You are on page 1of 22

Naclear Physics A223 (1974) 541-562; @ North-Holland Publishing Co.

, Amsterdam
1.D.l:
I l.D.2 I Not to be reproduced by photoprint or microfilm without written permission from the publisher

ISOTOPE EFFECTS IN TIN


WITH THE LOCAL DENSITY APPROXIMATION

X. CAMP1 t and D. W. L. SPRUNG


Physics Department, McMaster University, Hamilton, Ontario, Canada L8S4Ml

and

J. MARTORELL
Institat de Physique NucIPaire, Division de Physique ThPorique tt, 91406 Orsay, France and
Departamento de Fisica, Universidad Autdnoma, Madrid, Spain

Received 7 February 1974

Abstract: Properties of the even isotopes of tin are calculated using the density dependent effective
interaction G-O. Excellent agreement is found for the gross structure ;of these nuclei. The
isotope shift is studied and the factors governing this effect are elucidated.

1. Introduction

In a previous paper ‘) it was shown that Hartree-Fock calculations using the density
dependent effective force G-O give a good overall description of the properties of
doubly magic nuclei: their density distributions, binding and single particle energies.
This agreement was found to extend to the isotopes of He and 0. In this paper we
consider the even isotopes of Sn, which pose a severe test on the model. Considerable
experimental information is available on these nuclei. The binding energies are
known for 104 5 A 5 128, while elastic electron scattering data, optical, X-ray and
muonic atom data are available for the region A = 112 to 124. The single particle
energies and their occupations vi are fairly well known. It is a challenge to our
model to give a reasonable accord with this data, and to see whether such fine effects
as the isotope shift are reasonably well described.
Calculations of these effects have already been done by some authors 2*3), using
semiphenomenological theories. Those by Bunatyan and by Krainov and Miku-
linskii ‘) were done in the framework of the Migdal theory, whereas those byUher
and Sorensen “) use a phenomenological description of the core and treat the residual
interaction with the pairing plus quadrupole model. In both cases the core response
to the addition of neutrons is described in a parametrized way, and thus strongly
depends on the model utilized. In our calculation the core behaviour is given by the
t Present address: Institut de Physique NuclBaire, Division de Physique Theorique tt, 91406 Orsay,
France.
tt Laboratoire AssociC au CNRS.
541
542 X. CAMP1 et al.

microscopic HF theory, which we expect to be more reliable in describing these effects.


With our method, information on the strength and shape of the polarizing field
created by the added neutrons can be extracted. Similar monopole core polarization
effects in the Pb-Bi isotone shift have been described by Sick et al. “) also with a
microscopic selfconsistent formalism using Skyrme effective forces.
The gross effects of the residual interaction are included in our calculation in a
simplified way (BCS theory plus quadrupole vibrations introduced phenomenologi-
tally), so that their effects can be accounted for satisfactorily, and their importance
discussed. Contributions due to the neutron electromagnetic structure have been
included following the work of Bertozzi, Friar, Heisenberg and Negele [BFHN] 5),
and their effect in the isotopic changes in charge densities is found to be important.
The experimental data now available appear to determine the density shifts Q(Y)
between tin isotopes in some detail, so that properties other than 6(r*) can be ex-
tracted. Radial moment analysis “) has been used to correlate such experimental
data with our calculated charge densities (p,) and isotopic charge density differences
(6Pc).

2. Description of the calculation


2.1. THE NUCLEAR MODEL
Our model, described at length in ref. l), uses the Hartree-Fock approximation
with a density dependent effective interaction. This effective force G-O was developed
to reproduce the G-matrix elements of the Reid potential, as a function of density
and energy in nuclear matter. Since it was intended to be used in Hartree-Fock cal-
culations of doubly magic nuclei, averages over the J and L dependence of the G-
matrix elements were made so that an effective central force resulted. A param-
etrization in terms of Gaussians was selected, giving finally

V(r, lir, W) = C (ai+ biki) exp [ -(r/~i)*] +A(kF)(W- WO(kF))S(r). (2.1)

In nuclear matter, the total density p equals (2/3x*)ks and W,(k,) is at each density
twice the average single particle energy; W is the corresponding quantity in the finite
nucleus under consideration.
The parameters Ui, bi and A differ in each spin, isospin subspace and their values
may be found in ref. ‘). The parameter A = 3 gives a density dependence which is
rapid at subnuclear densities but less rapid at higher density, in agreement with
nuclear matter calculations of the G-matrix. In a finite nucleus, k, is determined by an
arithmetic average of the densities of the two types of particles concerned (e.g.
2pP for a pp interaction) at their positions. Finally, the parameters a,, b, of the short-
est range Gaussians in the even states were renormalized so that the force G-O
saturates nuclear matter at k, = 1.35 fm-‘, with B/A = 16.5 MeV per particle. Two
phenomenological parameters thus introduced allow the good overall agreement
with experiment mentioned in the introduction.
ISOTOPE EFFECTS 543

