Sei sulla pagina 1di 9

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.

5.3 Multiparticle Configurations j": The Seniority Scheme


When there are numerous particles outside closed shells, they can enter
different shell model orbits. For example, in 40"Zr59 the nine valence neutrons
might be in a configuration (dM)6 (g^)3. Here, the dM shell is filled and the ear-
lier arguments on the effect of closed shells on the values of AE( j2'1 +1 /"2J) tell
us that the dM orbit can be neglected, so this configuration is equivalent to
(g7/2)3. Now, consider 95Zr. In this case, the lowest expected configuration
would be (d5/2)5. By the particle-hole equivalency discussed earlier, this is
exactly equivalent to a single neutron in the d5/2 orbit, leading to a one-state
configuration with J =j = 5/2 and, indeed, the ground state of 95Zr is 5/2+.
However, one could also imagine excited states in 95Zr of the form (d5/2)3 (g7/2)2.
Normally, at least near closed shells, such configurations are rather high-lying
excited states: our primary interest is usually in the lowest-lying levels in which
as many particles as possible are packed into the lowest accessible j value.
Thus, at least in simple shell model treatments, one is frequently interested in
;" configurations. Moreover, even though realistic shell model calculations will
often involve important components coupling two / values, an understanding
of the single ;' case greatly helps to interpret and even anticipate such calcula-
tions. So far, we have ignored the possibility of both valence protons and neu-
trons. This clearly complicates the situation, as seen in the discussion earlier of
the 5-function interaction for p-n systems. Moreover, as we shall see later,
once one has nucleons of both types outside of closed shells, collective effects
rapidly accumulate and other models provide alternate, and often better,
approaches. Therefore, it is appropriate to again stress the/1 configuration of
identical nucleons. Despite this restriction, the following considerations have
extremely wide applicability.
The tendency of particles to pair to J = 0+ leads to a scheme in which this
property is explicitly recognized and exploited. Consider the /" configuration.
We ask what is the smallest value of n that can produce a given / value.
Denoting this value by v, it is clear that there can be no particles coupled in
pairs to J - 0 in the configuration f j . (Otherwise, a f-2 configuration would
have a spin/.) Such a state is then said to have seniority v. From a configuration
y+2 we can make a state of the same spin J by coupling one pair of particles to
7 = 0. This state is also said to have seniority v. Physically, v is simply the
number of unpaired particles in a state of angular momentum J in the
configuration;". The number of paired particles is (n - v) and the number of
such pairs is (n - v)/2. For v = 0, all particles are paired and 7 = 0.
Let us further illustrate this concept with a simple example. Consider the
116 Shell Model and Residual Interactions

(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

of sums over operators acting on each particle. A one-body odd-tensor


operator acting in a f configuration is given by U* = "Z^U/1. Since an odd
tensor operator cannot change any }2J = 0) pair to one with J * 0 (J even), odd
tensor operators must conserve seniority. Equation 5.4 shows that there is no
contribution to U* from such pairs of particles coupled to / = 0. Thus matrix
elements of one-body odd-tensor operators in/1 configurations with seniority
v, can be reduced to those in the f configuration. Moreover, they are inde-
pendent of n (for n > v).
These results follow so trivially from Eq. 5.4 that the preceding comments
essentially constitute a derivation. However, they are so important and basic
that it is worthwhile to go through the arguments more explicitly. Consider a
matrix element such as (j"vJ' ".=1U* \j"vJ). Since v < n, the left side can be
rewritten in terms of wave functions of the configuration ly^vC/')/2^ = O)/)
and similarly for the right side. For simplicity, we take the particles thus
separated off as the (n-l)th and nth particles. Application of Eq. 5.4 to these
two particles contributes nothing to the overall matrix element, and we can
replace the operator "Z.=1IJ * with "~2Z.=1U* extending over n-2 particles. Since
the matrix element is now independent of the last two particles, we can
integrate over them. Since they are in the same state j2J = 0), this integral is
unity by orthogonality. If (n - v) > 4, we can repeat this procedure for another
pair of particles. We continue this procedure for any even v until we are
dealing with an operator "L.^U* acting on the states |;vJ}and 1/7'). Thus we
obtain

