning now to these nondiagonal Coriolis effects, there are two significant
observable effects. One concerns rotational energies and stems from a propa-
gation of the highly perturbed level order in the K = 1/2 band to higher K bands
via successive AK = 1 Coriolis mixings. Precisely because of the large decou-
pling parameters, this is most important for unique parity states and, as we
shall see, accounts for their importance in high-spin studies where the (7 + 1/2)
factor in Eq. 8.8 becomes crucial. The other effect occurs in "normal"
rotational bands (especially in their impact on single nucleon transfer cross
sections). We shall discuss this first.
Consider the admixture of two bands as shown in Fig. 8.5. Recalling our
discussion in Chapter 1 of two-state mixing and the fact that the Coriolis
mixing increases with /, the perturbed energies will behave as illustrated. It is
easy to show that, to first order, the effect of the Coriolis interaction is to
decrease the effective inertial parameter, h2/2I, for the lower band and to
increase it for the higher band.
Equation 1.6 gives the energy shifts of the two interacting states relative to
the unperturbed spacing for a given spin /
where the last step assumes small mixing (Eq. 1.12). Thus AE^ °c V1^.
Isolating the spin dependence Eq. 8.4 gives
Nilsson Model: Applications and Refinements 285
Fig. 8.5. Illustration of the changes in effective rotational parameters, h2/2I, resulting from two-
state Coriolis mixing.
which is just the rotational energy expression (J, K dependent parts). The
Coriolis interaction merely alters the effective rotational spacings. The lower
band is compressed, the upper one expanded. This simple result breaks down
for very large or multistate mixing, but even in these cases gives a useful
framework.
In general, there can be many low-lying Nilsson bands with assorted K
values; thus a realistic Coriolis mixing calculation will be a multistate diag-
onalization. The admixed wave functions can be written
change by one unit, but their sum remains constant (An = -AA). (The unique
parity case is, of course, in this class.) Examples from the table are the matrix
elements between the 3/2-[512] and l/2-[521] orbits or between the 7/2-[514]
and 9/2~[505] orbits. Such allowed j matrix elements are typically on the order
of NI2 - M3. All of the others are nonallowed matrix elements and are
typically < 1. Examples of these are the matrix elements between 7/2~[514] and
5/2-[512] or l/2-[510] and 3/2- [512].
It is important to note that Coriolis mixing between orbits in the latter class
is not always negligible, especially if they have the same nz values. As we have
seen, the Nilsson diagram separates approximately according to nz values for
large deformations. Therefore, it can often occur that two orbits with identical
nz values differing in K by &.K = ±1 lay very close to each other. Their Coriolis
mixing can be large even with a small matrix element. A classic example of this
occurs for the l/2-[510] and 3/2~[512] orbits in ™-™W. We shall discuss this
case in some detail momentarily.
Having dealt with typical values of the j± matrix elements, it is useful to
develop a feeling for the absolute magnitudes of the full Coriolis matrix
elements in Eq. 8.4. For well-deformed nuclei, h2l2I» £2^/6 where E2\ is
given by a neighboring even-even nucleus. For rare earth nuclei, fi2!2I~ 14-18
keV, while for the actinides, fi 2/2I» 7 keV. The matrix elements of J± given by
the square root factor are typically 2-3 for moderate spin states, although they
can become very large for high spins. Finally, as noted, the pairing factor
becomes very small for high-lying orbits on opposite sides of the Fermi surface,
while for orbits near the Fermi surface, or for those on the same side, this
factor is typically between 0.7 and 1.0. Thus, typical nonunique parity allowed
Coriolis matrix elements in the rare earth region are roughly
VCor= (16)(3)(1)(0.8) <= 40 keV. This estimate is only accurate to a factor of
2-3. Nonallowed Coriolis matrix elements will, of course, be less.
For unique parity orbits, the j matrix elements are - N. In addition, the
observed states are typically of rather high /, since single nucleon transfer
reactions preferentially populate the / = N + 1/2 states for which C. ~ 1 and
heavy ion reactions tend to feed the high-spin unique parity levels. Taking
V(./-.KV./+Ar+l) = N, the Coriolis matrix elements linking unique parity
orbits can be extremely large, typically reaching V Cor (unique parity) ~
16(6)(6)(0.8) <= 400 keV. Such matrix elements mixing states often only a
couple of hundred keV apart have enormous structural effects.
From extensive experience with Coriolis mixing calculations, it has been
found that the actual empirical matrix elements are generally about 20-50%
lower than these theoretical estimates. This conclusion emerges from com-
parisons of extensive data on level energies and single nucleon transfer cross
sections in many deformed nuclei. We will not detail this evidence here, but
one example of it is trivially evident in Fig. 8.1, which shows that the 7/2~ states
of the 7/2-[514] and 5/2~[512] bands in 183W are separated by only 70 keV. If we
consider this an isolated two-state system (an approximation good enough for
the present purposes although not for detailed calculations) and recall that
such states can never be closer than twice their mixing matrix element, then the
288 Collectivity, Phase Transitions, Deformation
Coriolis matrix element between these two states must be < 35 ke V. Since both
bands are hole excitations, the pairing factor is near unity. Using h2/2I = 18 keV
(E2f (184W) = 111 keV) and the j matrix element from Table 8.3 gives a
predicted Coriolis matrix element of 55 keV. The maximum matrix element
allowed empirically is only 65% of this. Other similar examples abound.
Despite this attenuation, Coriolis mixing effects, especially among unique
parity orbits, represent a substantial perturbation to the rotational picture and
can seldom be ignored. One final point to emphasize before considering some
actual calculations is that in weakly deformed and transitional nuclei, Coriolis
matrix elements are far larger than in well-deformed nuclei because of the
smaller moments of inertia. Matrix elements between unique parity orbits
may reach an MeV or more, and under certain circumstances may even lead to
a new coupling scheme, the so-called rotation aligned scheme we shall discuss
later.
As an example of multistate Coriolis mixing, let us consider the level
scheme of 183W shown in Fig. 8.1. In principle, a full calculation cannot neglect
the unseen bands that occur at higher energies, but in practice, one assumes
that their effects are small (at least for nonunique parity orbits). We will see
one way to estimate whether such an assumption is grossly violated. Under
this assumption, the Coriolis mixing among the negative parity bands involves
diagonalizing matrices of varying size, 2 x 2 for 7 = 1/2, up to 6 x 6 for / = 11/2.
For the positive (unique) parity states, the strength of the Coriolis mixing
precludes safely ignoring unseen bands and therefore one usually carries out a
full 7 x 7 diagonalization.
We consider first the negative parity states. We note the empirical result
(from ratios of (d, p) to (d, t) cross sections) that the Fermi surface is slightly
below the 1/2~[510] orbit. For simplicity we ignore the 9/2[505] band. (In any
case, it can only affect/ 9/2 states.) From our earlier discussion, we anticipate
that the principle mixing effects will occur between the l/2~[521]-3/2~[512] and
5/2-[512]-7/2-[503] pairs. That this is not quite true highlights the other factors
that must be taken into account in practical situations. Although the matrix
element connecting the l/2-[510] and 3/2~[512] orbits is rather small (=1), they
lay so close to each other that the mixing is substantial. Likewise, the
"forbidden" matrix element between the 5/2-[512] and the 7/2~[514] orbits
(=1.1), strongly admixes these close-lying bands. In contrast, despite the large
j matrix element between the 5/2~[512] hole orbit and the particle excitation
7/2~[503], the pairing factor substantially reduces the overall matrix element.
The one exception to the simple rule given above for estimating Coriolis
matrix elements among these bands occurs for the 1/2~[510] and 1/2~[521] pair:
the supposedly forbidden j matrix element has a value = 2.5. Although these
bands are nearly an MeV apart, the coupling between them is nonnegligible.
Thus, the principle Coriolis admixtures will be between the 1/2~[510] and
3/2-[512] bands, the 5/2~[512] and 7/2-[514] bands, and the l/2-[510] and
l/2-[521] bands. Second-order mixtures of, say, the l/2-[510] into the 5/2-[512]
band, will be very small. Thus, a rather good simulation of the full diagonali-
zation should be obtainable by considering sequential two-state mixing of the
Nilsson Model: Applications and Refinements 289
Table 8.4. Calculated mixing amplitudes (a. x 100) for the 1/2' states inm 185W
preceding three pairs. As an example, consider the 7/2* states with a Coriolis
attenuation factor 0.7. With H2I2I = 18 keV and a pairing factor of 0.9, we
obtain =41 keV for the full l/2-[510]-3/2-[512] Coriolis matrix element. The
final spacing of the 7/2 l/2-[510] and 7/2 3/2-[512] states is 205 keV. Working
backwards in Fig. 1.7, we see that R must be rather large and therefore the
energy shift induced in each state by the mixing is a small fraction (=0.05) of
their unperturbed spacing. Reducing the full separation by 10% to estimate
the unperturbed splitting gives R ~ 3.8. Another application of Fig. 1.7 or Eq.
1.8 gives the admixed wave functions as y ("7/2 l/2-[510]") =
(0.97) 7/21/2-[510] + (0.24) 7/2 3/2-[512] and the orthogonal combination.
For the 7/2 1/2- [510J-7/2 l/2-[521] mixing we take a pairing factor of 0.6,
giving the full Coriolis matrix element of = 88 keV. AE^ = 1058 keV, so we can
use the final spacings to obtain R »12. The admixed wave functions are
y ("7/21/2-[510]") = (0.99) 7/21/2-[510] + (0.08) 7/21/2-[521] and the orthogo-
nal combination. Finally, for 7/2 3/2-[512] and 7/2 l/2-[521], the pairing factor
is ~ 0.5. Calculations again give R ~ 12, and final wave functions of
V/-("7/2 3/2-[512]") = (0.99) 7/2 3/2~[512] + (0.08) 7/21/2[521], and the orthogo-
nal combination. In all three cases the signs of the amplitudes are arbitrary.
We can test these estimates by reference to the detailed wave functions
resulting from a full diagonalization given in Table 8.4. The three admixtures
just calculated are 1/2- [510]-3/2-[512] = 0.24, l/2-[510]-l/2-[521] = 0.08, and
3/2-[512]-l/2-[521] = 0.08. The exact calculations give 0.26,0.05 and 0.08!
We can also estimate the energy shifts. Using the same R values we get the
following results (in keV):
7/2 l/2-[510]: -12.0(3/2~[512]) - 7.2(l/2-[521]) = -19.2 keV
7/2 3/2-[512]: +12.0(l/2-[510]) - 5.8(l/2-[521]) = + 6.2 keV
7/21/2-[521]: + 7.2(l/2-[510]) + 5.8(3/2-[512]) = +13.0 keV
290 Collectivity, Phase Transitions, Deformation
where the orbits in parenthesis give the mixing partner that induced each shift.
Again, these estimates are close to the results of an exact calculation.
With these shifts, and similar ones for other / values, the 1/2"[510] band is
compressed and the 1/2~[521] and 3/2~[512] bands are expanded, reflecting the
derivation in Eq. 8.10 that, to first order, Coriolis induced energy shifts can be
absorbed into changes in ft2/27. Indeed, one clue to the presence of Coriolis
effects in empirical level schemes is unequal empirical h2/2I values (i.e., after
mixing), with larger magnitudes for the higher-lying (expanded) bands and
smaller values for the lower (compressed) states. A measure of the adequacy
of a calculation is whether the input (unperturbed) ft 2/27 values are substan-
tially closer: they should be if the deformation is the same for all excited states
and small microscopic "blocking" effects are neglected.
where the a.'s are the Coriolis mixing amplitudes and the sum is over the
admixed bands. Note that the sum is coherent, thus magnifying the effects.
Simple manipulations also show that the total cross section is conserved for
each spin /: that which is lost by some states must be gained by others.
Before considering the example of 183W in detail, it is useful to emphasize
how small mixing amplitudes can have significant effects. For simplicity,
assume a two-state mixing of bands with identical C coefficients and pairing
factors for some spin J. Then, if the mixing amplitude of each band in the other
is 0.22 (meaning that the amplitude for the "parent" state is still 0.975), this
gives a 50% increase in the cross section of one state and a 50% decrease in the
other [o^ (1.22)2, a2<x (0.78)2]. The two cross sections that would have been
equal without mixing now differ by a factor of three!
Another feature is evident from Eq. 8.12. If two admixed states have very
different unperturbed Ci values for some J, the state with the larger C. value
will be relatively unaffected while that with the smaller may be drastically
altered. Indeed, much of the resultant cross section may easily come from the
small admixture rather than from the parent orbit itself. To be specific,
suppose the two bands have equal pairing factors, that O =0.2 and C2 = 0.8,
Nilsson Model: Applications and Refinements 291
and that the mutual mixing amplitudes are ±0.22. Then, assuming that the
phases are such that the cross section for the state of spin J in band 1 is
increased, the ratio of perturbed to unperturbed cross sections is =3.5 for band
1 [(0.2 + 0.22(0.8)]2 and =0.9 for band 2 [(0.8 - 0.22(0.2)]2. This is another
example (bandmixing in even nuclei was the first) of how relatively small
mixing interactions and amplitudes can lead to drastic effects on certain
observables, especially when one of the unperturbed transition rates is small
or forbidden. If the phases were reversed (which would happen if the unper-
turbed positions of the two bands were exchanged), the same analysis shows
that despite the small mixing, the cross section for band 1 would essentially
vanish while that for band 2 would increase only by about 6%. Finally, if one
C. coefficient is nearly zero, the cross section will come only from the mixing.
It will therefore be independent of the signs of the mixing amplitudes and will
always be increased by the mixing.
Simple application of Eq. 8.12 to the mixing amplitudes such as those given
in Table 8.4 for the 7/2 - states of the negative parity bands in 183W gives the
cross sections labeled "perturbed" in Table 8.1. The point of this section is
highlighted by the enormous differences between perturbed and unperturbed
Fig. 8.6. Systematics of experimental, unmixed, and Coriolis coupled (d, p) and (d, 1) cross
sections in W isotopes.