The present calculations were carried out in essentially the same way as those of
ref. ‘). The principal differences will be briefly described here. First of all, we have
altered our choice of the prescription for calculating the “starting energy” W in a
finite nucleus. In ref. ‘) we discussed a number of possible prescriptions, but opted
for “case b”, as used by Negele ‘). In principle, W should be just the sum of the single
particle energies of the two interacting particles, with their Coulomb energy sub-
tracted because one compares to W, from nuclear matter in which the Coulomb force
does not exist. This would lead to a problem of orthogonality between states of the
same Z, (e.g., Is, 2s, 3s) so in such a case a weighted average value of the energies
was made and used for any one of the states of that 1. In the example proposed, this
would be essentially equal to the energy of the 2s level. Prescription (d) is even
simpler, it is simply to average the (Coulomb corrected) single particle energies for
all the occupied states, and use this single value everywhere required.
In the present study we found it preferable to adopt the simpler prescription (d).
The single particle energies so obtained have a better order and spacing, and lead to
occupations uf in better agreement with experiment in the BCS calculation to
be described in a moment. For example, in case (b) the lh, level, being weakly bound,
receives a strong attractive effect from the energy dependent term (W- W,). This
results in this level being more bound than the 3s, or 2d+ level, and in the BCS cal-
culation it becomes too fully occupied. When the overall averagevalue of W is used,
(W- W,) becomes small and the lhK finds a better position. A similar situation
holds for the lg, level. Using prescription (d) we find the correct order for the single
particle levels, albeit the spread is about 20 % too great as compared with experiment
for the five least bound levels. [By experiment we mean the single particle levels
inferred from solution of the inverse gap equations by Gillet et al. “).I
The change from case (b) to case (d) increases the binding energies of doubly
magic nuclei by about 0.05 MeV/A. The rms radii are unchanged for light nuclei but
reduced by 0.005 fm in Pb. The BFHN contributions to the charge density, to be
discussed in subsect. 2.3, reduce the radii of light nuclei by 0.03 fm and those of heavy
nuclei by about 0.015 fm. This reduction is welcome as one of the problems noted in
ref. ‘) was the tendency of radii to be too small in Pb and too large in 0. It suggests
however that our chosen renormalization of the force is not optimal, and we should
seek saturation in nuclear matter at a slightly lower density in order to increase all the
charge radii. For the present work we have not made this readjustment, preferring
to keep the same force as used in ref. ‘).

2.2. PAIRING CORRELATIONS

A number of authors ‘) have applied BCS theory to the tin isotopes, in a model
where the single particle energies are adjusted phenomenologically and are fixed
from nucleus to nucleus as in the shell model. This model is successful in reproducing
the mass surface. A preliminary HF calculation using our force indicated too much
structure in the mass parabola, due to the discrete jumps in occupation of orbitals.
544 X. CAMP1 et af.

We, therefore, included pairing correlation effects in an approximate but self-consis-


tent manner. This leads to a double self consistency requirement. At each iteration,
the BCS equations are solved to produce the occupation probabilities v,” for the
varions orbitals, and the new HF fields are constructed taking into account these
partial occupations. We believe that in this way the most important effects of pairing
are included. Both the neutron and the proton fields are modified by the pairing effects.
The equations (2.9) to (2.23) of ref. ‘) for construction of the HF fields remain
valid providing that factors of j, = 2j,+ 1 representing the occupation of an orbital
are replaced by&u,“. The total energy of the nucleus is calculated from

We note that the rearrangement energy Ex [defined in subsect. 3.2 of ref. ‘)I now
depends on the occupations v,“.
The pairing theory used is the simple BCS theory with a constant pairing force
strength G = lS.O~(ll.O+~) MeV acting between a set of neutron levels within 10
MeV of the Fermi surface. One solves the equations

(2.3)

&,-- / Ir

- ---.-.-- , u,2+v,2
= 1,
d(~,--A)~i-A21

for the uuk~owns 1, d and u,“. The value adopted for G is that obtained by Flocard,
Kerman, Quentin and Vautherin [FKQV] lo) by an empirical fit in all the mass
table, but made 20 % stronger because our single particle spectrum is dilated by about
20 % compared to experiment. This slightly stronger G is able to give occupations
ug close to the experimental ones, as is shown in fig. 1. Here the boxes and x’s indicate
values of u,” extracted by (d, p) [ref. “‘>I experiments on the even isotopes. The
triangles are values of ZJ,”taken from (d, p) [ref. “)I on odd isotopes and the circles
from (p, d) [ref. I”)] ex p eriments. The orbitals entering into the pairing calculation
are Ih+, 2d,, 3s,, lg,, 2d,, lge and 2~~.
We found no pairing among the protons, using the corresponding value G =
18.0/(11.0+2) MeV, because there is a significant gap of about 6 MeV between the
occupied I & and the empty levels 1g*, 2d, +
The argument in favour of using the simple BCS theory is two-fold. It is simple
and it gives reasonable occupation probabilities. On the other hand, the force G-O was
developed so that its diagonal matrix elements reproduce properties of nuclear
matter. There is no reason to believe that its pairing matrix elements are realistic,
so it would be questionable to use G-O to calculate the pairing matrix elements.
ISOTOPE EFFECTS 545

100 IO8 II6 124 132


.\
Fig. 1. Occupation probabilities for the last neutron orbitals. Experimental points: boxes and crosses,
values from (d, p) experiments (even isotopes); triangles, values from (d, p) experiments (odd
isotopes) ref. “); circles, values from (p, d) experiments ref. I?). Theoretical results given by the
continuous line.

