Lecture Notes 2
Lecture Notes 2
2 Nuclear Properties
An atomic nucleus is the small, heavy & central part of the atom, consisting of A nucleons (Z
protons and N neutron), where A is the mass number and Z is the atomic number. Nuclear
size is measured in fermis (femtometres); 1 fm = 10−15 m. The basic properties of the atomic
constituents are given in the table below.
The nucleus is somehow enigmatic and mysterious. Its properties are much more difficult
to characterise than those of macroscopic objects. However, we can describe the nucleus by
parameters like; mass, radius, electric charge, binding energy, and so on. These are considered
in this section, and they form the static properties of the nucleus.
Radius of a nucleus is not a precisely defined quantity. Both the Coulomb potential that binds
the atoms and the resulting electronic charge distribution extend to infinity, although both
become negligibly small at distances far beyond the atomic radius (∼ 10−10 m).
The density of nucleons and the nuclear potential have a similar spatial dependence; relatively
constant over short distances, beyond which they drop rapidly to zero. It is therefore natural to
characterise the nuclear shape with two parameters:
1. The mean radius, where the density is half its central value.
2. The skin thickness, over which the density drops from near its maximum to near its
minimum.
The radius that we measure depends on the kind of experiment we do to measure the nuclear
2 NUCLEAR PROPERTIES 24
To determine the size and shape of an object, we usually examine the radiation scattered by
it. To see the object and its details, the wavelength of the radiation must be smaller than the
dimensions of the object, otherwise diffraction will partially or completely obscure the image.
Beams of electron with energies in the range 100 MeV to 1 GeV can be produced with high-energy
accelerators, and can be analysed with a precise spectrometer to select only those electrons that
are elastically scattered from the selected nuclear target.
~
Let the initial electron wave function be of the form eiki ·~r , appropriate for a free particle of
momentum ~pi = ~~ki . The scattered electron can also be regarded as a free particle of momentum
~
p~f = ~~kf and wave funciton eikf ·~r .
The interaction V (r) converts the initial wave into the scattered wave. The probability for the
transition will be proportional to the square of
Z
F (~ki , ~kf ) = ψf∗ V (r)ψi dv (44)
Z
F (q) = ei~q·~r V (r)dv, (45)
where ~q = ~ki - ~kf , i.e. the momentum change of the scattered electron.
The interaction potential V (r) depends on the nuclear charge density Zeρe (~r′ ), where ~r′ is a
coordinate describing a point in the nuclear volume and ρe is the distribution of nuclear charge.
An electron located at ~r feels a potential energy due to element of charge dQ located at ~r′ given
by
edQ Ze2 ρe (r ′ )dv ′
dV = − = − . (46)
4πε◦ |~r − ~r′ | 4πε◦ |~r − ~r′ |
Then, the complete interaction energy is
Z
Ze2 ρe (r ′ ) ′
V =− dv . (47)
4πε◦ |~r − ~r′ |
2 NUCLEAR PROPERTIES 25
Figure 9: The vector r′ locates an element of charge dQ within the nucleus, and the vector r defines the position
of the electron.
Writing ~q · ~r = qrsinθ in (45) and integrating over r, the properly normalized result is
Z
′
F (~q) = ei~q·~r ρe (~r′)dv ′ (48)
Since we initially assumed that the scattering was elastic, then |~pi | = |~pf |, and q is a function of
the scattering angle (α), between ~pi and p~f . It can be shown that
2p
q= sinα/2,
~
where p is the momentum of the electron. The quantity F (q) is called the form factor, and the
numerical inversion of (49), i.e. inverse Fourier transformation, then gives ρe (r ′ ).
The central nuclear charge density is nearly the same for all nuclei. Nucleons do not seem to
congregate near the centre of the nucleus, but instead have a fairly constant distribution out to
the surface. Thus, the number of nucleons per unit volume is roughly constant, i.e.
A
4πR3
≈ Constant,
3
1
where R is the mean radius. We notice that R ∝ A 3 , i.e.