292 Collectivity, Phase Transitions, Deformation
cross sections even in cases where the mixing amplitudes of Table 8.4 are small.
For example, because of its small value of Cm, nearly all the cross section to the
1/2 l/2-[510] state stems from weak mixing with the 1/2 l/2-[521] state. The
cross sections for the 3/2~ levels of the 1/2~[510] and 3/2~[512] bands are
significantly shifted by the mixing. The same applies to the / = 5/2 and 7/2
states of these bands and to the 7/2 states of the 5/2-[512] and 7/2-[514] bands.
Table 8.2 shows similar Coriolis mixing results in 185W.
Figure 8.6 shows the systematics of some empirical and Coriolis calculated
cross sections across the odd mass W isotopes. It includes a comparison with
the unmixed cross sections. The latter are constant except for small, smooth
changes in the pairing factor P2. in Eq. 8.1. The figure highlights the changes
in single nucleon transfer cross sections brought about by the Coriolis interac-
tion as well as the dramatic shifts that can occur from one isotope to another.
This is particularly evident for the l/2-[510] and 3/2~[512] bands in 183W and
185
W. An understanding of this is obvious from our discussion of two-state
mixing in Chapter 1. As we have stated before, these two bands have
interchanged positions (energies) in these two nuclei. In 183W, the 1/2~[510]
band is the lower, ground state orbital, while the 3/2~[512] band is the first
excited intrinsic excitation. In 185W, the 3/2~[512] orbital forms the ground state
and the l/2-[510] band is close, lying just above it. Therefore, the signs of the
mixing amplitudes are inverted between 185W and 183W. In the familiar termi-
nology of first-order perturbation theory, the sign of the energy denominator
has changed from one nucleus to the other. Therefore, for those states where
the two unperturbed Nilsson wave functions have comparable C coefficients,
cross sections that were increased in 183W will be decreased in ™5W and vice
versa. For cases where the C. coefficient in one is negligible, the cross section
to that state will increase relative to the unmixed case in both nuclei. Another
example of inversion concerns the 7/2~[514] and 5/2~[512] bands, in which the
7/2~levels interchange positions between 181W and 183W. Figure 8.6 shows the
dramatic effect on the weaker cross section.
Before turning to the positive parity levels, it is worth re-emphasizing the
extremely large effects involved here. Empirical fingerprint patterns auto-
matically incorporate the effects of Coriolis mixing and can differ from those
predicted by the Nilsson model by sufficiently large quantities as to completely
obscure the identification of the bands if Coriolis mixing is not taken into
account. Moreover, the mistakes that one would make would not even
necessarily be the same in neighboring nuclei, and the systematics of the
Nilsson orbits deduced could be completely wrong.
orbits stemming from the same/ shell (although second-order (A/C = 2) mixing
is significant).
Before discussing practical calculations, let us take a schematic model.
Assume the Fermi surface lies below the whole group of unique parity orbits in
some nucleus. The order of their excitation energies is K = 1/2,3/2 (Af+1/2).
Fig. 8.7. Illustration of how the irregular rotational spacings in strongly decoupled K =1/2 bands
can propagate to K & 1/2 bands via Coriolis mixing.
294 Collectivity, Phase Transitions, Deformation
Fig. 8.8. (3He, a) cross sections for 183W (based on Kleinheinz, 1973).
Each mixes with the K + 1 and K - 1 member of the series. We thus have a
situation analogous to one discussed in Chapter 1, in which equally spaced
states each mix with adjacent levels. One general result is that lowest band will
be pushed much lower. Moreover, given the increase of the Coriolis matrix
elements with J, it will be severely compressed, and its wave functions will be
a complex mixture of several components with all wave function components
in phase.
The principle difference between this schematic situation and the real one
arises because of the large decoupling parameter for the K = 1/2 band. For
N = 6, this has the effect of greatly lowering the 13/2+, 9/2+, 5/2*, l/2+ states and
raising the ll/2 + , 7/2+, 3/2 + ,... states of that band prior to mixing. Consider now
the effect of Coriolis mixing on the nearby K = 3/2 band. The situation is
illustrated in Fig. 8.7. The reordering of energies in the K = 1/2 band because
of the large diagonal Coriolis effect (decoupling parameter) causes the unper-
turbed spacings between the 3/2, 7/2, and 11/2 states of the two bands to be
much larger than between the 5/2,9/2, and 13/2 states. Therefore, in the lower-
lying K = 3/2 band, the 5/2, 9/2, ..., group is shifted down substantially more
than the 3/2,7/2,..., group. The perturbed energies of the K = 3/2 band take on
an alternating pattern as well, relative to a pure .7(7+1) rotational spacing, and
appear as iflhe. K - 3/2 band had a decoupling parameter of the same sign and
slightly smaller magnitude than the K = 1/2 band. When the K = 3/2 band in
Nilsson Model: Applications and Refinements 295
turn mixes with the K = 5/2 band, this "signature" is passed on in a somewhat
reduced form. In effect, the Coriolis mixing "propagates" the decoupling
parameter throughout the entire sequence of unique parity orbitals.
If the Fermi surface is below the K = 1/2 orbit, the effect is reversed. The
K = 1/2 band lies below the K = 3/2 band and therefore the closest lying pairs
are the 3/2, 7/2, and 11/2 states. Also the propagation is severely damped by
the pairing factor as one goes from hole states to particle slates. In the W
isotopes that we have been considering, the Fermi surface is near the 9/2" [624]
and ll/2+[615] orbits, and the effect of the K = 1/2 band is negligible. (This is
linked with a point we will make shortly, that strong diagonal Coriolis matrix
elements are most effective in inducing a rotation aligned coupling scheme
when the Fermi surface lies near the low K orbits.) In W, the primary
observable mixing among the unique parity orbits should be in the K = 5/2,
112,9/2, and 11/2 orbits. As contrasted with the normal parity states, here the
matrix elements and spacings are comparable and a two-state mixing calcula-
tion is hopelessly crude. The results for 183W of an explicit calculation of the
single nucleon transfer strengths (C 13/2 coefficients) to 13/2' states is shown in
the top two panels of Fig. 8.8. Since the j matrix elements arise almost solely
from they = 13/2 term in Eq. 8.6, and since the C. 13/2 coefficients all have the
same sign, the phases of the resulting wave functions are such that in any two-
296 Collectivity, Ph ase Transitions, Deformation
state mixing of these orbits, the cross section to the lower state is increased
while that to the higher is decreased. This persists in the multistate extension,
and the net effect is to transfer cross section from the higher-lying bands to the
lower ones. This is the point alluded to in the schematic model at the beginning
of this discussion. We now compare this calculation with the empirical situ-
ation.
In 183W, the lowest-lying unique parity orbit is ll/2+[615]. The 9/2+[624] and
7/2+[633] orbits are hole excitations. An ideal reaction to study these unique
parity orbits is the (3He, a) reaction, which preferentially excites higher-spin
hole levels. In the rare earth region it can almost be used as a "/ = 13/2+ meter."
Typical (3He, a) spectra, for 181-185W, are shown in Fig. 8.9. Comparison with
the (d, t) reaction in Fig. 8.2 vividly illustrates the selectivity. In the absence of
Coriolis mixing, the (3He, a) reaction to 13/2+ states in 183W should look like
that shown schematically at the top of Fig. 8.8, in which there are five nearly
equally strong peaks. (That for the K = ll/2+ is weaker than the others due to
the smaller V2 factor and the peak from the 13/2+[606] orbit is absent since
V ~ 0.) In contrast, the data show only four peaks but with the same total cross
section expected for the six unperturbed states. This illustrates both the
shifting and the descent of strength just discussed. (The empirical (3He, a)
spectrum (Fig. 8.9) for 181W shows this effect even more; only two peaks
consume nearly all the 13/2+ strength.) The bottom-most panel of Fig. 8.8
summarizes the empirical C =13/2 coefficients in bar graph form for 183W.
The second panel of Fig. 8.8 includes Coriolis mixing, and is somewhat
better than the unmixed calculations shown in the top panel. Further improve-
ment requires, as we shall now see, the introduction of hexadecapole deforma-
tions. The study of such shape components offers us an ideal situation to apply
the same kind of intuitive approach we used for the Nilsson model itself.
with the largest radii are the unique parity orbits since they stem from the next
higher oscillator shell. From the relation sin0 ~ Klj, we see that 6 = 45°
corresponds to Klj ~ 0.7. For the im neutron orbit in the rare earth region, this
gives K ~ 9/2. Thus, the largest contributions to an increasing positive
hexadecapole deformation in this region occurs when the 9/2+ [624] orbit is
filling near the Fermi surface. Simple counting in the Nilsson scheme shows
that this occurs around A ~ 180.
At the other extreme, analogous reasoning shows that negative e4 values are
favored when very low or very high K i13/2 orbits are filling. The former occur
near the beginning of the deformed region at neutron number N ~ 92. The
latter occur near the end of the shell and serve to reduce the positive e4 values
for A > 180. The expected e4 systematics should therefore consist of large
negative values at the beginning of the deformed region that decrease in
absolute value with mass, cross zero, turn positive, and maximize around the
W isotopes, followed by a rather rapid decrease towards zero as the 208Pb
closed shell is approached. This is exactly the systematics observed empiri-
cally, as shown in Fig. 8.10.
[A technical point is worth mentioning here. In this discussion, we have
used e4 as the deformation parameter. Figure 8.10 is expressed in terms of j84.
In Nilsson's original paper, the principle discussion was carried out in terms of
deformations /32 and $,. The disadvantage of these parameters when discussing
AN = 2 mixing and hexadecapole deformations is that, even for fa = 0, there
will be finite AN = 2 mixing. In contrast, the e representation discussed in an
appendix to Nilsson's original article, was designed so that A/V - 2 mixing
vanishes when £4 = 0. The relation between e2, £4 and j32, j94 is complex and
coupled. Either set of deformation parameters may be converted into the
other by the use of Fig. 9 of the article by S. G. Nilsson et al (1969): However,
one should note that there is a mistake in this figure and that its proper use
requires the reversal in the sign of e4. Very crudely, e4 —/?4.]
We can go one step further. Since the orbit inclination 6 changes slowly for
low K, there will be more low-angled orbits (9 < 45°) than orbits near 45°.
Therefore, negative e4 deformations should predominate and the "crossing
point" to positive values should occur past midshell. This is also seen in Fig.
8.10.
The principle effect of hexadecapole deformations on the Nilsson wave
functions is to admix components with AN = ±2. Thus, the N = 6 im Nilsson
orbits will now contain some components from the N = 4, 8 shells, and the
N = 5 normal parity orbits will contain contributions from TV = 3,7. Normally,
this AN = 2 mixing is miniscule because of the large energy separation of
oscillator shells. However, inspection of the characteristic form of the Nilsson
diagram (downsloping early, upsloping late), shows that there are a few
isolated regions where steeply downsloping unique parity orbits from one shell
(e.g., i^) intersect upsloping orbits from the next lower shell (here, N = 4). If
the nuclei in such regions have large e4, then substantial AN = 2 mixing can
occur. An example of such interacting orbits was discussed earlier in Chapter
7 in the context of a comment on the relative purity of the nz quantum number
at large deformations.
Thus far we have discussed the origin of hexadecapole deformations, their
systematics, and their relatively minor effects on Nilsson wave functions. It
remains to discuss their enormous impact on Coriolis mixing and single
nucleon transfer cross sections. This impact arises mostly from the effect of e4
on Nilsson energies of orbits that can Coriolis mix. It is easy to see what the
main effects will be. Consider, for example, a large positive e4 and the i13/2 or-
bits. It is obvious that both equatorial (K = 1/2, 3/2,5/2) and polar (K = 13/2
and perhaps 11/2) orbits will be, on average, further from the nuclear matter
than for e4 = 0, and therefore their energies will increase. The mid-A" orbits
(K = 7/2, 9/2) will be closer to the nuclear matter and their energies will
decrease. Hence the overall effect will be a compression of the energy
separations from K = 1/2 to 7/2 or 9/2. Moreover, this compression will become
more extreme as e4 increases. In fact one can imagine sufficiently large e4
values that some of these K orbits may actually cross and interchange their
relative positions. Figure 8.11 shows an explicit calculation of the i13/2 energies
for fixed e2 as a function of e4. All these features appear. There is a compres-
sion, and even a crossing, near e4 ~ 0.1. The envelop of the energies can easily
be compressed by a factor of two and, therefore, the already large Coriolis
mixing among the unique parity orbits will increase still further (the Coriolis
matrix elements themselves will not substantially change).
Recall from our discussion of cross sections to 13+/2 states in the W isotopes
that Coriolis mixing calculations produce some improvement in the predic-
tions, but that significant discrepancies remain. The mixing casts some cross
section from higher lying levels into the lower ones. With the increased
Coriolis mixing that now occurs with a large positive e,, this effect will be
exaggerated, as shown in the third panel in Fig. 8.8, where we see much better
agreement with the empirical cross sections.
Nilsson Model: Applications and Refinements 299
One of the beauties of the Nilsson model is its easy extendibility. We see
here an excellent example where the basic model predictions are in strong
disagreement with the data but where simple and physically reasonable refine-
ments easily remove most of the discrepancies, and thereby show both the
usefulness of the model as a starting point and also the absolutely crucial need
to incorporate certain of these extensions. The particular case we have been
considering is historically interesting as well: the large changes in im energies,
Coriolis mixing, and (3He, a) cross sections as e4 varies from 0 to 0.06 provided
the first definitive evidence for large hexadecapole deformations in the odd W
isotopes. Another interesting point is that while e4 values of zero and 0.06 can
be easily distinguished in this way, the approximate constancy of the envelope
of K = 1/2 to 7/2 orbits from e4 = 0.06-0.16 precludes a further refinement in the
actual £4 values.