The coupled HF and BCS equations which we use are not based on a variational
principle, because of the density and energy dependence of the force G-O. In the case
of the Skyrme interaction, Vautherin 13) was able to formulate a variational
procedure leading to these coupled equations. Our solution is based on the
principle of maximum self-consistency rather than minimum energy. It can be regard-
ed as an approximation to the HF Bogolyubov calculation, in which one would
use the same force to construct all the matrix elements, both those used in the HF
part and those involved in the pairing calculation. In this case, however, as we have
stated above, one would have to return to the original G-matrix and study the validity
of the construction of an equivalent local force (G-O) for the purpose of calculating
the pairing matrix elements.+

t Recently D. Gogny has attempted to develop a phenomenological interaction which will give
reasonable results in HFB calculations. The results he presented at Munich r4) are promising. A
successful conclusion of such a program would show what properties the interaction should have, to
work in HFB calculations, and would pose a new challenge to the “realistic” forces deduced in the
local density approximation.
546 X. CAMP1 et al.

3. Construction of the charge density

3.1. ELECTROMAGNETIC STRUCTURE

The charge density was constructed from the proton and neutron densities pP,
p,, following the theory in BFHN 5), and allowing for c.m. motion according to the
harmonic oscillator prescription adopted in our previous work ‘). The protonic
contribution, pz is written

“:=G
1
s-e
Ir-r’pjd
pp(r’)d3r’, (3.1)

a2 = *(r,‘) - B2, J(ri) = 0.8 fm.

and B = Jh/moA is the c.m. oscillator length. We choose ho so that a density con-
structed from harmonic oscillator functions had the same rms radius as our actual
mass density. The proton electric form factor GE is taken to be of Gaussian form,
which allows us to combine the c.m. and finite size corrections as in eq. (3.1). We
checked that the use of an exponential charge distribution, c: (q ') = 1 .O/( 1.O+ a2q2)2,
a2 = 0.05561 fm2 made no perceptible difference to our result. Furthermore we
have found that a sum of two Gaussians which gives an excellent fit to the exponential
form factor G”up to q2 = 25 fme2 gives the same pz to four digit accuracy. We con-
clude that the use of a one Gaussian form factor is sufficient, provided it is adjusted
to give the same rms radius of the proton charge distribution.
According to BFHN there is a contribution to the charge density due to the
electrostatic structure of the neutron, p:, and a spin orbital contribution pz' coming

from non spin-saturated sub-shells.


For the first of these it is convenient to fit the neutron form factor as the difference
of two Gaussians, so that two applications of eq. (3.1) suffice to calculate pz in
coordinate space. We choose

-q2(R2-0.06)/6.0_,-q*(R'+0.06)/6.0,
GE(q’) = e (34
As suggested in BFHN, the 0.06 term has the effect that the slope of GE (q ‘) at q2 = 0
reproduces the very accurately measured thermal neutron slope. We find that R2 =
0.75 fm2 gives a fine visual “fit” to the same neutron form factor data of BFHN,
as shown in fig. 2. The Gaussian choice seems superior to their exponential form in so
far as it falls off faster at large q2.
The spin-orbital contribution is evaluated using an effective spin-orbit density
pt = - (~/Mc)~ +(l/r2)(d/dr)[rp,,(r)]. Here

Pd = C nIj,iX<~ ’ l)lj Pljr 2 (3.3)


Ijr

where /I,, = - 1.9135 is the neutron magnetic moment and $, = 2.29278 is the proton
ISOTOPE EFFECTS 547

magnetic moment minus 0.5 (since there is an orbital contribution for protons).
The occupation of the orbital nlj, is (2j+ 1) for a filled orbital, and the expectation
value of r - 2 is + I, - (1+ 1) for spin up, down levels respectively. Finally plj, is the

_04 I I I i I I I I I I I I I I I I I I I I I I I I1 I
0 5 IO IS 2.0 25

q2(frn'l
Fig. 2. Neutron electromagnetic form factor. Experimental points from ref. 2Q). The solid line corre-
sponds to the best fit: R* = 0.75 fm2; upper Iine: RZ = 0.5 fm’; lower line: lpz = 1.5 fm*.

radial density normalized to unity for particles in the orbital ljz. When both members
of a spin-orbit doublet are occupied, there is only a very small contribution to psi,
because the densities are very similar and the weights cancel. Since we include pairing,
many orbitals near the Fermi surface contribute to pst. Notice that in eq. (14) of
ref. “) there is an error of sign, and that we have included the contributions from all
the orbitals rather than just the last few. In the case of 4BCa we verified that this in-
clusion had no effect on the result reported in BFHN ‘).