1
R = R◦ A 3 ,
2 NUCLEAR PROPERTIES 26
We define the skin thickness parameter t as the distance over which the charge density falls from
90% of its central value to 10%. Experimentally, it is ∼ 2.3 fm.
The nuclear charge density can also be examined by studying atomic transitions. Consider the
case of an atom with a single electron, where we assume that the electron feels the Coulomb
attraction of a point nucleus;
Ze2
V (r) = − .
4πε◦ r
Since nuclei are not points, the electron wave function can penetrate to r < R, thus the electron
spends part of its time inside the nuclear charge distribution, where it feels a very different
interaction. As r → 0, V (r) 9 ∞, for a nucleus with non-zero radius.
Assuming that the nucleus is a uniformly charged sphere of radius R, the potential energy of
electron for r 6 R is
′ Ze2 3 1 r 2
V (r) = − − ,
4πε◦ R 2 2 R
while for r > R, the potential energy has the point-nuclear form.
The total energy E of the electron in a state ψn of a point nucleus depends partly on the
expectation value of the potential energy, i.e.
Z
< V >= ψn∗ V ψn dv.
If we assume that changing from a point nucleus to a uniformly charged spherical nucleus does
not significantly change the electronic wave function ψn , then the energy E ′ of the electron in a
state of uniformily charged spherical nucleus depends on the expectation value of the potential
V ′ , i.e. Z Z
′
< V >= ψn∗ V ψn dv + ψn∗ V ′ ψn dv,
r<R r>R
where the second integral involves only the 1/r potential energy. The effect of the spherical
nucleus is thus to change the energy of the electronic states, relative to the point-nucleus value
by,
△ E = E′ − E
i.e. following from the assumption that the wave functions are the same, so the K. E. terms in
E ′ and E are identical.
2 Z 4 e2 R2
△E= . (51)
5 4πε◦ a3◦
We can measure and compare the K X-ray energies (resulting from 2p to 1s transitions) in two
neighbouring isotopes of mass numbers A and A′ . Let EK (A) and EK (A′ ) represent the observed
K X-ray energies. Then
EK (A) − EK (A′ ) = E2p (A) − E1s (A) − E2p (A′ ) + E1s (A′ ).
If we assume that the 2p energy difference is negligible, i.e. p-electron wave functions vanish at
r = 0, the remaining 1s energy difference reduces to the difference between the △ E values of
(51) because
E1s ≡ E ′ = E△ E,
and the point-nucleus values E would be the same for both isotopes. Thus,
2 Z 4 e2 1
EK (A) − EK (A′ ) =△ E(A′ )− △ E(A) = − R◦ (A2/3 − A′2/3 ),
5 4πε◦ a3◦
It is also possible to measure isotope shifts for the optical radiations in atoms, i.e. the transition
among the other electronic shells that produce visible light. Since these orbits lie much further
from the nucleus than the 1s orbit, their wave functions integrated over the nuclear volume
give much smaller shifts. The s states give non-zero limits on the wave function at small r. If
the optical transitions involve s states, the isotope shifts can be large enough to be measured
precisely using modern techniques of laser interferrometry.
2 NUCLEAR PROPERTIES 28
The effects of the nuclear size on X-ray and optical transitions are very small, i.e. ∼ 10−4 to
10−6 of the transition energy. This is due to the difference in scale of 104 between the Bohr
radius a◦ and the nuclear radius R.
For integrals of the form in (50) to give large effects, the atomic wave function should be large
for r near R, but instead the atomic wave functions are large near r = a◦ /Z ≫ R. This situation
can be improved by using a muonic atom.
The muon is a particle identical to the electron in all its characteristics, except its mass which is
207 times the electronic mass. Since the Bohr radius depends inversely on the mass, the muonic
1
orbits have 207
the radius of the corresponding electronic orbits. Thus, for the heavier nuclei like
Pb, the muonic 1s orbit has its mean radius inside the nuclear R. The effect of the nuclear size
is a factor of 2 in the transition energies, which is a considerable improvement over the factor of
∼ 10−4 to 10−6 in electronic transitions.