There is one other consequence of large hexadecapole deformations that
should be mentioned. We have been discussing permanent or static hexadeca-
pole shape components. However, it is also possible that the nuclear potential
energy surface may be "soft" in £4, and that this will lead to hexadecapole
vibrations, just as softness in p and 7 leads to ft and 7 vibrations. It is not
surprising that the heavy even-even rare earth nuclei, especially the Os
isotopes, display rather low-lying (~1 MeV) K = 4 bands that have been
interpreted in terms of hexadecapole vibrations by Baker and co-workers. Of
300 Collectivity, Phase Transitions, Deformation
Fig. 8.12. Diagram for strong coupling and rotation aligned or decoupled level schemes (based on
Stephens, 1975).
This is the Nilsson energy of the odd particle plus the total rotational energy.
At the top of Fig. 8.12 K is large and the K2 term dominates the VCor term.
Since the coefficient of K2 is linear in 8 and since the inertial parameter, h 2/2I,
decreases with increasing deformation, this is a situation that is valid for large
deformation and/or large K, The dominance of the K2 term implies that K is a
good quantum number (the K mixing terms are relatively small) and one has
the so-called deformation aligned or strong coupling scheme we have been dis-
cussing. Coriolis mixing effects are a small (but important) perturbation that
causes the angle of the vector; to "wobble" slightly as it precesses about the z
axis.
However, there are situations in which this coupling scheme does not occur.
An obvious one is for high spins for which \Cor <* 7; when this term dominates
the K2 term, the solutions must be approximate eigenfunctions of VCop that
correspond to a new coupling scheme in which K specifically is not a good
quantum number. One then has the situation illustrated in the lower part of
Fig. 8.12, where the particle angular momentum vector, j, precesses about an
axis perpendicular to the symmetry axis (i.e., about the rotation axis). Clearly,
A" will vary significantly and include negative values. It is now the alignment
along the rotation axis, commonly called a, that is the good projection quan-
tum number. This coupling scheme will be realized when
that is, especially for low K values. This is physically plausible since j already
points nearly along the rotation axis. Clearly, if K is large, an enormous
Coriolis interaction (extremely high J values) would be required to enforce
precession about the rotation axis, whereas for low K, such precession can
302 Collectivity, Phase Transitions, Deformation
occur at relatively low spins. Inspection of Eq. 8.14 shows that the rotation
aligned scheme can also be realized for low J if the coefficient of K2 vanishes.
Since ft2/27 ~ 1/5, it is clearly possible to choose a 5 value that satisfies this
cancellation requirement. For the A ~ 130 region, numerical estimates give
<5= 0.17 (/?=0.18). This is a rather moderate deformation and accounts for the
fact that the rotation aligned scheme often manifests itself in transitional,
moderately deformed prolate nuclei early in a shell (where the low K orbits are
filling). For larger deformations, the inertial parameter ft2/27 drops rapidly
while the Nilsson energies further split and the Coriolis effects decrease below
a critical value. In most well-deformed nuclei we see normal (strongly coupled)
rotational behavior.
The energies in the rotation aligned scheme are very easy to visualize. The
total angular momentum points essentially along the rotation axis and is
composed of the particle angular momentum j plus a core rotation R. Thus,
from Eq. 8.13
For high spin unique parity states, low K values and moderate deformations,
JJ » K and / is nearly a good quantum number. Moreover, in the rotation
aligned scheme, I / M a i . Therefore, using Eq. 8.7 and neglecting terms
independent of J, we have
This equation is simply that for a rotor of spin (/-a). But |(/-a)| = \R ,
the core rotational angular momentum! So the energies do not behave like
those of a rotor with spin /, but rather like those of the rotational core.
Moreover, the lowest energies occur for the highest alignments, a. The reader
may recall that when we derived Eq. 7.17, we set the problem up as the solution
to why rotational bands were not upside down, and why for example, the core
angular momenta R were not 0,2,4... for states with J = 13/2 and (9/2,17/2),
(5/2,21/2),..., respectively. The solution involved recognizing the variation of
I R | values that occurs when the particle angular momentum vector precesses
around the symmetry axis. We alluded to the possibility that rotational bands
with core rotational spacings did indeed exist in certain circumstances. We
now see those circumstances—when the precession is no longer about the
symmetry axis but rather about the rotation axis—so that R is nearly a constant
of the motion.
To understand the implications of Eq. 8.15, let us take as an example a
situation of maximum alignment for the rare earth nuclei where the unique
parity orbit has; = 13/2. We take a = 13/2. The energy difference £17/2 - E13/2 is
given by the energy difference between R = 2 and R = 0, that is, by the energy
of the first 2+ state in the even-even core nucleus or 6(ft2/27). This is com-
pletely different from the strongly coupled case in which £17/2- El3/2 =
ft2/2I [17/2(19/2) - 13/2(15/2)] = 64 (ft2/27). This difference is enormous, as is
the energy saving if the rotation aligned scheme is applicable. It was precisely
Nilsson Model: Applications and Refinements 303
Fig. 8.13. Comparison of rotational spacings in Ba and La nuclei. This is the classic example of
decoupled band structure (based on Stephens, 1975).
mate those of a weak coupling model. The other, unfavored states are pushed
considerably higher.
Of course the exact solution for the rotation aligned scheme can be ob-
tained by explicit diagonalization of the Nilsson particle-plus-rotor Hamil-
tonian. An example taken from Stephens is shown in Fig. 8.14, for the case
where the unique parity orbit is the h11/2 and the Fermi surface is below the
K = 1/2 orbit. (Note that in all of the preceding discussion of the im orbits, the
only relevant property of the unique parity orbits was the purity in j, and
therefore nearly identical effects result for any other unique parity orbits
simply by substituting a different/. For example, if the unique parity orbit is
hn/2 rather than im, the lowest spin state will be J = 11/2 and the favored
aligned states will have spins J = 11/2,15/2, 19/2, .... This makes for a very
generally applicable scheme with close correlations from mass region to mass
region.) In Fig. 8.14, the favored aligned states are given by the thick lines, the
others as thin lines. The characteristic feature of the decoupled band emerges
clearly on the prolate side, whereas on the oblate side, the Fermi surface is
near the high K orbits, so the lowest states form a normal strongly coupled
band. For the rotation aligned scheme, the favored aligned energies remain
remarkably close to those of the core energies (which can be seen at /? = 0) even
out to relatively large deformations.
This brief summary of rotational alignment shows that it can be a rather
widely applicable phenomenon, occurring especially in moderately deformed
nuclei, where Coriolis effects are strong and deformation effects still rather
weak, whenever the Fermi surface is near the low K unique parity Nilsson
orbits. The rotation aligned scheme relies on the notion that it is energetically
Fig. 8.14. Behavior of particle-rotor level energies with [3 for unique parity levels including
Coriolis mixing. The Fermi surface is below the entire h ra set of orbits. Note the strongly
coupled pattern (&J = l,J(J+ 1) spacings) on the left and the decoupled pattern (A/= 2, £' mjn for
/ = 11/2, compressed (core) rotational spacings) on the right (Stephens, 1975).
Nilsson Model: Applications and Refinements 305
Fig. 8.15. Empirical proton unique parity (n = +, g9/2) levels of odd mass Rb isotopes. The isotopes
span a strongly coupled (deformed) toward decoupled (weakly deformed) transition. The inset
indicates the proton orbits and the movement of the proton Fermi surface (with decreasing N),
Solid levels are favored states (dashed are unfavored) in the rotational aligned picture (based on
Tabor, 1989).
Fig. 8.16. Plot of K (see text) against A for several energy differences in the Er nuclei (based on
Stephens, 1975).
odd mass nucleus has full rotation alignment (a= 13/2), then Reo»1 independ-
ent of spin.
If the rotational band in the odd mass nucleus is strongly coupled, the larger
spacings will lead to Rto » 1. Empirically, near 157Er Rfo » 1, but a clear
transition toward a strongly coupled limit is observed with increasing neutron
number. This is caused both by an increase in deformation, which reduces the
strength of the Coriolis interaction (VCor <* ft2/27), and by an increase in the
Fermi surface from near the low K unique parity orbits toward the mid-K
orbits. In addition, we note that the transition proceeds much more slowly for
higher spins. The energy spacing E2m- EK/2 remains very close to the rotation
aligned limit. This is simply because, as we have noted several times, the
Coriolis interaction increases with J for high spins states and therefore the
rotation aligned coupling scheme persists longer.
One last point, of some interest in terms of testing this picture of rotation
alignment and favored and unfavored states, concerns the relative role of high-
and low-spin levels. The entire picture described so far assumes a simple
axially symmetric rotational core and its coupling to the odd particle motion.
It entirely neglects any effects due to rotation-vibration coupling, axially
asymmetry, mixing with quasi-particle states, and so on. If one considers the
rather general situation of a fixed number of valence nucleons spanning a
specific set of single-particle states it is clear that, while there are many ways of
constructing low and intermediate spin states from different combinations of
the single-particle angular momenta, there is only one way of constructing the
highest spin level—by aligning all these individual particle angular momenta
along the same direction in space. Therefore, any model for this highest spin
state has the same structure, independent of the assumptions of the model.
For other high-spin (but not the highest-spin) states there will, in general, be
only a few ways of constructing them and different models may present
somewhat different, but mostly likely not very different, predictions. For low-
spin states, however, different models with different interactions may have
entirely different effects on specific subsets of states. Their energies and
structure may differ markedly from one model to another. Although the
beauty of and evidence for rotation aligned behavior is most dramatic in the
high-spin states, perhaps the most sensitive tests of such models occur in the
low-spin, unfavored, antialigned levels. Study of such levels may provide
evidence for other degrees of freedom of some importance.
We should briefly apply our understanding of Coriolis effects to even-even
nuclei. This will provide a simple understanding of the widespread phenom-
ena of backbending. Consider a specific two-quasi-particle state with both
particles in a low K unique parity orbit paired to J* = 0' and with the Fermi
surface below all the unique parity orbits. Neglect for a moment the interac-
tion between the two particles. Since the particles are in the lowest unique
parity orbits, their energies are greatly lowered due to the strong Coriolis
interaction with particles in higher K orbits. This is analogous to the situation
discussed in Chapter 1 of a set of equally spaced states with equal matrix
elements connecting adjacent levels. The lowest level is pushed down and
becomes a collective (strongly admixed) combination of amplitudes. From the
308 Collectivity, Phase Transitions, Deformation
size of the unique parity Coriolis mixing matrix elements, we have seen that
this energy lowering can be substantial. The tendency will be for each particle
to align its angular momentum with the rotation axis. Since the Coriolis
interaction grows with spin, it may well be that at some J value it becomes
energetically favorable to form a state, not from a core angular momentum
Fig. 8.17. Dlustration of the idea of crossing bands and frequencies in a plot of E(J) against J
(lower part) leading to the backbending phenomenon (top). The insets give an example, appropri-
ate to the Ce region, of how one can understand the systematics of proton (n) and neutron (v)
crossing frequencies (sketched in the small boxes) against Z and N (based on Wyss, 1989).
Nilsson Model: Applications and Refinements 309
only), and aptly reflects the origin of the name backbending. The smaller
energy spacing between the J =12 and J = 14 states in the lower part becomes
the sharp backbend in the upper part. The concept of crossing frequency now
replaces that of 7crit.
One can view the reason for the upbend or backbend in two equivalent
ways, both based on the rotational energy expression (fi2/27) R(R + 1). In one,
the R values are effectively lower, as we have described. This is the view in
terms of a new aligned coupling scheme in which R suddenly decreases to take
advantage of the angular momentum gained from the aligned particle. The
other view is phrased directly in terms of Coriolis coupling in that the coupling
causes a substantial lowering of the states of the two-quasi-particle aligned
band. This lowering is larger for higher spins so that, at some spin, they cross
the "normal" levels. (Of course, there can be other origins for backbending
behavior, such as shape changes of nucleus as a whole: we do not discuss these
here.)
In a realistic situation, there will be an interaction between the ground band
and the decoupled band near the critical or crossing frequency. Just below Jctit,
the decoupled band is higher. Its interaction with a ground band will lower the
energies of the latter, effectively increasing 2I/h2, while simultaneously reduc-
ing the transition energies. This causes a slight upbend for J < 7crii: mixing
smooths out the sharply angled ideal pattern. The net effect is to lead to an S
shaped curve where the sharpness of the backbending or upbending depends
on the strength of the interaction between the two bands.
The data for many nuclei have now been accumulated and backbending is a
widespread phenomenon. As expected, it is typically observed when the
Fermi surface is near the low K orbits of the unique parity orbit. It tends to
disappear with increasing mass for a given sequence of isotopes as the Fermi
surface rises toward the higher K orbits.
How this works in practice can be illustrated by the example in Fig. 8.17.
The lower part, discussed earlier, shows a normal plot of £(/) vs. J and depicts
two crossings occurring at different E, J and crossing frequencies (h(o). The
insets show typical Nilsson diagrams for proton and neutron unique parity
orbits and indicate where the Fermi energies (EF) are in this example. We can
now understand the expected systematics of backbending in this region. As Z
increases, E*F (proton Fermi energy) increases. Moreover, the deformation ft
increases as deformation driving orbits are filled. For both reasons, the energy
required to occupy the low K unique parity orbits decreases. Hence, the
proton crossing (labeled n in the inset) occurs at lower energy and angular
momentum and hence lower hco. As N increases, the deformation decreases as
polar orbits are encountered. Hence the proton crossing frequency increases
with jV as the energies of the proton unique parity orbits rise. The neutron
crossing frequencies have an opposite behavior. They decrease with increasing
N (decreasing j3) since the energy separation to the low K orbits decreases, but
increase with Z because the deformation increases. These ideas are sketched
in the figure as crossing frequencies for protons and neutrons. While the
arguments are qualitative, they are borne out both experimentally and thco-
Nilsson Model: Applications and Refinements 311
Fig. 8.18. Example of the odd particle blocking technique for intrinsic excitations in 164 165
Yb
(Riedinger, 1974).