3.2. NON-STATIC CONTRIBUTIONS TO THE CHARGE DENSITY


Our calculation treats monopole polarization of the nucleus and pairing correIa-
tions selfconsistently, but does not allow for a possible quadrupole polarization. We
have estimated this effect by supposing that the HF charge density performs small
oscillations around a spherical equilibrium shape. Taking the nuclear medium to be
incompressible, we can write the vibrating charge density as

The s,~ are the collective variables for quadrupole oscillations and A’ is a normaliza-
548 X. CAMP1 et al.

tion factor conserving volume. The time average can be written, to order p2, as

X(r) = p,(r) + 2 B2C4d(r)+ r’~b’(r)l~ (3.6)

where /3” = (~ojZ,,,a~~,l+o) is the ground state expectation value. In the framework
of the vibrational model, following for example Owen and Satchler 15), /I” can be
related to the transition probability B(E2; 2: + O+), by the equation

p* = B(E’4 !? 21 . (3.7)
[ 52 1 (r,2)2

This relation reduces to the one widely used [Uher-Sorensen 3), Stacey ‘“)I if we
express (r,“) = $Rg (R. = 1.2 A*) in terms of the equivalent spherical radius.
Relation (3.7) is preferable since it takes account of the proper variation of the charge
radius (slower than A”) and of the diffusivity. Experimental values of B(E2) for all

(rm-3)I pc“%”

- Go

- Go
-_-__Go+vib ceri-
= e-

Fig. 3. The “9n charge density and isotopic change in charge density 124Sn-*‘aSn, theoretical
(G-O) versus phenomenological [electron scattering, ref. 20)]; Sp is the change in the theoretical
charge density given by the zero point quadrupole motion.
ISOTOPE EFFECTS 549

the isotopes of tin were taken from a compilation I’). The vibrating charge density
differs from the static one only in the surface, since the corrections involve derivatives
of p, . In fig. 3 we show our density for l1 %n and ten times the vibrational correction
included in this density. Clearly it is a vet-; small effect which increases both the
diffuseness and the radius. Although our treatment is very phenomenological, we
hope that it will reflect correctly the magnitude of the dynamic corrections to the static
HF theory+.

4. Results
4.1. BINDING ENERGY AND SINGLE PARTICLE ENERGIES

The binding energy per particle B/A is compared with experiment in fig. 4. The
pairing effects, even though they are a small contribution, have given the smooth
parabolic shape, avoiding irregularities due to the shell structure which is over-
emphasized in pure HF calculations. We note a tendency towards too much binding
between A = 118 and 124, which could have been mitigated by a more rapid depen-

I
..t-,I
,...I.. .I .L.

100 108 116 124


‘32.,
Fig. 4. Calculated binding energies per particle compared to experiment.

t Dr. L. Kisslinger has remarked to us that where large momentum transfers are involved (as in
elastic electron scattering at large angles, where q 2 1.0 -1.5 fm-I), collective contributions to nuclear
form factors may become ineffective due to loss of coherence. Therefore, for higher momentum
transfe’r it is necessary to use a microscopic description, so that one can follow the single particle
transition matrix elements as they go out of phase.
550 X. CAMP1 et al.

TABLE 1
3inding energies per particle and separation energies
A WA S(P) SO-4
G-O em G-O em G-O exp

100 8.235 3.60


102 8.310 4.51 24.12
104 8.370 8.377 5.42 4.42 22.86
106 8.416 8.429 6.33 5.20 21.62 22.32
108 8.457 8.469 7.18 5.79 21.26 21.05
110 8.488 8.496 8.04 6.59 20.32 19.97
112 8.503 8.514 8.86 7.51 18.66 18.97
114 8.516 8.522 9.66 8.51 18.49 18.06
116 8.520 8.523 10.41 9.27 17.50 17.10
118 8.515 8.517 11.12 10.00 16.45 16.27
120 8.503 8.505 11.84 10.67 15.59 15.59
122 8.485 8.488 12.56 11.40 14.81 14.98
124 8.463 8.468 13.28 12.10 14.24 14.44
126 8.435 8.444 14.00 13.40 13.93
128 8.405 8.418 14.72 13.03 13.53
130 8.371 15.44 12.39
132 8.337 16.14 12.25

Experimental values are from ref. Is).

dence of the pairing strength G upon N; we have, however, kept to FKQV lo)
formula for G. Towards the end of the shell, EBcs + 0 so the effect is unimportant
after A = 124. Even with that schematic treatment of pairing the agreement between
our results and the experimental B/A curve is excellent. The error in B/A varies be-
tween 5 and 13 keV per particle, or 0.6 to 1.6 MeV for the entire nucleus. Table 1
summarizes the calculated values for B/A and for the two neutron separation energies
S(2n) = B(Z, N) - B(Z, N-2); for completeness are also shown the one proton
separation energies S(p), defined as the energies of the last occupied orbit.
Probably the most remarkable aspect of B/A is the fact that the calculation follows
the curvature of the mass parabola quite accurately, indicating a good value for the
symmetry energy. In He and 0 [ref. ‘)I we found a similar good agreement for the
dependence of B/A on N. In nuclear matter our force has Esym = 36 MeV (including
the rearrangement part) in excellent agreement with the most recent liquid drop
model fits to nuclear masses 19).
For loOSn, the calculated B/A is 8.235 MeV. We find it to be a good doubly closed
shell nucleus, with no pairing energy. If we correct the calculated value by the average
error between A = 104, 128 (which is 8 keV) we predict B/A = 8.243 MeV. We have
also calculated 90Sn, the mirror nucleus to 9oZr, with N = 40, Z = 50. We find B/A
of only 7.337 MeV. However, this nucleus is unlikely to exist because the last proton
orbit, lg% is unbound. In the calculation it is bound at -0.27 MeV. Looking at the
proton separation energies for the other isotopes, one sees that they follow the ex-
ISOTOPE EFFECTS 551
552 X. CAMP1 et af.