Muons are not present in ordinary matter, but are made artificially using large accelerators that
produce intense beams of π muons, which then undergo rapid decay (∼ 10−8 s) to muons. Beams
of the resulting muons are then focused onto a suitable chosen target, where the atoms of the
target capture the muons into orbits similar to the electronic orbits. Initially, the muon is in a
state of very high principal quantum number (n), but cascades down toward its 1s ground state,
emitting photons in analogy with electronic transitions between energy levels.
The muonic energy levels and transition energies are 207 times their counterparts (electronic).
Since ordinary K X-rays are in the energy range of tens of keV, muonic K X-rays will have energies
of a few MeV. The isotope shift of electronic K X-rays. We can use the observed muonic X-ray
energies directly to measure/compute the parameters of the nuclear charge distribution.
Another way to determine the nuclear charge radius is from direct measurement of the Coulomb
energy differences of nuclei. For example, consider 31 H2 and 32 He1 . To move from 3 He to 3 H,
we must change a proton into a neutron, which changes only the Coulomb energy, but not the
nuclear energy, because the two protons in the 3 He experience a repulsion that is not present in
3
H.
2 NUCLEAR PROPERTIES 29
The formula in the Coulomb repulsion energy can be used to calculate the distance between
the protons and thus the size of the nucleus. In this consideration, Z of the first nucleus must
be equal to N of the second nucleus and vice versa. Such pairs of nuclei are called “Mirror
Nuclei”because one is changed into the other by reflecting in a mirror that exchanges protons
13 13 39 39
and neutrons. Examples of pairs of mirror nuclei are 7 N6 and 6 C7 , or 20 Ca19 and 19 K20 .
1. One of the nuclei can decay to the other through β-decay, in which a proton changes to a
neutron with the emission of a positron. The maximum energy of the positron is a measure
of the energy difference between the nuclei.
11
2. Through nuclear reactions; for example, when a nucleus such as B is bombarded with
11
protons, a neutron will be emitted leaving C. The maximum proton energy required to
cause the reaction is a measure of the energy difference between the two nuclei (11 B and
11
C).
Although the measurements of the nuclear charge radius use very different techniques, they all
give the same results, i.e. the nuclear radius (R) varies with mass number (A), as
1
R = R◦ A 3 ,
Any experiment that involves nuclear force between two nuclei will often provide a measure of
the nuclear radius. In this case, the radius is characteristic of the nuclear force, rather than
Coulomb force. The radius therefore reflects the distribution of all the nucleons in the nucleus,
and not only protons.
As another example, we can consider a form a radioactive decay in which an α-particle is emitted
from the nucleus. The α-particle must escape the nuclear potential and penetrate a Coulomb
potential barrier. The α-decay probabilities can be calculated using the Schrödinger equation
(the standard barrier-penetration approach). The values depend on the nuclear matter radius
R, and comparisons with measurements enable values of R to be deduced.
A third approach to determine the nuclear matter radius is the measurement of the energy of the
π-mesic X-rays, similar to the muonic X-ray technique discussed above. The difference though
results from differences between muons and π-mesons, i.e. the muons interact with the nucleus
through the Coulomb force, while the π-mesons interact with the nucleus through the nuclear
and Coulomb forces.
Like the muons, the negatively charged π-mesons cascade through electronlike orbits and emit
photons (π-mesic X-rays). When the π-meson wave functions begin to overlap with the nucleus,
the energy levels are shifted somewhat from values calculated using only the Coulomb interaction.
In addition, the π-mesons can be directly absorbed into the nucleus, especially from the inner
orbits. Thus, there are fewer X-ray transitions among the inner levels. The disappearance rate
of π-mesons gives another way to determine the nuclear radius.