312 Collectivity, Phase Transitions, Deformation
We call *P0 the vacuum state and will see exactly what this means later.
Consider now an arbitrary operator OJ that acts on *F0, giving a new wave
function ¥a
Now, suppose, and this is the key point on which all else depends, that Oaf
happens to satisfy the operator relation
Although we have yet to specify what form OJ must have for Eq. 9.3 to be
obeyed, or what coa is, let us center our attention for the moment on the
implications of this equation. Writing out Eq. 9.3 explicitly and acting on 4*,,,
we have
or, finally
This is basically a definition of the expansion coefficients Xar. Using Eq. 9.5
in Eq. 9.3 gives
or,
This, however, is just a matrix equation that must be satisfied for each r and
can be written
where M is the matrix whose elements are M and X is a column vector with
TS OC
elements Xm.
Thus, Eq. 9.4 and the condition on the OJ, Eq. 9.3, follow if Eq. 9.9 holds.
Alternatively, if Eq. 9.9 is true, then defining the OJ by Eq. 9.5, we find that
Eqs. 9.3 and 9.4 are obeyed.
At this point, it is probably not clear why we have indulged in this process of
piling definition upon definition. The aim was to produce Eq. 9.9. The reason,
and the practical use of all this is as follows.
The basic idea of the TDA and RPA (or higher-order approximations) is
a) to make assumptions about the operators O, defined in Eq. 9.5
b) to then use the definition given in Eq. 9.8 to solve for M (that is for the
elements Ma)
c) to then use these A/w's in Eq. 9.9 to solve for the XJs.
By Eq. 9.5 we then know the detailed structure of the operators Oaf that create
the excitation *Fa! If this seems amazing and magical, good. If it seems abstract
and artificial, be patient.
316 Collectivity, Phase Transitions, Deformation
To remove a little of the magic (but hopefully not the awe), let us consider
two simple assumptions for the O/s and see what results. Suppose the Or's
have the schematic form:
This will hold if A^ = 0 for all r. Therefore, this *F0 has no particle-hole
excitations r. If it had any, Ar could destroy one, giving a nonzero wave
function that did not contain a particle in that orbit. Thus, *P0 must be a closed
shell nucleus. Alternatively phrased, *P0 has no built in correlations. The TDA
therefore corresponds to creating excitations from a very simple, uncorre-
lated, uncollective ground state.
Assumption (ii), the RRA, on the other hand, means that we define OJ in
terms of Eq. 9.5 by
where, for convenience, we have separated those arbitrary operators "Or" that
are particle-hole creation operators (denoted A/) from those that are
particle-hole destruction operators (denoted Ar). Thus, we have
Fig. 9.1. Highly schematic illustration of the difference between the TDA and RPA approxima-
tions to the microscopic structure of collective excitations.
Multiplying by A. on the left, writing out the commutator, and taking the
matrix element of Eq. 9.15 for state *F , gives
Recalling that A.^ and As simply create and destroy particles in orbit s, and
abbreviating wave functions *Fs = A^^ by s, (and similarly for /'), we have
where we also used the fact that the quantities Mrs are just numbers.
Now, the right side of Eq. 9.16 vanishes unless r = z (since f or r * i it contains
the factor (V0\ A .Arf 4*) = {4y0 A. *V), which vanishes by orthogonality since
HMias no particle in orbit;' to be destroyed). For r = /', we have on the right side,
318 Collectivity, Phase Transitions, Deformation
where the Q's are proportional to the transition quadrupole matrix element
from the ground state to a particle-hole state with the particle in orbit i or j.
That is, Q. means (i IQ, ¥„> = (i \ r% | *P0). C is the strength of the interaction.
For an attractive interaction, C < 0. To keep the notation simple, recall that we
have specified each particle-hole excitation by a single subscript. Technically
this is incomplete; each such excitation involves elevating a particle to some
empty orbit and vacating a filled one. As long as no confusion results, we shall
keep to this notation but the reader should keep in mind that each Ae. involves
a pair of orbits and the energy difference between them.
Substituting Eq. 9.18 in Eq. 9.9 and using Eq. 9.17, we have
or
or
or
Microscopic Treatment of Collective Vibrations 319
or, finally, we get the expansion coefficients or amplitudes for the wave
function Ya of Eq. 9.6
TDA
To find the eigenvalues coa, we multiply Eq. 9.19 by Qr and sum over r,
obtaining
TDA
which is the desired result. Note that the wave functions specified by the Xm
and the energy (oa are now written in terms of completely known quantities:
single-particle energies and single-particle transition quadrupole moments.
Given a single-particle model (shell, Nilsson, etc.), they are easily calculable.
Indeed, we shall see shortly that, using these results, we can for example
virtually "derive" y vibrational wave functions by inspection of the Nilsson
diagram without detailed calculations.
We emphasize that in Eq. 9.19, the quantity £<2,^a,is J ust a number, so that
Fig. 9.2. Illustration of the solution to Eq. 9.20TDA. The energies ^correspond to those points
where the vertical curves cross the "1/C" lines. The lowest (rightmost) solution is the collective
one (Ring, 1980).
320 Collectivity, Phase Transitions, Deformation
the coefficients of the vibrational wave functions for the excitation a are
simply proportional to QJ(&er - <»a). This fundamental result is of great
importance.
Recall that the Qr are the quadrupole transition matrix elements connecting
the ground state with a particular (>th) particle-hole excitation. Thus the
amplitude for particle-hole excitation r in the vibrational wave function is
proportional to the matrix element for a quadrupole transition to this
particle-hole excitation and is inversely proportional to the difference be-
tween the particle-hole energy, Aer, and the vibrational energy, a>a.
Particle-hole excitations that require energies close to ct)a and have large
quadrupole matrix elements are strongly favored.
The eigenvalue Eq. 9.20 has several possible solutions that are labeled a.
These correspond to those energies coa such that the right-hand side takes on
the value 1/C. These solutions are illustrated in Fig. 9.2. One solution always
appears significantly lowered in energy. This is the collective vibration.
It is easy to see this if we take a highly simplified example. Suppose there
are N identical particle-hole energies Ae and that the Qr values are also all
equal. Then the lowest solution has the wave function amplitudes
Hence,
which is N times the single-particle B(E2) value. Since this exhausts the total
strength, it follows that no other solution has any strength. Of course, this is an
extreme simplification, but it does demonstrate what is meant by collectivity
and coherence and how they arise from this microscopic formalism.
Finally, the eigenvalue for the collective solution in this degenerate case is
given from Eq. 9.20 by setting all (<aa- Ae) constant, giving
r-i
The vibrational energy is lowered (recall that C < 0 for an attractive residual
interaction), relative to the (common) particle-hole energy by the large amount
Microscopic Treatment of Collective Vibrations 321
NCQr2 This leads to an alternate viewpoint on the structure and origin of the
vibrational wave functions since they are identical to those we obtained in
Chapter 1 for the mixing of a set of degenerate levels. The lowest wave
function (see Eq. 1.13) was a sum of equal amplitudes for all basis states: this
state was lowered by an amount proportional to the number of admixed states
while all of the others were raised in energy.
Before closing this discussion and looking at applications of these ideas, it is
worth recapitulating the key ingredients that lead to the results obtained:
• We discussed a simple but abstract operator formalism leading to the
essential Eqs. 9.3, 9.4, 9.5, and 9.8. Equations 9.3 and 9.4 show the
properties of the excitation a whose structure we are interested in.
Equation 9.5 defines the creation operators OJ and Eq. 9.9, with the
definitions of Eq. 9.8, gives a means for solving for the coefficients Xm
that define the wave function *Fa of Eq. 9.6, and for the energies a>a.
• We made simple choices for the/orm of the various operators Oaf. These
correspond to the TDA (Oaf contains only particle-hole creation opera-
tors) and the RPA (Oaf contains both particle-hole creation and destruc-
tion operators).
• We made a simple choice for the residual interaction V. This step is often
called a schematic model. Doing this allowed us to calculate the A^'s,
hence the OJ's, and hence the structure and energy (fi>o) of the vibra-
tional state YB.
The TDA represents the gross approximation that the ground state is a
pure, undisturbed, closed shell nucleus. This is generally not realistic and the
use of the RPA to incorporate ground state correlations is more common. We
therefore present (without derivation) the eigenvalue results analogous to Eq.
9.20TDA, for the RPA:
or
RPA
This is similar to the TDA expression except for the more complicated
factor multiplying the Q\ Note that, in the degenerate case (all Aerequal) this
gives
or
Using Eq. 9.20TDA and labeling the energies a>a as ft>aTDA or fi>aRPA, we have
322 Collectivity, Phase Transitions, Deformation
if all the Qr are equal. This shows that <oaRPA is always less than coaTDA. The RPA
leads to greater collectivity.
We close this section by noting that the discussion has been phrased in terms
of spherical single-particle-hole energies e; and single-particle quadrupole
moments (or other moments for other choices of V, such as octupole mo-
ments). However, as hinted at a couple of times, the single-particle energies
can be Nilsson energies equally well, and the A.\*s and A.'s can be creation and
destruction operators for Nilsson orbits. Then *¥a is the ground state of the
deformed nucleus. A related point concerns our choice for V in our example
of a schematic model in Eq. 9.18. It seems to ignore the short-range parts of the
residual (nonsingle particle) interaction such as the pairing interaction.
However, the whole formalism is identical if the pairing interaction is incorpo-
rated into the definition of the single-particle or Nilsson energies by defining
these as quasi-particle energies instead of single-particle energies. Then the
TDA or RPA can give the structure of vibrations in deformed nuclei in terms
of amplitudes for specific Nilsson quasi-particle excitations. We thus see how a
rather formidable looking formalism leads simply and elegantly to an easy way
of deducing the particle (or quasi-particle) composition of specific collective
vibrations. This is one answer to the apparent dichotomy that we noted earlier
between the independent particle picture of the nucleus and the existence of
collective excitations and correlations. The key element, of course, is in a
sense inserted a priori by defining the operators Oaf as linear combinations of
single-particle (or two-quasi-particle) operators and by simplifying the
interaction V. Ultimately, the method is tested by its agreement with experi-
ment. This test has been passed many times, making the result a useful,
elegant, and powerful approach to the microscopic structure of collective
vibrations.
disappear. Excitations involving orbits from the shells below or above the
valence shell must be incorporated in realistic calculations, and the RPA
becomes a necessary refinement.
Even higher-order forms of the operator Oaf are sometimes applied and are
often called higher RPA (HRPA) for obvious reasons. Other approaches such
as that of Kumar and Baranger or the more recent refinement, the dynamic
deformation theory, maybe used. The full sweep of many-body theory encom-
passes many varied and complementary approaches.
One of the most interesting, informative, and physically transparent appli-
cations of the RPA is to the /bands in deformed nuclei. The relevant
calculations were carried out by Bes and co-workers in the early 1960s and
remain the standard for the microscopic structure of these modes.
The basis states here are naturally the Nilsson orbits. Pairing must be
included via a BCS calculation, so that, for the ground state, there is a
distribution of occupation amplitudes over several orbits near the Fermi
energy /I. This can be seen quantitatively in a modification to Eq. 9.20RPA. For
this discussion, we must specifically label both orbits involved in the quasi-
particle excitation as we must specify occupation amplitudes for each. Replac-
ing the particle-hole energy Aer with Ei + E., where the £'s are quasi-particle
energies defined by
where Qtj now means (i \ r'Y^ \f). The wave functions are now linear combina-
tions of two-quasi-particle excitations rather than single-particle-hole excita-
tions, but the physical idea is identical. Note that in Eq. 9.22, E. changes more
slowly than e when (e.- A) < A. Therefore, a wider range of e. values and orbits
can contribute. Also, the pairing factor favors pairs of quasi-particle excita-
tions on opposite sides of the Fermi surface; as an analogue to particle-hole
excitations, this is not surprising.
The matrix elements of r2Y2±2 that determine the important y-vibrational
amplitudes are easily deduced by writing r2Y2±2 in Cartesian coordinates as
This is a field (operator) that does not change nz (there is no effect in the z-
direction) but that changes A by ±2. (For a table giving the structure of all the
low multipole operators in Cartesian form, see Mottelson and Nilsson, 1959).
Another practical selection rule is that the sum nz + nf + ny = N should be
conserved; otherwise the matrix element would involve Nilsson wave func-
tions [Nnz\] and [N'n'^A'] with N' = N±2. Such states are far apart (about 10
MeV). Thus, the important components of the y-vibration will be two-quasi-
particle excitations involving a Nilsson orbit within about 2 MeV of the Fermi
324 Collectivity, Phase Transitions, Deformation
The analogous proton amplitudes can be similarly deduced. To see how well
this estimate works, we show the detailed results of RPA calculations in Table
9.1, from which it is evident that we have indeed identified the most important
two-quasi-particle components.
Table 9.1. Principle neutron two-quasi-particle amplitudes (xlOO) for the y-vihration in several
rare earth nuclei*
154 160 1M 170 174Yb !78Hf 184W 186W
Two-quasi-particle states Gd Dy Dy Er
ll/2-[505]®7/2-[514] — 7 7 — —
— — —
11/21505]® 7/21503] 11 12 15 25 26 21 11 8
9/21514]® 5/21523] 9 7 — — —
—
— —
9/21514]® 5/21512] 25 23 21 20 14 — — —
9/21505] ® 5/21503] — — — — 11 16
— —
7/21523]® 3/21521] 19 12 — — — — — —
7/21514]® 3/21512] — 17 35 48 34 26
— —
7/21503]® 3/21501] — — — — — 7 20 26
5/21523] ® 1/21541] 9 — — —
—
—
— —
5/21523] ® 1/21521] 20 36 51 29 14 8 — —
5/21512]® 3/21521] — — —
9 7 — — —
5/21512] ® 1/21510] — — 9 31 56 56 32 24
3/2~[532]® 1/21530] 24 8 — — —
—
— —
3/21521] ® 1/21541] 9 — — — —
— — —
3/2~[521] ® 1/21521] 27 45 46 25 13 8 — —
3/21512]® 1/21510] — — — 8 13 22 66 70
Comparison with Table 9.1 again shows that these comprise most of the main
neutron components of the j-vibration in 170Er. Perusal of this table shows how
various two-quasi-particle amplitudes systematically grow and decay across
the region as the Fermi surface rises, and how different amplitudes are favored
by the energy denominator and pairing factors in Eqs. 9.20RPA and 9.22.