perimental trend closely, although about 1 MeV too large. Hence, the lg, level is
almost certainly unbound for 90Sn.
The variation of the single particle energies, shown in figs. 5a and 5b, is quite
regular. As we remarked for the oxygen isotopes [ref. ‘)I, the nearly constant en and
the rapidly falling ep are consistent with a picture of forces only between unlike
particles. A comparison with experiment would require that the HF energies be de-
duced from experiment, for example, by solving the inverse gap equation (IGE) “).
Unfortunately, this requires an assumption on the form of the pairing interaction.
In ref. “) a single Gaussian force with an exchange mixture was used. Our spectrum
is dilated about 20 y0 more than theirs, but the order of the states is correct. Compared
to fig. 13 of ref. 8), we remark that the lgI-2d, levels do indeed cross between
A = 120 and 122. On the other hand, the lh, level falls more slowly in our result,
and does not cross the 3~ level. All the variations are more regular in our calculation
than in the IGE, which makes us think that the residual interaction has not been
completely eliminated by solving the IGE. For example, if the states really contain a
mixture of three quasiparticles as well as one quasiparticle, the IGE are insufficient.
Finally, the neglect of the tensor force in our model puts some uncertainty in the
calculated single particle energies, so that in any case a closer agreement with the
experimental results would be probably of doubtful significance.

4.2. CHARGE DENSITIES

We have already shown in fig. 3 a sample of our charge densities. The accord with
phenomenological 20) densities is similar to that found in ref. ‘), for closed shell
nuclei: generally very close agreement in the nuclear surface, but with distinctive
oscillations in the interior coming from the shell structure. In fig. 6 our rms charge
radii are compared with those of phenomenological charge densities fitted to electron
scattering data by Ficenec et al. 20). The various lines show that the trend of the radii
is correct, but in all cases they are about 0.02 fm too small. The contributions of the
neutrons and the electromagnetic spin-orbit go in the opposite direction to the vi-
brational contribution; so together they roughly cancel. It is easy to see that the neutron
contribution tends to make the surface sharper and to reduce the charge radius,
because the neutron charge distribution is positive at the centre and negative outside;
hence at the surface of the nucleus there will be a tail of negative charge introduced.
If we suppose that the neutron form factor is of short range compared to the nuclear
dimensions, one can make an expansion leading to

p:(r) = ~p~(r)<r~)+~~‘,V(r)(r~)+ . . ., (4.1)


where the mean square radius of the neutron, -0.12 fm2, was imposed on our
neutron form factor. The leading term is by far the largest and shows that pz will
have a positive region at the shoulder of the neutron density, followed by a negative
region in the tail. This leads to the known “) estimate 6r, = -0.06 N/Zr, for the
change in charge radius.
ISOTOPE EFFECTS 553

112 tl6 1x1 I24 128 A


Fig. 6. Theoretical charge radii compared to experiment [electron scattering ref. 20)].

Compared to the phenomenological charge densities of Ficenec et al. lo), our


densities suffer from having too sharp a corner at the inner surface region, and the
neutron contribution to the charge density accentuates this defect. The spin-orbital
contribution however helps, since (unlike the case of 48Ca) it goes in the opposite
direction and increases the radius. In the case of tin isotopes there are both protons
and neutrons in non spin saturated shells, and their contributions tend to cancel.
Also, the neutrons are filling several levels at once, including the d, and d, levels,
which again give cancelling contributions from the <a * l> factor. The levels which
make an appreciable contribution are the lg, proton, and Id,, lh+, Id% neutron
levels. At 1’6Sn, this effect increases the charge radius by 0.006 fm while at lz4Sn,
where the Id, level is filled, the increase is only 0.003 fm. Compared to 48Ca, the
additional nodes in the wave functions cause further reduction in the effect.
More significant than the comparison with phenomenological densities is the direct
test of our densities against electron scattering and muonic atom data. In fig. 7 the
scattering of 330 MeV electrons on lz4Sn is seen to be qualitatively correct, although
our error in the charge radius entails a poor value of x2. In table 2, the energies of a
number of muonic atom transitions for isotopes and isotones of the tin nuclei are
compared with experiment “). Again, the fact of having a slightly too small radius
results in all our transition energies being 10 keV too large, The usual corrections for
Lamb shift and vacuum polarization were included, whereas the more uncertain
554 X. CAMP1 er al.

20' 30' LO' 50" 60' 7c" 80'

Fig. 7. Differential cross section of elastic electron scattering for lz%n.

nuclear polarization correction 12.5 keV for K, lines, according to Wu and Wilets “‘)I
was omitted (see also ref. ““).
The good overall systematics of the results presented above give us a measure of
confidence in the validity of our model for nuclei in this region of the periodic table.
We now turn to examine fine details, namely the differences in structure between
neighbouring isotopes. The differences in the neutron densities between pairs of even
isotopes are dispiayed in fig. 8b, which shows ~~(~~~~( 112). The three humps reflect
the filhng of the 3s+, lh, and 2d, neutron orbits. Since the neutron radius increases
steadily at the same time, the proton density is pulled outward. The effect on the
proton density is illustrated in fig. 8a, showing p,(A)-p,( 112). The charge density is
similar, with flattened oscillations. Remarkably, the calculation shows a common

TABLE 2
p-mesie transition energies

2Pg + ls+ 2P+ + Is+ 3d$ -+ 2pg 3d+ -+ 2~4


____-._
th. exp. th. exp. th. exp. th. exp.

‘S&i 3271.5 3263.6 3231.7 3223.7 906.0 905.8 940.7 940.6


‘Z&e 3471.7 3464.8 3425.8 3419.1 982.5 982.2 1022.4 1022.1
%W4 3455.1 3444.3 3409.5 3398.9 982.3 981.7 1021.9 1021.4
%‘eT4 3639.1 3629.3 3587.2 3577.6 1061.1 1060.7 1106.1 1105.7

Experimental values quoted from ref. 21).