which shows both the reduction of a matrix element in the f configuration to


one in;'" and the independence of n.
The other result, conservation of seniority, is equally obtainable. Suppose
the two wave functions in the above matrix element have different seniorities
v, v' < n. There is some point in the successive reduction (the successive
peeling off of pairs of particles) where an overlap integral over the wave
functions fJ = 0} and \j2J 0} appears. Clearly, by orthogonality, this
vanishes. To reiterate, we have the absolutely critical results:
Odd-tensor single-particle operators conserve seniority in ;'" configura-
tions.
The matrix elements of odd-tensor single-particle operators in;'" configu-
rations in the seniority scheme can be reduced to ones in the f configura-
tion.
These matrix elements are independent of n.
These rather abstract results have many practical applications. They imply,
for example, that the magnetic moment of the 7/2" state of an (f7/2)3 configura-
tion is identical to that in the single particle L,I2 configuration: in general,
magnetic moments in odd mass nuclei where the valence particles occupy a
given; orbit should be independent of the (odd) number of valence nucleons.
Similar arguments cannot be applied to even tensor operators like the quad-
rupole operator. It turns out that these operators are not diagonal in the
118 Shell Model and Residual Interactions

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 rarethey can only connect states of the same senioritywhile 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

Therefore, the energies of the v = 2 states are independent of n. Let us also


calculate the spacings within the v = 2 configuration. These are given by
120 Shell Model and Residual Interactions

Fig. 5.2. Illustration of the constancy of seniority v = 2 levels in/1 configurations.

Thus, all energy differences of seniority v = 0 and v = 2 states in the n-particle


configuration are identical to those in the two-particle system and are inde-
pendent of n. This is illustrated in Fig. 5.2. This result is crucial because its
absence would make it virtually impossible to apply the shell model in a simple
way to nuclei other than those within one or two nucleons of closed shells.
Indeed, this result was anticipated in the previous chapter in arguing that the
shell model has broad applicability. The low-lying levels of good seniority in a
f configuration are independent of n. In practice, more than one orbit; will be
occupied by the valence nucleons. Nevertheless, the present result can be
approximately generalized if one writes the wave function in the schematic
form

In fact, the incorporation of such two-body configuration mixing is essentially


equivalent to a modification of the interaction itself, and thus Eq. 5.6 is widely
applicable. The Sn nuclei (see Fig. 2.6) provide a classic example of Eq. 5.10
and its generalization to the mulli-/ case: the entire known set of v = 2 levels,
J = 0+, 2+, 4+, 6+, is virtually constant across an entire major shell. The Ca
isotopes (Fig. 2.3) provide another example that will be discussed later.
To recapitulate, for the matrix elements of odd-tensor operators and inter-
actions between states of good seniority in/ 1 configurations:
One-body matrix elements (e.g., dipole moments) are

Two-body interactions arc linear in the number of


Multiparticle Configurations 121

It is possible to derive analogous results for even-tensor operators and


interactions. The derivations involve the manipulation of, and recursion
relations for, CFP coefficients. These are tedious but straightforward. We will
cite two important results. Once again, we look first at one-body operators.
As noted, even-tensor operators do not necessarily conserve seniority and can
link states with Av = 2. The expression for such matrix elements is given by:

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

That is, in the/" configuration, the B(E2) value, defined as

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

This equation again expresses an n-particle matrix element for states of


seniority v in terms of the v-particle matrix element. It has an interesting
behavior as a function of n, as given by the factor outside the matrix element.
In terms of/(the fractional filling of the shell), the numerator behaves simply
as (1 - If). It therefore has opposite signs in the first and second halves of the
shell and hence must vanish identically at midshell. This is, of course, an
extremely important result, indicating that, for example, quadrupole moments
of;'" configurations in even-even nuclei change sign in midshell. The generali-

Potrebbero piacerti anche