It is sometimes possible to test such predictions using single nucleon trans-
fer reactions. Since the 183W target ground state is the l/2-[510] orbit, the (d, p)
reaction can only populate two-quasi-particle states in 184W of the specific form
1/2"[510] ® v/-Nils- Moreover, they will contribute proportionally to IP-, the
emptiness of yNils in 183W. In 183W, therefore, the hole state 5/2-[512] component
should not contribute significantly to o(d, p), for the /-vibration, while neutron
transfer involving the 3/2~[512] orbit should. (Of course, in 182W the 7-vibration
must have similar structure. Hence, 5/2"[512] transfer will be important for the
183
W(d, t)182W cross section.) In any case, the (d, p) cross section to the various
spin states of the 7 rotational band in 184W will be determined primarily by the
C coefficients for the 3/2~[512] orbit.
The explicit expression for o(d, p) into an even-even nucleus is not quite as
simple as given in Chapter 8 for an odd-mass final nucleus, since each state of
spin / can be constructed by coupling a component (/j) of the 1/2~[510] orbit
with one;2 from the 3/2~[512] orbit. In general, several/^ pairs can lead to the
same/. Their relative contributions are given by Clebsch-Gordon coefficients.
The population of a two-quasi-particle state in an even-even nucleus in (d, p)
is thus given by a generalization of Eq. 8.1
where the 0( are DWBA cross sections (/ = ;'±l/2). It can easily happen that
several /values contribute to this expression for a given final state. In the case
of 184W, for example, the J* = 2+ state can arise by coupling the target;' = l/2~
state with; = 3/2" or; = 5/2' components of the 3/2~[512] wave function; these
proceed by / = 1 and / = 3 transfer, respectively. For a higher spin target such as
167
Er (7/2+), there are even more possibilities. A 4- level can be formed by
326 Collectivity, Phase Transitions, Deformation
coupling the 7/2+ ground state with;2 = 1/2, 3/2, 5/2, 7/2, 9/2,11/2 (the upper
limit here is provided not by angular momentum conservation but by the
available orbits in the N = 5 shell where ;max = 11/2). This involves six
contributions, with /=!,!, 3,3,5,5, respectively.
Thus, generally the "fingerprint patterns" in even-even nuclei will be
somewhat "washed out" and less orbit sensitive than in odd nuclei. Neverthe-
less, the (d, p) data for population of the y-band in 184W does have a pattern
very close to that predicted for the transfer of a 3/2-[512] neutron Nilsson orbit,
and the corresponding (d, t) data in 182W confirms the presence of the 5/2~[512]
orbit in the y-vibration there. Note that not only can the important orbits be
identified in this way, but their amplitudes in the /-vibration may be deduced
(approximately) by comparing, for example, the cross sections into the
even-even nucleus to those for transfer of the same orbit into the neighboring
odd A nucleus. In this way, even the detailed structure of RPA calculations are
confirmed, at least for a couple of important amplitudes. Other amplitudes,
such as 3/2-[512] ® 7/2-[514] cannot be directly tested.
In a few special cases it is possible to test for amplitudes involving a non-
target-ground state orbit. As an example, 185W has a ground state 3/2"[512] and
a low-lying 1/2~[510] excited state. If 185W were stable, permitting the relevant
(d, t) reaction to 184W, it would be possible to study the l/2-[510] ® 5/2~[512]
component in the 184W y-vibration, because the (3/2~) ground state of 185W has
a strong Coriolis admixture of the 3/2 1/2~[510] state. (The case starting with
183
W is not the same, even though here too the 1/2[510] and 3/2~[512] orbits are
admixed, since the l/2~ 183W ground state cannot contain a 3/2 3/2"[512] admix-
ture.) In the 185W(d, t)184W case, the y band would be primarily populated by a
coherent combination of the components 3/2~[512] ® 7/2[514] and 1/2~[510] ®
5/2[512].
To return to our discussion of the structure of collective excitations, we see
how an understanding of the basic formalism and philosophy of the RPA
allows us to understand and even anticipate the detailed microscopic structure
of excitations such as the y vibrations of deformed nuclei. A little further
thought reveals five additional basic features of these excitations.
First, the unique parity orbit is generally unimportant in the y vibration,
since this vibration has positive parity and therefore both quasi-particles have
to be in a unique parity orbit. But, from the sequence of asymptotic Nilsson
quantum numbers, we see that any pair with AK = 2 also has Anz = 2, and
therefore will not be connected by a r2Y2±2 operator.
Second, since the important orbits are those within a certain distance of the
Fermi surface, the y vibrational wave functions will not change radically from
one nucleus to the next or even over a small region. This is almost a
requirement of a collective mode. We can go even further and predict how the
structure will vary. From our discussion of the relevant (large) matrix ele-
ments and of the role of the energy denominator, it is clear that a given Nilsson
orbit will, for some mass A, be high above the Fermi surface and will contribute
little. As A (Noi Z really) increases, this Nilsson orbit will drop closer to the
Fermi surface and gain importance. Later, it will become a hole state and begin
Microscopic Treatment of Collective Vibrations 327
decreasing in amplitude as it recedes further from the Fermi energy. Thus, for
most two-quasi-particle excitations, a plot of their contributions against NOT Z
(whichever is relevant) will be bell shaped. Inspection of Table 9.1 shows that,
for most orbits, this is precisely the observed behavior. Amplitudes for two-
quasi-particle states such as 9/2~[514] and 5/2^[512] containing an excitation
from the h^ orbit from the next lower major shell peak early in the deformed
region. The 5/2~[523] ® 1/2"[521] component increases into the deformed
region and attains its maximum amplitude near N = 9S (164Dy), but contributes
little for N > 104. Finally, the 7/2 -[514] ® 3/2~[512] combination is negligible
until rather late in the shell, but becomes strong near N = 108 and again drops
off in importance near jV = 112.
Third, we note that for any given nucleus, the number of significant compo-
nents is surprisingly low, typically three to five for neutrons and slightly fewer
for protons (since the protons are filling the lower shell with fewer orbits).
Thus the word "collective" must be taken in context. One should not imagine
50 to 100 nucleons or amplitudes significantly involved in this or other collec-
tive excitations.
Fourth, one should address the question of how the collectivity (as meas-
ured, for example, by large B(E2:7-> g) values) arises if only a few orbits
contribute strongly. This is especially so since "single-particle" B(E2) values
in odd-mass deformed nuclei (i.e., B(E2) values where one particle changes its
Nilsson orbit) are much smaller than in "shell model" nuclei. The former are
typically 10~3 to l(h* s.p.u., since the shell model strength is fragmented by
configuration mixing in deformed nuclei. Yet B(E2:y->g) values are typically
=10 s.p.u. This can only arise then as a specific effect of coherence: the
dominant contributions to these E2 matrix elements must add in phase.
This can be understood (at least qualitatively) by expressing the structure of
the vibration in an equivalent but, in a sense, inverted picture that we alluded
to briefly earlier. Instead of conceiving the vibration as built up of components
that arise by operating with r2Y2±2 on the orbits near the Fermi surface,
imagine a set of closely spaced two-quasi-particle states at some excitation
energy (e.g., 2 to 3 MeV) that mix by an appropriate interaction (that need not
be specified). Then, by the arguments concerning two- and multistate mixing
in Chapter 1, one level will be pushed down and its wave function will have the
most coherent admixture of amplitudes. More specifically, if one starts with
the idealized case of a set of degenerate levels with equal mixing matrix
elements, the lowest state after mixing is described by a thoroughly mixed
wave function (see Eq. 1.13) with all amplitudes identical and with the same
phase. For the still schematic case of nondegenerate but equally spaced initial
states that mix with equal matrix elements, a similar result is obtained. For the
more realistic case of more or less random initial spacings but roughly
comparable ^Y^ matrix elements (since otherwise the orbits in question
would not be important in the /vibration), the same qualitative feature
persists. Sample diagonalizations show this. In particular, the lowest state
always has its various wave function amplitudes in phase. Thus the B (E2:y-> g)
328 Collectivity, Phase Transitions, Deformation
values will have the maximum possible collectivity consistent with this set of
orbits and matrix elements.
This leads to the fifth feature—the energy systematics of the /vibration—
which we can again deduce qualitatively without explicit calculation. We need
only refer once again to the aforementioned mixing calculations and recall that
(taking for simplicity the case of N degenerate levels mixing with each other
with equal matrix elements) one state is pushed down well below all others and
that its energy lowering is (A7 - l)V, where V is the mixing matrix element.
Thus the y vibration will be lower when there are more contributing (mixing)
states. This occurs precisely when there is an abundance of Nilsson orbits near
the Fermi surface with identical nz values so that many large r 2 Y 2i2 matrix
elements with A« = 0, AA = ±2 contribute.
It was shown in our earlier discussion of the Nilsson model that at the
beginning of a shell, the steepest downsloping orbits have high nz (for prolate
nuclei). Since these orbit energies are so strongly correlated with the extent of
the wave function in the z direction, even neighboring orbits will have different
nz values. This makes it difficult to find nearby orbit pairs with AA = ±2 and
An = 0. Near the top of a shell, the large changes in orbit angles for different
K values means that there will be fewer orbits and that they will be further
apart on average. Thus, both cases only permit a few important y-vibrational
amplitudes. Just before midshell, however, one encounters a region where
orbits with various nz and A values congregate so that there will be more
contributing Anz = 0, AA = ±2 orbit pairs, hence greater collectivity and a lower
y-band energy. This qualitative behavior is sketched in Fig. 9.3 (which also
includes actual values calculated by Bes). These may be compared with the
observed systematics shown earlier in Fig. 2.17. The similarity of both ob-
served and calculated patterns to our simplified estimates is remarkable. (The
sharp drop empirically observed in E near A = 190 is beyond this approach
since it involves the onset of /-softness and large yms values even in the ground
state.)
It is worth noting that this systematics stems from the specific detailed
microscopic structure of the y vibration. It is not simply a handwaving
statement that vibrations will be lowest, and be most collective, at midshell
where there are the most valence nucleons. Few of the valence nucleons
actually participate. Moreover, other excitations, such as the j8 vibration (see
the following), have radically different systematics.
Similar arguments can be applied to other vibrational modes such as octu-
pole (or hexadecapole vibrations). For example, for K - 0~ octupole vibra-
tions, the relevant operator is Y3 0 and hence A«2 = ±3 AA = ±0 amplitudes are
critical. Being of negative parity, octupole vibrations need two orbits of
opposite parity. Thus, they must involve the unique parity orbit, and therefore,
there will not generally be as many available two-quasi-particle states that can
contribute. Octupole vibrations thus tend to be rather high-lying and not very
collective. They should lie lowest early in the deformed region where there are
a number of A nz= ± 3, AK = 0 combinations involving excitations from the high
/ normal parity orbits into the unique parity orbits.
Microscopic Treatment of Collective Vibrations 329
Fig. 9.3. Qualitative estimate of the systematics of 7, /3, and octupole vibrations across ashell. The
second panel shows the actual calculations for a yvibration from Bes (1965).
For neutrons in the rare earth region, for example, amplitudes such as
l/2-[530] ® l/2+[660], 3/2-[521] ® 3/2+[651], and 5I2~[S12] ® 5/2+[642] can con-
tribute. As the shell fills, the low K unique parity orbits, whose participation
is essential, begin to fill in the ground state, effectively blocking the collectivity.
Thus K = 0, 3" excitations should rise in energy. Late in the shell, the normal
parity orbits are all low K, while the nearest unique parity ones are high K.
Hence K - 0~ octupole excitations are inhibited. Now, however, K = 3~
octupole vibrations, with amplitudes satisfying An = 0, AA = ±3 such as
ll/2+[615] ® 5/2-[512] or 9/2+[624] ® 3/2-[521], begin to increase in collectivity
and drop in energy. Thus one expects the qualitative systematics shown in Fig.
9.3, where the K sequence of successive octupole excitations in a given nucleus
is seen to change from K = 0-3 to K = 3 - 0 as a major shell fills. This
inversion of K ordering across a shell is verified in detailed calculations by
Neergaard and Vogel and confirmed by experimental systematics. It is inter-
esting that this inversion arises from the interplay of the quadrupole (Nilsson
330 Collectivity, Phase Transitions, Deformation
The calculation of (^ I ^Y^ | vgs) is just a special case of the general result
discussed in Chapter if We have now
The last term vanishes since Y2 0 cannot change the orbits of both particles.
The first two matrix elements are just the quadrupole moments of two par-
ticles in orbits b and a, respectively. Thus,
and this is small unless the orbits a and b have very different quadrupole
moments.
As we have discussed, the orientation of a Nilsson orbit (and hence its
quadrupole moment) is closely linked to its up- or downsloping character. For
prolate nuclei, downsloping orbits are equatorial and have prolate quadrupole
moments, while upsloping orbits are oblate. So ft vibrations should be rela-
tively collective and low-lying only in regions of the Nilsson diagram where
orbits with very different slopes lie near each other. Inspection of Fig. 7.4
Microscopic Treatment of Collective Vibrations 331
shows that at the beginning and end of major shells, regions of strongly up- and
downsloping orbits approach each other from different shells. (The fact that
they have different /lvalues is inconsequential, since the allowed r^Y^ matrix
elements in Eqs. 9.27 and 9.28 are diagonal: they do not connect the two orbits
involved.) Near midshell, there are virtually no such orbit combinations.