ISOTOPE EFFECTS 555

-002 0

.OOlO

0
-.OOLO

- 0020

-.003 0

-.004 0

-.005 0

-0060
0. 2.0 4.3 6.0 80 10.0
(fm)
Fig. 8a.

( fm-3)

0. 2.0 4.0 6.0 8.0 IO.@


(fm)
Fig. 8b.

Fig. 8. Isotopic changes in neutron and proton densities; Ap defined as p(A)-~(112).

crossing point dividing the positive from the negative charge regions, reflecting a
linear response of the proton density to the addition of neutrons.
The validity of the charge density differences is tested by comparing with relative
electron scattering cross sections. Since our calculation indicates a regular progression
in 6p,, we show only one example, [o(l16)-o(124)]~[o(116)+cr(124)] in fig. 9. The
solid line is calculated with our densities, while the dashed line represents a best fit
with phenomenological charge densities by Ficenec et at, 20). The points however
are from an older experiment, by Litvinenko et al. ‘“). Out to a momentum transfer
q 5 1.4 fm-’ both curves give similar results.
5.56 X. CAMP1 et af.

4.3. RADIAL MOMENT ANALYSIS


Almost model independent information about the charge density can be extracted
in a rigorous way from elastic electron scattering and muonic data using the method
developed by Friar and Negele ““). Nevertheless the more relevant features of the
isotope shift can be analyzed with the simplest, but also appreciably model indepen-
dent, radial moment analysis method of Ford and Wills “). They showed that each
muonic transition fixes a distinct moment of the charge density. Ford and Rinker 2‘)

Fig. 9. Relative change in cross section between “‘%I and ‘24Sn. Experimental points from ref. 23).
Continuous line: theoretical cross section given by our calculation, dashed line: cross section calculated
with the phenomenological densities of ref. 20).

have shown that electron scattering can be similarly analyzed, serving mainly to
determine the slope of the curve Rk (see below) and that certain moments are fixed with
high accuracy by the muonic data. Following Ford and Wills we define the equivalent
radius for each moment as
R, = [y (rk)]‘h, (4.2)

(rk)

We do not follow the more elaborate suggestion


=: f
Z spc rkd3r.

of Barrett 26) by including an ad-


(4.3)

ditional radial factor u(r) in the definition of (rk). Although less precise, the older
prescription here used gives simpler relations between momenta and more intuitive
properties like rms radius and diffuseness, and makes easier the discussion. The
usual rms radius is 4: R, in this notation.
For a Fermi-like density distribution,

p = p. [I+exp~~)]-i (4.4)
ISOTOPE EFFECTS 551

one finds that

Rk = R l+‘n26 yqZn+...]. (4.5)


The next term is of order (u/R)~“. The slope of a plot of R,

d&t 1 a’*
-_= &_ __
dh:
,I2 R2n-I (4.6)
contains information about the diffuseness of the charge density. Of course, a, R
and IZ are here parameters of a model, and we are simply relating them to quantities
which can be extracted from the experimental data. The best way to compare phenom-
enological models with our model is by translating both of them into equivalent radii
&-
The comparison of rms radii as a function of A has already been given in fig. 6.
In fig. 10, we show the moment analysis for lz4Sn. The curve marked “exp”. is the
moment analysis from the phenomenological density of ref. ‘“). The three curves
for our model show separately the effect of adding in the neutronic plus spin-orbit
contributions, and the vibrational contribution, to the charge density. That our model
gives nearly the correct slope indicates that the diffuseness of the charge density is
good.
The isotope shift in the charge density is reflected in the difference AR, between
lz4Sn and “%, in fig. 11. The curves have the same sense as in the preceeding
figure. The dash dot curve is obtained when only the proton density is used to construct
the charge. Adding the neutronic plus spin-orbital contributions gives the dash line,
finally including the vibrational contribution gives the light solid line. Inclusion of the

Fig. 10. The ‘*%n equivalent radii [eq. (4.2)].


558 X. CAMP1 et al.

“Rk [..------
IX-‘14Stl !_ Exp.(r w:m )
09
I I ----- GO
I
i

_; ~~~

---_ ---_
-05 --__ ---
-----N
---.