Thus, P vibrations should be low in energy at the start and end of a shell and
should peak (and have the lowest collectivity) near midshell. This is illustrated
in Fig. 9.3 and can be compared to the empirical situation in Fig. 2.17. The
qualitative agreement is reasonable.
The brief discussion of the TDA and RPA formalism in this chapter, and its
application to some of the most important low-lying vibrations in medium and
heavy nuclei is meant only as an introduction to the field. In practical
calculations there are numerous subtleties (e.g., detailed choices of single-
particle energies) and sophistications (e.g., higher-order operators, O r , than in
Eqs. 9.11 and 9.12, or other related approaches involving self-consistent many-
body theory). Nevertheless, the underlying physics is always similar to that
outlined here. It is hoped that the present discussion will have removed some
of the mystery from such calculations and indicated how simple arguments,
based on the form (operator) for each type of vibration and the available
single-particle orbits, can lead to reasonable deductions of the principle wave
function components, their mass dependence, collectivity, and energy sys-
tematics.
This page intentionally left blank
PART IV
EXPERIMENTAL TECHNIQUES
This page intentionally left blank
10
A FEW SELECTED EXPERIMENTAL TECHNIQUES
where Z, and Z2 are the projectile and target atomic numbers, and r12 is the
time-dependent distance between target and projectile. Clearly, the probabil-
ity for exciting the nucleus by this Coulomb field is related to the magnitude of
V(r12) along the hyperbolic collision trajectory. The crucial ingredients are
A Few Selected Experimental Techniques 337
Fig. 10.1. (Top) Diagram of some important parameters in the Coulomb excitation process.
(Bottom) Some E2 matrix elements involved in multiple Coulomb excitation.
then the product Z1Z2 and the sequence of values taken by r12. Since most of
the excitation will take place when the Coulomb field is strongest, the most
important physical parameter is p, commonly called the distance of closest
approach.
Clearly, the scattering angle 6 is closely connected with p: smaller values of
p during the collision lead to greater Coulomb interaction strengths and
therefore larger scattering angles. Conversely, the smallest distance of closest
approach, pmin, will be obtained when 9= 180°, and is given by
where df(9, £,) arises from an integral over the orbit of the projectile along the
classical trajectory. This expression embodies our previous comments. The
cross section increases as the distance of closest approach of the projectile
decreases, or as its charge increases. It also increases with increasing Ep : the
1/v factor is more than compensated by large df(6, £) values for small £ (large
Ep ) as we will discuss next. The df(8, £,) factor contains a sin-40/2 dependence
representing the underlying classical Rutherford scattering. (Often, Coulomb
excitation yields are given as a ratio to Rutherford scattering, removing this
strong dependence.) More importantly, df(9, %) decreases exponentially with
£. For large TJ, the integral of df(9, £) over 9 drops by a factor of 3 as \ goes
A Few Selected Experimental Techniques 339
from 0 to 0.5 and by a factor of 10 as £goes from 0.5 to 1.0. In our example of
60 MeV 16O ions on 150Sm, these three <fj values correspond to excitation
energies of 0,1.5, and 3.0 MeV, respectively. Clearly, Coulomb excitation is a
tool for exciting low-lying states!
The preceding discussion implicitly assumed a single-step Coulomb excita-
tion process. There is no inherent reason, however, why multiple excitations
cannot occur during the collision process. Such multiple Coulomb excitation is
actually very common with heavy ion projectiles. The idea is illustrated in Fig.
10.1 (bottom). It is convenient to view this as a sequential process in which the
target is excited first to one level, then subsequently to another. However,
recognizing the extremely short collision times involved (= 10"22 seconds) the
uncertainty principle AtAE > h implies a large uncertainty in the system energy
during the collision. This permits the process of "virtual" excitation over short
time periods during which the conservation of energy is not required. Since
the probability of direct Coulomb excitation decreases exponentially with
excitation energy while virtual excitation decreases roughly linearly, it turns
out that the virtual excitation process is dominant in most multiple Coulomb
excitation situations. While high-lying states can be virtually excited during
the collisions, the range of final excitation energies produced is governed by an
exponential behavior. Nevertheless, the presence of virtual excitation both
offers an opportunity to excite levels not otherwise accessible and imposes a
complexity on the resulting analysis because the extraction of nuclear matrix
elements from Coulomb excitation data ultimately involves solving a set of
simultaneous equations that involve all matrix elements (and their generally
unknown relative signs) that connect all levels involved, virtually or otherwise,
in the excitation process.
The solution of these equations can be grea tly facilitated by certain experi-
mental tactics. One of these is based on the recognition that the only direction
fixed in space in a Coulomb excitation collision is that provided by the incident
beam direction. By conservation of momentum, those excitation processes in
which the projectile is scattered backwards at 180° can involve no transverse
momentum transfer to the target nucleus and, therefore, only the m = 0
magnetic substates of the final nuclear levels are excited. This immediately
reduces the number of simultaneous equations to be solved by a factor on the
order of (2/+1)2, where J is the average spin of the levels excited in the target
nucleus.
This quantum mechanical constraint can be imposed experimentally by
detecting the de-excitation y-rays in coincidence with back scattered projec-
tiles. Excitation of m * 0 substates increases slowly for 0 <180°, so that in
practice, one measures the backscattered particles at angle, ranging from 160°
to 175°. This technique uses annular detectors in which the incident beam
passes through a central hole and the backscattered particles are recorded in
the back angle annulus. Coincidence with backscattered particles has another
advantage: it preferentially selects those collisions with high excitation proba-
bilities (p small), thereby enhancing yields and leading to increased multiple
excitation so that B(E2) values for higher-lying states can be measured.
340 Experimental Techniques
There is another aspect of the fact that the incident beam defines a direction
in space. De-excitation y-rays have a particular angular distribution. In
principle, one should measure the yields at a series of angles and integrate over
angle. However, since virtually all Coulomb excitation proceeds by E2 excita-
tion (these are the strongest matrix elements connecting low-lying states), any
angular distribution involved can only include Legendre polynomials P((cos0)
of order 0, 2, and 4. Moreover, the coefficient of the quartic term is small in
most cases. Therefore, many Coulomb excitation experiments are performed
with the y-ray detectors situated at 55° relative to the beam axis since, at this
angle, P2 also vanishes, making the yields nearly identical to those obtained by
integrating over all angles.
Another simplification exploits the Z1 and Eprof dependence of Coulomb
excitation cross sections by carrying out a sequence of experiments starting
with light, low-energy projectiles. Typically, in even-even nuclei, Coulomb ex-
citation by protons or a-particles excites only the first level or the first couple
of levels. It is easy to extract the nuclear matrix elements for these. Indeed, it
was in the early days of Coulomb excitation, when only these projectiles were
available, that many B(E2:2+1 -» 0*1) values were first determined. With these
matrix elements in hand, a second experiment is performed, using a heavier,
and/or more energetic, projectile to excite the next set of levels. This process
continues until all of the matrix elements are deduced sequentially. Most
Coulomb excitation studies of heavy nuclei have been carried out in this
manner. In recent years, however, the development of sophisticated search
algorithms in conjunction with the deBoer-Winther code has led to the possi-
bility of simultaneously fitting an extensive set of Coulomb excitation meas-
urements made with high energy, heavy ion projectiles to extract a reasonably
unique and consistent set of matrix elements.
Thus far we have discussed experiments in which the presence of Coulomb
excitation is indicated by the observation of de-excitation y-rays. This is the
most common approach. It is hampered, however, by a built in ambiguity: one
does not generally know what fraction of each level's population (which is
deduced from the number of detected y-rays de-exciting that level) resulted
from direct excitations and from the decay of higher-lying levels. The detec-
tion of the outgoing, scattered projectile gives the number of excitations of
each level directly. Although the energy resolution of typical particle detec-
tors, or the poor efficiency of high resolution magnetic spectrometers, ham-
pers the use of this technique, it is often a valuable tool. It also leads to an
alternative to the preceding step-by-step approach to unravelling the E2 ma-
trix elements. In the traditional approach just described, this is achieved by
choosing the experimental conditions (Z1 and Ep) so that pmin attains different
values sequentially, the smallest corresponding to the most complex excitation
processes. From Fig. 10.1 it is clear that this can also be obtained with a single
projectile and fixed energy, simply by varying the angle 6 at which the scattered
projectile is detected. With the use of multiple particle detectors positioned at
different scattering angles, such a program can be carried out in a single
experiment, thus saving valuable running time and assuring identical experi-
mental conditions for each angle measured. The cost is that, for most angles,
the full set of magnetic substates must be included in the calculation.
Fig. 10.2. Data from two Coulomb excitation experiments. (Top) a-induced excitation of 160Gd
(Ronningen, 1977). (Bottom) Excitation of 232Th by 208Pb ions (Ower, 1982). The diagram shows
the experimental arrangement (see text). The spectrum at bottom illustrates the relative yields of
different y-rays while the boxes at the right give the population of two spin states as a function of
6. Note that the higher spin level is not seen at forward angles and is also much weaker at back
angles although, as expected, its yield increases faster with 9 than the J = 8 level.
342 Experimental Techniques
A final criterion that controls the choice of many of the projectiles used in
Coulomb excitation is that one generally wants to avoid projectile excitation
itself. Figure 10.1 is drawn in the laboratory system, but in the center of mass
system, it is impossible to distinguish projectile and target. One therefore
usually chooses projectiles that have very high-lying first excited states, gener-
ally magic or doubly magic nuclei such as a particles, 16O or M8Pb. Sometimes
it is more convenient to reverse the kinematic conditions and use the nucleus
conceptually considered to be the target as a projectile. This is useful when the
appropriate projectiles are not available, but it is also used, by choice, because
of the different scattering angles then obtained in the laboratory system. For
example, by using a Pb projectile on a Mg target, all yields will be in the
forward direction. In some cases such reversed kinematics can be advanta-
geous.
To obtain a concrete feeling for Coulomb excitation spectra, it is useful to
inspect the two examples shown in Fig. 10.2. The top one nicely illustrates low
energy, a induced, Coulomb excitation with detection of the inelastically
scattered projectiles. The strongest peak corresponds to elastic scattering.
The 2+ and 4+ members of the ground band are also strong. The excitation
probabilities fall off very rapidly as more excitation steps are required and the
6+, 8 + ,... states are not seen. Weaker, higher-lying peaks correspond to the
vibrational levels, 2+r and 3". The former is much stronger because there are
two excitation routes and because it involves E2 rather than E3 excitation.
Also note that Coulomb excitation is almost always limited to natural parity
states [it = (-1)' such as 0+, 2+, 4+, 3", but not 1+ or 3+] simply because of the
angular momentum conservation rules. This is rigorous in a single-step
process and nearly always characterizes multiple Coulomb excitation yields
also.
The lower part of Fig. 10.2 shows the experimental set up for 208pb induced
Coulomb excitation of 232Th. Such massive projectiles impart considerable
recoil to the target leading to Doppler shifts in the y-ray energies that depend
on the angle of recoil. The "Doppler" detector tags a specific angle. The
typical Ge(Li) y-ray spectrum shown demonstrates the ability to populate high
ground band and even side band states (n=-, here). (The 2+ -» 0+ and 4+ -» 2+
y-ray transitions are actually the strongest in this reaction; their weakness in
the spectrum is due to the fact they are highly converted.) Figure 10.2 includes
typical yields as a function of the angle of the scattered projectile. Note the
strong increase with 9 corresponding to smaller impact parameters p.
To understand multiple excitation more thoroughly, let us consider the
excitation of the 2+ band in Fig. 10.1. Even for relatively light projectiles there
are two possible excitation routes, a direct one from the ground state and a
double excitation via the 2\ state. In typical experimental conditions in which
backscattered particles are selected, the probability of exciting the 2+t state
may well be greater than 50 percent per detected backscattered particle. In
rotational nuclei, the Coulomb excitation matrix elements, <2+r E212 + t ) and
<2+rl E210 + j>, are simply related by Clebsch-Gordon coefficients. The latter is
slightly larger. However, it also involves a larger A£ than the step linking the
A Few Selected Experimental Techniques 343
Fig. 10.3. Coulomb/nuclear interference effects in ascattering near the Coulomb barrier (Bemis,
1973).
illustrated in Fig. 10.3, which shows the excitation cross sections for the 2*l and
4 + j levels divided by the pure Coulomb excitation cross sections. Deviations
from unity represent the effect of the Coulomb/nuclear interference. This
technique has been applied to many nuclei, in particular those in the W region.
The measured yields can also give hexadecapole shape components because of
the large interference effects for 4+ levels.
Until recently, Coulomb excitation experiments were carried out almost
exclusively below or at the Coulomb barrier. The reasoning was that interfer-
ence with direct or compound nuclear reactions would obscure the simplicity
of the technique. However, the increase of Coulomb excitation amplitudes
with projectile energy is not vitiated by competition with other processes. And
higher incident energies can provide access to higher-lying excitations. This
point has recently be exploited in very high-energy Coulomb excitation studies
of collective states, such as giant resonances (single or multiple) where other
excitation amplitudes are small. A technique thought by some to have already
witnessed its greatest popularity is thus being rejuvenated.
The (n, y) reaction has been used for several decades to elucidate nuclear
structure. It has some severe limitations and some powerful advantages. Its
primary attribute is nonselectivity, which makes it complementary to selective
processes such as Coulomb excitation or transfer reactions. In certain forms,
such as average resonance capture (ARC), the (n, y) reaction is actually
designed to be completely nonselective and to provide an a priori guarantee of
disclosing all states in certain spins and excitation energy ranges. The basic
idea of the (n, 7) reaction can be seen in Fig. 10.4 (left). An incident low-energy
Fig. 10.4. (Left) Schematic illustration of the difference between thermal neutron capture and ARC. (Right) Examples of the spin values and relative averaged in-
tensities available in the ARC process on even- and odd-mass targets.
346 Experimental Techniques
sequentially populating, say, 50 to 100 individual resonances, and then sums all
the spectra so obtained. The statistical Porter Thomas distributions of each
resonance would then be nearly completely averaged out and all states of a
given final spin would be populated with nearly the same summed intensity.
Such an approach, although conceptually appealing, is clearly tedious and time
consuming. An alternative that achieves the same end automatically is to use
a beam of neutrons that has been rendered specifically non-monoenergetic.
This can be achieved by passing a reactor neutron beam through certain filter
materials, notably scandium and ^Fe, which have minima in their neutron
scattering cross sections near 2 keV and 24 keV, respectively. Having passed
through such materials, a reactor neutron beam will have an energy centered
around 2 or 24 keV with a FWHM (typically 1 to 2 keV) determined by the
thickness of the filter material. The FWHM must be sufficiently broad so that
a large number (Nr) of resonances are simultaneously populated, but not so
broad that it degrades the energy resolution of the primary y-ray transitions
following capture. The width of these lines is a convolution of the detector
(usually Ge) resolution (typically = 5 to 6 keV at 6 MeV) with the energy width
of the incident neutron beam.
To illustrate the process more concretely, consider a typical deformed
even-even target nucleus. Since the ground state is 0+ and 2 keV neutrons are
predominantly s wave, the capture state will be J = 112*. For nuclei such as
155
Sm, 167Er, 183W, the level spacing of l/2+ states at 1 MeV is «10 eV. Therefore,
a 2 keV neutron beam with 1 keV FWHM will populate =100 resonances. The
fluctuations in the combined Porter Thomas distribution go approximately as
21'i~N~r. Therefore, instead of orders of magnitude variations in primary
transition intensities, the fluctuations in the ARC spectrum intensities will be
± 2/VlOO = ± 20 %. (Actually, this variance refers to the reduced intensities,
defined as IJE5^ since it is necessary to remove a secular energy dependence.)
If we consider just El primary transitions (they dominate Ml transitions by a
factor of = 6), the final states will be 1/2-, 3/2- (Fig. 10.4, right).
When the target is odd mass, there are two possible capture state spins.
Each will decay by El (and weaker Ml, which we ignore) primary transitions.
In the simplest approximation, neglecting a spin dependence of the level
density near the neutron separation energy, both of these spins contribute
equally. Thus, as illustrated in Fig. 10.4 (right), the relative reduced intensities
will fall into two groups differing by a factor of = 2. An ARC spectrum was
shown for 196Pt in Fig. 5.8. The 195Pt target has J* = 1/2-, so that the (y-wave)
capture state is Jc = 0~, 1". Decay of the 1- levels by El transitions lead to 0+, 1+,
and 2+ states. The 0~ level only feeds 1+ levels. Final states with J* = 1+ should
thus be twice as intense as 0% 2+ levels. Of course, as we have seen in many
models, low-lying 1+ states are very rare. So, in practice, we expect rather
uniform intensities. The good averaging in Fig. 5.8 is immediately apparent,
especially when contrasted with typical intensity fluctuations of one to two
orders of magnitude in thermal capture experiments.
The preceding examples correspond to 2 keV ARC measurements where ,y
wave capture still dominates. At 24 ke V, both s and p wave components are
348 Experimental Techniques
Fig. 10.5. Ratio of reduced primary intensities in ARC at 2 and 24 keV in 168Er, for states of known
parity (Davidson, 1981).
roughly comparable and a wider range of final J" values becomes accessible.
The ratio of reduced ARC intensities at 2 and 24 keV provides a nearly ideal
parity meter for the final states. An example is shown in Fig. 10.5.
The limitation on excitation energies accessible in ARC stems from two
sources. First, with increasing Ex, the level density increases, which may make
it impossible to resolve individual levels. At first this is not a serious problem
since an unresolved pair of levels appears in the spectrum as a single peak of
twice the area and is therefore readily identified. At higher excitation ener-
gies, however, and especially in odd-odd nuclei, the probability of missed
levels rapidly increases. Second, and more importantly, the secular decrease of
the observed reduced intensities / (recall, it is the quantity /r IE 5 that is
constant) eventually leads to a point at which they can no longer be safely
detected above the background.
It is important to realize the consequences of the ARC-based guarantee of
completeness. It is not simply a question of finding an additional level or two
beyond that obtainable with other techniques. Its most profound importance
lies not so much in finding all levels (of certain J" and Ex ranges), but in
knowing that one has found all levels. This rather subtle point is best discussed
by two illustrative examples. Consider the imaginary level scheme shown on
the top in Fig. 10.6 and suppose that the ARC process populates levels with
spins 2+, 3+, 4+ via El primary transitions. The figure clearly depicts a de-
formed rotational nucleus with an excited vibrational band containing at least
2+ and 4+ levels. However, the figure is ambiguous as to whether this is a 7 band
with an undetected 3+ state or a p" band with undetected 0+ state. Even though
0+ states are not accessible to the ARC process in our example, this question
can be unambiguously decided. If the ARC data discloses a 3+ level between
the 2+ and 4+ states, then the band is clearly a 7 vibrational excitation. Equally
certain, if no such 3+ level is detected, there must be an undetected lower-lying
0+ level below the 2+ state. (Only K* - 0+ bands have a spin sequence 0', 2 + ,
A Few Selected Experimental Techniques 349
Fig. 10.6. Examples of "completeness" arguments using ARC. (Bottom left) All n = -levels in
168
Er; (Bottomright) Arrangement into rotational bands. The numbers of bands of each lvalue
above - 1700 keV (right of dashed line) are discussed in the text.
Table 10.1. 168Er. Numbers of negative-parity states and deduced rotational band structures!
J*
r T 3- 4~ 5-
Number of observed states 2 1 4 5 6
Unique allocation of these
to rotational bands
one £*=0~ band 1 — 1 — 1
one K*=l~ band 1 1 1 1 1
two .£*= 3~ bands — — 2 2 2
two K*= 4~ bands — — — 2 2
tThe tabulation includes all negative-parity levels with spins from 1 to 5~ above 1780 keV not already allocated
to rotational bands with band-heads below 1570 keV.
'Davidson, 1981
tional hands. Between 1400 and 2000 keV, an ARC measurement disclosed
the set of negative parity states shown on the left in the lower part of Fig. 10.6.
The 167Er ground state is 7/2+, so the capture spins are 3+, 4+: therefore E1(M1)
primary transitions following s and p wave ARC populate all l~-5~ levels (up to
2 MeV) and distinguish the 2', 5~ group from 3", 4~ states from the factor of two
difference in reduced intensities. Combined with other data, a full set of
unambiguous J* assignments was made, but no structural information about
these states was known. However, given just the knowledge that this set of
negative parity states is complete, one can not only deduce the number of
rotational bands present, but their precise K values and the number of levels of
spin 0~, without having observed them. The idea is illustrated in Table 10.1.
The top gives the number of levels of each J. The key point is that the number
of 5~ levels is one greater than the number of 4- states. There are only two ways
to produce a 5~ state without an accompanying 4~level, either in a K* = 0~ band
with levels 1~, 3~, 5~, or in a K* = 5~ band with levels 5~, 6~, 7,.... It is easy to rule
out the latter. There are two 1~ levels and only one 2~ level. Again, the only
way this can occur is if there is a K* - 0^ band. Thus there cannot be a K* = 5~
band. Having established that there is one K = 0~ band, we have accounted for
one 1", one 3~, and one 5" state. Since there are two 1" levels and only one
K* = 0~ band, the other 1- state must be the bandhead of a K* = I' band. This
accounts for one each of the 1~, 2~, 3~, 4~, and 5~ levels, and thereby exhausts the
1~ and 2' levels but leaves two 3" levels unaccounted for. Each of these must be
K* = 3~ bandheads that account for two additional 4" and 5~ levels. There are
still two levels of each of these spins unaccounted for, and thus there must also
be two K* = 4' bands.
This now accounts for all of the states disclosed in the ARC measurement.
We have deduced the exact number of negative parity rotational bands and
their K values, and the fact that there is only one 0~ band and no K = 5 bands
below 2 MeV in 168Er. And we have deduced all this without assigning a single
state to a single rotational band, attempting to fit any of them into specific
rotational energy patterns, or inspecting their y-ray decay. The key was the
knowledge that the set of levels was complete. Some other experiment that
A Few Selected Experimental Techniques 351
disclosed the same set of levels without providing this knowledge could not
have led to the same analysis.
Fig. 10.7. Excitation functions for / = 4 and J = 6 states in the reaction 123Te(a, ny)126Xe (Lieberz
1989). The yields (7SF are for "side-feeding" transitions that exclude discrete feeding. Note the
nearly monotonic behavior against Ef (labeled for each line in keV) and the dependence of the
slopes on J.
rays also provide information on level spins, but they are seldom as unambigu-
ous as conversion electron measurements.
We will illustrate some of these techniques by focusing on the exceptional
capabilities of two instruments at the Institut Laue Langevin in Grenoble,
namely the electron magnetic spectrometer, BILL, and the y-ray spectrometer,
GAMS. Both are characterized by very high resolution and reasonable
efficiency. BILL achieves this by magnetic focussing, while GAMS measures
y-ray energies by diffracting the y-rays at precisely measured angles. Solid
state Ge y-ray detectors normally have resolutions ranging from = 800 eV at
100 keV to = 1.5 keV at = 800 keV. The corresponding numbers for the GAMS
spectrometers in normal operation are = 50 eV and = 400 eV. The corresponding
y-ray energy precision is typically 50 to 200 eV for Ge detection and 1 to 20 eV
for GAMS. A portion of a typical GAMS spectrum is shown in Fig. 10.9. The
primary impact of such resolution is not the higher level energy precision that,
in any case, far exceeds the predictive capabilities of modern nuclear models.
It is rather in the avoidance of accidental energy combinations ("Ritz combi-
nations ") that can lead to incorrect y-ray placements and, thereby, incorrectly
deduced structure information.
Consider an example of a typical deformed nucleus in which 250 y-rays are
Fig. 10.8. Typical level scheme from the (a, wf) reaction (positive parity levels only). (Lieberz, 1989).
Fig. 10.9. Examples of GAMS and BILL data. Note the remarkable resolution and energy precision of GAMS and the sensitivity to transition multipolarities
resulting from the resolution of L and M lines with the BILL spectrometer. From Krusche, 1989 (GAMS) and based on data discussed in Gelletly, 1981 (BILL).
A Few Selected Experimental Techniques 355
Fig. 10.10. K conversion coefficients and L subshell ratios for El, E2, and Ml transitions.
Lifetimes for transitions of "single-particle" strength (including conversion) as a function of tran-
sition energy are also shown (right).
356 Experimental Techniques
Fig. 10.11. Schematic idea of the flow of population and the de-excitation mechanisms in (heavy
ion,;t«y) reactions.
decay that can carry off considerable energy but relatively little angular
momentum. It therefore proceeds along nearly vertical de-excitation paths.
Given the high complexity of the compound state just prior to y-ray emission,
there are a myriad of possible routes and the decay is characterized by a nearly
statistical "rain cloud" pattern.
There have been many studies in recent years of the average properties of
the statistical continuum of y-rays including measurements of their average
multipolarities, average energies, and so on. We will not delve into this here.
The net result is that, finally, states relatively near the yrast levels are reached.
At this point the de-excitation path must closely follow the yrast line. The de-
excitation converts to the emission of a sequence of E2 y-ray transitions
feeding down any of several nearly parallel rotational bands in the yrasl
the short time since, numerous other superdeformed sequences have been
found. At first, only high spin levels were associated with such structure.
Recently, in the Hg region, superdeformed states with / ~ 10 have been
assigned.
Initial interpretations of the superdeformed states involve the concept of
energy gaps in the Nilsson diagram at large deformations caused by the
presence of strongly downsloping Nilsson orbits from the N' = N+2 oscillator
shell. These deformed "shell gaps" act in a similar way to spherical shell gaps
and lend greater stability to those nuclei whose Fermi surface is near the gap.
The detailed characterization of superdeformed levels and the correct predic-
tion of their systematics offers a challenge to modern theories, both of high
spin states and of the behavior of single particle levels well above the Fermi
surface. Critical to such predictions has been the development of sophisti-
cated "cranked" shell model calculations that properly account for the effects
of extremely fast rotational motion, along with the incorporation of specific
corrections for shell structure not included by simply specifying the deforma-
tion parameters. The effects of the monopole proton-neutron interaction may
well be important in a detailed understanding of some of the latter effects.
Fig. 10.14. Illustration of three trajectories, and the associated angular distribution, for a heavy
ion transfer reaction.
Fig. 10.15. Spectra for three heavy ion single neutron transfer reactions leading to 14!lSm (Bond,
1983).
A Few Selected Experimental Techniques 363
This list includes two categories of references. One consists of those specifically cited in
the text (for example, as the source for figures or tables). (The author apologizes for
the disproportionate and semi-obscene number of citations of his own work here. This
reflects both the fact that the topics covered in the book tend to be those that I am
interested in and may have worked on, and a degree of laziness in choosing illustrative
figures from my own papers rather than searching the literature for the better ones.)
The other category gives a few classic articles and other excellent works, often text-
books, review articles, or compilations, which are suggested for additional study. All of
these are recommended. Papers that the author has personally found to be continually
and particularly helpful, some indispensable reference works, and a few others are
marked with an asterisk. No attempt is made to provide a complete reference list for
ideas and data discussed in this book. This list is for the benefit of the reader, not for the
recognition of earlier accomplishments. That is the task of appropriate review articles.
References are cited in the text by first author only, with date of publication.
Alaga, G., Alder, K., Bohr, A., and Mottelson, B.R. (1955). Mat. Fys. Medd. Dan. Vid.