Fig. 11. Change in equivalent radii between rzaSn and “%I

Go (proton f. f. only)

112-114 114.116 116-118 118120 lXLl22 122.124

Fig. 12, Changes in mean square radius &<P,~>for pairs of even isotopes. Experimental points from
refs. *O**’).

neutronic contribution greatly improves the slope of the ARk line, and we can conclude
that this contribution is indeed present in the isotope shift. The vibrational correction
gives aiso a similar but less marked effect. From the point of view of R, itself, the
slope is not improved by adding in the I~eutroni~ contribution; probably because it is
ISOTOPE EFFECTS 559

the large proton contribution which is at fault (we have already mentioned that our
renormalization of the force is not quite right). The single experimental point at
li = 2 represents electronic X-ray measurements 27), which measure the mean square
radius. That it does not overlap the phenomenological line may reflect over optimistic
assignment of the error.
In fig. 12, finally, are plotted the changes in charge radius between adjacent pairs of
isotopes. The triangles and dots represent the experimental values based on optical
spectra “) and elastic electron scattering respectively ‘O), the latter having no assigned
error although it would be very large. Here our model is quite successful, reflecting the
general downward trend as A increases. Again the neutronic and spin-orbit con-
tributions are important, improving the slope of the 6 (r2) curve. Adding in the
vibrational correction (dotted line) puts the value too low.

4.4. POLARIZING FIELD

It is useful to relate the isotope shift in the charge density to the properties of the
polarizing field, defined as the change in the equivalent local HF potential ‘) when
particles are added to the nucleus. For the state nrjz it is:

(4.7)
This polarizing field shows directly the regions to which particles will be attracted
preferentially. For the difference of (A = 124)-(A = 116), the neutron and proton
polarizing fields are represented in fig. 13. It is at once evident that the field is essen-
tially the same for all proton states, and that the effect on the neutrons is very small.
The A v&%ons) P eak at the surface appears to be slightly increasing with the energy of
the state. Once again we see evidence that the unlike particle interaction is determinant.

/ -___ J~pol neutrons ‘24Sn_ F%


I - ,rpo, protons

Fig. 13. Polarizing fields in the 1aJSn-“6Sn case.


560 X. CAMPI et al.

The form of the polarizing field can be roughly understood as follows. In a self
consistent theory the one body potential will follow the density. Having in mind
something like a Fermi function, depending on the variable (P - R)/a, we will have

Thus, we expect A VP”’to contain both a volume contribution and a surface peaked
contribution; the latter can be situated inside or outside the half density radius
depending upon the sign of the change in diffusivity. Noting the importance of the
unlike particle interaction, we expect AV, to vanish for neutrons (when no protons
are added) so the volume contribution will vanish. For protons there is a sizeablr
volume-contribution evident in fig. 13.
In fig. 1I we saw that the AR, curve for our model, taking only the proton density
into account, is nearly horizontal, implying a very small reduction of diffusivity for
the proton charge densities. This would imply Aa m 0 in our formula. We see that
the surface peaked term in AV “I is very close to the half potential radius. This is
clear from fig. 14 where the total HF field is drawn on a reduced scale along with the
A VP*‘. By comparing with Ap, on the same figure one sees a correspondence between
dp, and the polarizing field acting on the protons, which is consistent with the ap-

V(MeV)
Fig. 14. Change in neutron density between “%n and lZ4Sn compared with the polarizing fields
for proton states. The total field corresponding to the lga proton state in ‘r6Sn is given for comparison
(right scale)
ISOTOPE EFFECTS 561

proximation made in stating eq. (4.8). Also this means that the polarizing field is
strongly correlated to the single particle energies of the neutron states being filled.
It has been commonly stated that the isotope shift measures the compressibility
modulus of the nuclear medium; however this is seen not to be the case. First of all,
the neutrons and protons behave quite differently, with only one type of particle (the
protons) being pulled outwards by the added neutrons. As we remarked in another
context recently **), the nucleus is much more rigid against compression (or dilata-
tion) of only the protons. Secondly, the polarizing field acting on the protons is not
the quadratic field one would naively use in computing the compressibility, specially
because of this strong surface component. The symmetry energy should play a role
in the different behaviour of neutron and proton fields. As pointed out before, our
symmetry energy is correct, since we reproduce accurately the curvature of the mass
parabola, and also because the corresponding nuclear matter value agrees with those
of empirical mass formulas i9). Th us we expect that the main features of the like
and unlike particle interactions are well reproduced by our calculation.

5. Conchsions
Hartree-Fock calculations using the density dependent force G-O give very good
results for the binding and separation energies of the tin isotopes. The radii are about
0.02 fm too small, but this could be corrected by changing tha renormalization of the
force, to saturate nuclear matter at kr NN1.34 fm-i. This renormalization is required
in any case, to reinstate the charge radii of the ensemble of doubly closed shell nuclei,
now that the neutronic contribution to the charge radii is included. The very close
agreement of the binding energies as a function of neutron number is a sensitive test
for the symmetry energy of our force.
For the isotope shifts in the charge density, we hav? found a good qualitative agree-
ment; but there subsists some lo-20 % of the effect which is not explained within the
framework of our model. This will not be altered by the renormalization of the force.
We noted that the isotope shift depends sensitively upon the single particle energy of
the orbits which are being filled. Since in calculations of this kind it is usually found
that the levels near the Fermi surface are too spread out, this is an obstacle to finding
exactly the correct isotop:: shift. We have noted also that the n~utronic and spin
orbital contributions to the charge density are very important, they reduce the diffu-
sivity and give the good variation of this propsrty from isotope to isotope. Inclusion
of qua~upole vibrations has been found to be important in securing a good diffuseness
for the densities, although the crude way in which they have been introduced leaves
some indetermination on their precise effects. A more detailed calculation would be
required to see if part of the remaining IO-20 7; disagreement in isotope shift is ac-
counted for by these or by some other dynamical contributions.
We would like to thank Dr. Bellicard for providing us with the computer code
ALS 61’7,with which the elastic electron scattering cross sections have been calculated.
562 X. CAMP1 et al.

We are very indebted to Dr. 1. Sick for making electron scattering data available to
us prior to publication. Continued research support from NRC (Canada) under
operating grant A-3198 is gratefully acknowledged by two of us (X. C. and D. W. L. S.).
The other (J. M.) is very indebted to the Division de Physique ThtSorique of the
Institut de Physique Nucl6aire for its warm hospitality and thanks for support under
a grant Joliot-Curie; previous financial support by GIFT {Spain) is also gratefully
acknowledged.

References
1) X. Campi and D. W. L. Sprung, Nucl. Phys. Al94 (1972) 401
2) G. G. Bunatyan and M. A. Mikulinskii, J. NucI. Phys. (USSR) 1 (1965) 38 [Sov. J. Nucl. Phys.
1 (1965) 261;
G. G. Bunatyan, J. Nucl. Phys. (USSR) 4 (1966) 707 [Sov. J. Nucl. Phys. 4 (1967) 5021;
V. P. Krainov and M. A. Mikulinskii, J. Nucl. Phys. (USSR) 4 (1966) 928 [Sov. J. Nucl. Phys.
4 (1967) 6651
3) R. A. Uher and R. A. Sorensen, Nucl. Phys. 86 (1966) 1
4) I. Sick, H. Flocard and M. Vendroni, Phys. Lett. 39B (1972) 443
5) W. Bertozzi, J. Friar, J. Heisenberg and J. W. Negele, Phys. Lett. 41B (1972) 408
6) K. W. Ford and J. G. Wills, Phys. Rev. 185 (1969) 1429
7) J. W. Negele, Phys. Rev. Cl (1970) 1260
8) V. Gillet, B. Giraud and M. Rho, Nucl. Phys. AX03 (1967) 257
9) L. S. Kisslinger and R. A. Sorensen, Mat. Fys. Medd. Dan. Vid. Selsk. 32 (1959) no. 9;
R. Arvieu and M. Veneroni, Compt. Rend. 250 (1960) 992, 2155;
M. Baranger, Phys. Rev. 120 (1960) 9.57;
E. U. Baranger, Adv. Nucl. Phys. 4 (1971) 261, and references quoted therein
10) H. Flocard, A. K. Kerman, P. Quentin and D. Vautherin, IPNO.TH 73-40 preprint Orsay (1973)
I I) P. L. Carson and L. C. McIntyre, Jr., Nucl. Phys. Al98 (1972) 289;
E. 1. Schneid, A. Prakash and B. L. Cohen, Phys. Rev. 156 (f967) 1316
12) P. E. Cavanagh, United Kingdom Atomic Energy Authority Research Group report no. AERE-
R5801, 1968 unpublished
13) D. Vautherin, Phys. Rev. C7 (1973) 296
14) D. Gogny, Proc. lnt. Conf. on nuclear physics, Munich, 1973 tNorth-Holland, Amsterdam, 1973)
p. 48
15) L. W. Owen and Ci. Satchler, Nucl. Phys. 51 (1964) 155
16) D. N. Stacey, Rep. Prog. Phys. 29 (1966) 171
17) P. H. Stetson, F. K. MC. Gowan, R. L. Robinson and W. T. Mimer, Phys. Rev. CZ (1970) 2015
18) A. H. Wapstra and N. B. Gove, Nucl. Data Tables 9 (1971)
19) S. Ludwig, H. von Groote, E. Hiff, A. G. W. Cameron and J. Truran, Nucl. Phys. A203 (1973)
627
20) J. R. Ficenec, L. A. Fajardo, W. P. Trower and I. Sick, Phys. Lett. 42B (1972) 213
21) J. W. Kast, S. Bernow, S. C. Cheng, D. Hitlin, W, Y. Lee, E. R. Macagno, A. M. Rushton and
C. S. Wu, Nucl. Phys. Al69 (1971) 62
22) C. S. Wu and L. Wilets, Ann. Rev. Nucl. Sci. 19 (1969) 527
23) A. S. Litvinenko, N. G. Shevchenko, A, Yu. Buki, G. A. Savitsky, V. M. Khvastunov, A. A.
Khomich, V. N. Polischuk and I. 1. Chkalov, Nucl. Phys. A182 (1972) 265
24) 5. L. Friar and J. W. Negele, NucI. Phys. A212 (1973) 93
25) K. W. Ford and G. A. Rinker Jr., Phys. Rev. C7 (1973) 1206
26) R. C. Barrett, Phys. Lett. 338 (1970) 388
27) J. D. Silver and D. N. S. Stacey, Proc. Roy. Sot. A332 (1973) 139;
R. C. Barrett, Reports on progress in physics, to be published, 1973
28) 3. Martorell and X. Campi, Phys. Lett. 46B (1973) 296
29) S. Galster, H. Klein, J. Moritz, K. H. Schmidt, D. Wegener and J. Bleckwenn, Nucl. Phys. B32
(1971) 221

You might also like