Selsk. 29, No. 9.
*Alder, A., Bohr, B., Huus, T, Mottelson, B.R., and Winter, A. (1956). Rev. Mod. Phys.
28, 432.
Alder, K. and Winther, A. (1966). Coulomb Excitation (Academic Press, New York).
Aprahamian, A., Brenner, D.S., Casten, R.F., Gill, R.L., and Piotrowski, A. (1987).
Phys. Rev. Lett. 59, 535.
Aprahamian, A., Brenner, D.S., Casten, R.F., Gill, R.L., Piotrowski, A., and Heyde, K.
(1984). Phys. Lett. 140B, 22.
Arenas Peris, G.E. and Federman, P. (1988). Phys. Rev. C38, 493.
Arima, A., Harvey, M., and Shimizu, K. (1969). Phys. Lett. 30B, 517.
*Arima, A. and Iachello, F. (1976). Ann. Phys. (N.Y.) 99, 253.
*Arima, A. and Iachello, F. (1978a). Ann. Phys. (N.Y.) 111, 201.
*Arima, A. and Iachello, F. (1979). Ann. Phys. (N.Y.) 123, 468.
Arima, A. and Iachello, F. (1978b). Phys. Rev. Lett. 40, 385.
Arima, A., Otsuka, T., Iachello, F., and Talmi, I. (1977). Phys. Lett. 66B, 205.
Arvieu, R. and Moszkowski, S. A. (1966). Phys. Rev. 145, 830.
*Baranger, M. (1960). Phys. Rev. 120, 957.
Barrett, B.R. and Kirson, M. W. (1970). Nucl. Phys. A148, 145.
Bemis, C. E., Jr., Stelson, P. H., McGowan, F. K., Milner, W.T., Ford, J.L.C., Jr.,
Robinson, R.L., and Tuttle, W. (1973). Phys. Rev. C8, 1934.
Bengtsson, T., Ragnarsson, I., and Aberg, S. (1988). Phys. Lett. 208B, 39.
*Bes, D.R., Federman, P., Maqueda, E., and Zuker, A. (1965). Nucl. Phys. 65, 1.
Bes, D.R. (1961). Mat. Fys. Medd. Dan. Vid. Selsk. 33, No. 2.
Bes, D.R. (1963). Nucl. Phys. 49, 544.
Bes, D.R. and Sorensen, R. A. (1969). Adv. in Nucl. Phys. 2, 129.
Bohle, D., Richter, A., Steffen, W., Dieperink, A.E.L., Lo Iudice, N., Palumbo, F., and
Scholten, O. (1984). Phys. Lett. 137B, 27.
365
366 References
*Bohr, A. and Mottelson (1969, 1975). Nuclear Structure. Vols. I and II, Benjamin,
New York.
Bohr, A. and Mottelson (1953). Mat. Fys. Medd. Dan. Vid. Selsk. 24, No. 16.
Bond, P.D. (1983). Comments Nucl. Part. Phys. 11, 231.
Braun-Munzinger, P. (1984). Editor, Nuclear Physics with Heavy Ions, Nucl. Sci. Res.
Conf. Series, Vol. 6, Harwood Academic, New York.
Brink, D.M., de Toledo Piza, R., and Kerman, A.K. (1965). Phys. Lett. 19, 413.
Bromley, D.A., Gove, H.E., Paul, E.B., Litherland, A.E., and Almquist, E. (1957).
Can. J. Phys. 35, 1042.
Brown, G.E. and Bosterli, M. (1959). Phys. Rev. Lett. 3, 472.
*Brussaard, P.J. and Glaudemans, P.W.M. (1977). Shell Model Applications in Nuclear
Spectroscopy, North Holland, Amsterdam.
Bucurescu, D. G., Cata, G., Cutoiu, D., Dragelescu, E., Ivascu, M., Zamfir, N. V, Gizon,
A., and Gizon, J. (1989). Phys. Lett. 229B, 321.
Bunker, M. and Reich, C. W. (1971). Rev. Mod. Phys. 43, 348.
*Castanos, O. and Draayer, J. P. (1989). Nucl. Phys. A491, 349.
*Casten, R. F. (1985a). Nucl. Phys. A443, 1.
Casten, R. F. (1986). Phys. Rev. C33, 1819.
Casten, R. F, Aprahamian, A., and Warner, D.D. (1984). Phys. Rev. C29, 356.
Casten, R. F., Brenner, D.S., and Haustein, P.E. (1987). Phys. Rev. Lett. 58, 658.
Casten, R. F., von Brentano, P., and Haque, A.M.I. (1985b). Phys. Rev. C31, 1991.
Casten, R. F. and Cizewski, J.A. (1978). Nucl. Phys. A309, 477.
Casten, R. F., Frank, A., Moshinsky, M., and Pittel, S. (1988). (Editors), Contemporary
Topics in Nuclear Structure Physics, World Scientific, Singapore.
Casten, R. F, Heyde, K., and Wolf, A. (1988b). Phys. Lett. 208B, 33.
Casten, R. F, Kleinheinz, P., Daly, P.J., and Elbek, B. (1972). Mat. Fys. Medd. Dan. Vid.
Selsk., 38, No. 13.
*Casten, R. F., Warner, D.D., Brenner, D.S., and Gill, R.L. (1981). Phys. Rev. Lett. 47,
1433.
Casten, R. F., Warner, D.D., and Aprahamian, A. (1983). Phys. Rev. C28, 894.
*Casten, R. F. and Warner, D.D. (1988a). Rev. Mod. Phys. 60, 389.
Casten, R. F., Warner, D.D., Stelts, M.L, and Davidson, W.F (1980). Phys. Rev. Lett. 45,
1077.
Cizewski, J. A., Casten, R.F., Smith, G.J., Stelts, M.L., Kane, W.R., Borner, H.G., and
Davidson, W.F. (1978). Phys. Rev. Lett. 40, 167.
Cizewski, J.A., Casten, R.F., Smith, G.J., MacPhail, M.R., Stelts, M.L., Kane, W.R.,
Borner, H.G., and Davidson, W.F. (1979). Nucl. Phys. A323, 349.
Cohen, B.L. and Price, R.E. (1961). Phys. Rev. 121, 1441.
Cottle, P. D., and Bromley, D. A. (1987). Phys. Rev. C35, 1891.
Davidson, W.F., Warner, D.D., Casten, R.F., Schreckenbach, K., Borner, H.G., Simic, J.,
Bogdanovic, M., Koicki, S., Gelletly, W., Orr, G.B., and Stelts, M.L. (1981). J.
Phys. G7, 455.
Davydov, A.S. and Filippov, G.F. (1958). Nucl. Phys. 8, 237.
Deleplanque, M.A. and Diamond, R.M. (1988). GAMMASPHERE Proposal.
de Shalit, A. and Goldhaber, M. (1953). Phys. Rev. C92, 1211.
*de Shalit, A. and Feshbach, H. (1974). Theoretical Nuclear Physics, Vol. I, Nuclear
Structure, Wiley, New York.
*de Shalit, A. and Talmi, I. (1963). Nuclear Shell Theory, Academic Press, New York.
*Dobaczewski, J., Nazarewicz, W, Skalski, J., and Werner, T. (1988). Phys. Rev. Lett.
60, 2254.
Eisberg, R. and Resnick, R. (1974). Quantum Physics, Wiley, New York.
Eisenberg, J.M. and Greiner, W. (1970). Nuclear Models, Vol. 1, Excitation Mechanisms
of the Nucleus, Vol. 2, and Microscopic Theory of the Nucleus, Vol. 3, North Hol-
land, Amsterdam.
Endt, P.M. and Van der Leun, C. (1978). Nucl. Phys. A310, 1.
References 367
*Molinari, A., Johnson, M. B., Bethe, H.A., and Alberico, W. M. (1975). Nucl. Phys.
A239, 45.
Mottelson, B.R. and Nilsson, S. G. (1959). Mat. Fys. Skr. Dan. Vid. Selsk., 1, No. 8.
Nagai, Y., Styczen, J., Piiparinen, M., Kleinheinz, P., Bazzacco, D., von Brentano, P.,
Zell, K. O., and Blomqvist, J. (1981). Phys. Rev. Lett. 57, 1259.
*Nathan, O., and Nilsson, S. G. (1965). In Alpha, Beta, and Gamma Ray Spectroscopy,
Vol. 1 (ed. K. Siegbahn), North Holland, Amsterdam.
Neergaard, K. and Vogel, P. (1970). Nucl. Phys. A145, 33.
*Nilsson, S. G. (1955). Dan. Mat. Fys. Medd. 29, No. 16.
Nilsson, S. G. and Prior, O. (1961). Mat. Fys. Medd. Dan. Vid. Selsk 32, No. 16.
Nilsson, S. G., Tsang, C. F., Sobiczewski, A., Szymanski, Z., Wycech, S., Gustafson, C.,
Lamm, I.L., Moller, P., and Nilsson, B. (1969). Nucl. Phys. A131, 1.
Ower, H., Elze.Th. W., Idzko, J., Stelzer, K., Grosse, E., Emling, H.,Fuchs, P., Schwalm,
D., Wollersheim, H. J., Kaffrell, N., and Trautmann, N. (1982). Nucl. Phys. A388,
421.
Paar, V. (1979). Nucl. Phys. A331, 16.
Preston, M.A. and Bhaduri, R. K. (1975). Structure of the Nucleus, Addison-Wesley,
Reading, MA.
Preston, M.A. and Kiang, D. (1963). Can. J. Phys. 41, 742.
*Raman, S., Malarkey, C.H., Milner, W.T., Nestor, C.W., Jr., and Stelson, P.H. (1987).
At. Data and Nucl. Data Tables, 36, 1.
Ratna Raju, R.D., Draayer, J.P., and Hecht, K. T. (1973). Nucl. Phys. A202, 433.
*Riedinger, L.L., Johnson, N.R., and Hamilton, J.H. (1969). Phys. Rev. 179, 1214.
Riedinger, L.L., Smith, G.J., Stelson, PH., Eichler, E., Hagemann, G.B.,Hensley, D.C.,
Johnson, N.R., Robinson, R.L., and Sayer, R. O. (1974). Phys. Rev. Lett. 33,
1346.
Riley, M.A., Simpson, J., Sharpey-Schafer, J.F., Cresswell, J.R., Cranmer-Gordon,
H.W., Forsyth, P.D., Howe, D., Nelson, A.H., Nolan, P.J., Smith, P.J., Ward, N.J.,
Lisle, J.C., Paul, E., and Walker, P.M. (1988). Nucl. Phys. A486, 456.
Ring, P. and Schuck, P. (1980). The Nuclear Many-Body Problem, Springer, New York.
Ronningen, R.M., Hamilton, J.H., Ramayya, A.V., Varnell, L., Garcia Bermudez, G.,
Lange, J., Lourens, W., Riedinger, L.L., Robinson, R.L., Stelson, P.H., and Ford,
J.L.C., Jr. (1977). Phys. Rev. C15, 1671.
*Sakai, M. (1984). At. Data and Nucl. Data Tables 31, 399.
Scharff-Goldhaber, G. and Weneser, J. (1955). Phys. Rev. 98, 212.
Scholten, O., Iachello, F., and Arima, A. ( 1978). Ann. Phys. ( N.Y.) 115, 325.
Sheline, R. K. (1960). Rev. Mod. Phys. 32, 1.
*Schiffer, J.P. (1971). Ann. Phys. (N.Y.) 66, 798.
*Schiffer, J.P. and True, W. W. (1976). Rev. Mod. Phys. 48, 191.
*Stephens, F.S. (1975). Rev. Mod. Phys. 47, 43.
Stephens, F.S. and Simon, R.S. (1972). Nucl. Phys. A183, 257.
Tabor, S.L., Cottle, P.D., Gross, C.J., Huttmeier, U.J., Moore, E. F., and Nazarewicz, W.
(1989). Phys. Rev. C39, 1359.
*Talmi, I. (1962). Rev. Mod. Phys. 34, 704.
Twin, P., Nyako, B.M., Nelson, A. H., Simpson, J., Bentley, M.A., Cranmer-Gordon, H.
W., Forsyth, P.D., Howe, D., Morhtar, A. R., Morrison, J.D., Sharpey-Schafer,
J.F., and Sletten, G. (1986), Phys. Rev. Lett. 57, 811.
Van Duppen, P., Coenen, E., Deneffe, K., Huyse, M., and Wood, J.L. (1985). Phys. Lett.
154B, 354.
Van Maldeghem, J., Heyde, K., and Sau, J. (1985). Phys. Rev. C32, 1067.
Warner, D.D., Casten, R.F., and Davidson, W.F. (1980). Phys. Rev. Lett. 45, 1761.
Warner, D.D. (1984). Phys. Rev. Lett. 52, 259.
Warner, D.D. and Casten, R.F. (1982). Phys. Rev. Lett. 48, 1385.
*Warner, D.D. and Casten, R.F. (1983). Phys. Rev. C28, 1798.
Waroquier, M. and Heyde, K. (1971). Nucl. Phys. A164, 113.
References 369
This listing gives the locations of principal discussions of various ideas and
topics. Occasionally, casual references to an idea or technique (e.g., paren-
thetical comments in the text) are not listed. Also, themes running throughout
the text, sometimes implicitly, such as the attractive nature of the nuclear
force, the Pauli principle, two-state mixing, the nuclear potential, collectivity,
and so on are only cited where they are explicitly the principal topic. Ideas,
nuclei, and so on that are the subject of figures are cited by the page numbers
where those figures are discussed rather than where the figures themselves
occur. Greek letters and words are alphabetized according to their latin
spelling and acronyms (e.g., RPA, GAMS, CFP's) are generally given accord-
ing to the way they are commonly thought of: thus, CFP's are listed under "co-
efficients..." while GAMS is given according to the letters themselves. Such
decisions are often quite arbitrary.