All these effects could principally be used to deduce the nuclear radius. However, the calculations
2 NUCLEAR PROPERTIES 31
are very sensitive to the exact onset of the overlap between the probe particle and the nuclear
matter distribution. It would be very wrong to assume and use a uniform sphere model, where
density is considered constant out to radius R and zero beyond R. We instead use a distribution
with a proper tail beyond the mean radius.
In general, the charge and matter radii of nuclei are nearly the same, to within ∼ 0.1 fm. Both
show the A1/3 dependence with R◦ ≈ 1.2 fm. Because heavy nuclei have about 50% more
neutrons than protons, one might expect the neutron radius to be somewhat larger than the
proton radius. However, the proton repulsion tends to push the proton outward and the neutron-
proton force tends to pull the neutron inward, until the protons and neutrons are completely
intermixed that the charge and matter radius are nearly equal.
The mass energy mN c2 of a nuclide is its atomic mass energy mA c2 less total mass energy of Z
electrons and the total electronic binding energy. Thus,
Z
X
mN c2 = mA c2 − Zme c2 + Bi , (52)
i=1
where Bi is the binding energy of the ith electron. Electronic binding energies are of orders 10 -
100 keV in heavy atoms, while atomic mass energies are of orders of A × 1000 MeV. To a good
precision, the last term in (52) can be neglected. In nuclear physics, working with differences in
mass energies, the effects of electronic binding energies tend to cancel.
The binding energy (B) of a nucleus is defined as the difference in mass energy between the
nucleus A
Z XN and its constituent Z protons and N neutrons, i.e.
A
B = Zmp + Nmn − m X − Zme c2 .
By grouping the Z proton and electron masses into Z neutral hydrogen atoms, we can write
1
A
B = Zm H + Nmn − m X c2 .
Masses are generally given in atomic mass units (u). It is therefore convenient to include a
conversion factor for c2 , i.e. c2 = 931.50 MeV/u. Occasionally, we find atomic mass tables
in which the mass defect △= (m − A)c2 , is given instead of m A X . Nevertheless, the above
2 NUCLEAR PROPERTIES 32
equation can be used to deduce the atomic mass. Other useful and interesting properties that
are usually tabulated are the neutron and proton separation energies.
The neutron separation energy (Sn ) defines the amount of energy needed to remove a neutron
A−1
from a nucleus A A
Z XN . It is equal to the difference in binding energies between Z XN and Z XN −1 ,
i.e.
Sn = B A
Z XN − B A−1
Z X N −1
A−1 2
= m Z XN −1 − m A Z XN + mn c .
Similarly, the proton separation energy (Sp ) is the energy needed to remove a proton from a
nucleus A
Z XN , i.e.
Sp = B A A−1
Z XN − B Z−1 XN
A−1
= m Z−1 XN − m A Z XN + m
1
H c2 ,
where the hydrogen mass appears instead of proton mass, since we always work with atomic
masses. In both cases of (Sn ) and (Sp ), the Z electron masses cancel out.
As with many other nuclear properties, we gain valuable clues to nuclear structure from a
systematic study of nuclear binding energy. Binding energy increases more or less linearly with
A, i.e. B/A can also be plotted as a function of A, which shows a relatively constant curve
except for very light nuclei, as shown in Figure 10.
The average binding energy for most nuclei is about 8 MeV per nucleon. The curve peaks at
about A = 60, where the nuclei are most tightly bound. This suggests that we can gain energy
in two ways:
In either cases, we climb the curve of binding energy and liberate binding energy.
The positive binding energy is entirely contributed by the strong nuclear interaction, diminished
somewhat by surface effects and by the effects of a difference between N and Z. However,
2 NUCLEAR PROPERTIES 33
Figure 10: Variation of binding energy per nucleon with nucleon number.
the effect that actually causes B(Z, A)/A to decline above A ≈ 60 is the increasing Coulomb
repulsion among the nucleons. Because the strong force acts only between nearest neighbors, it
increases only as A (not A2 ), while the electrostatic repulsion increases about as Z 2 , and thus
becomes more important as A and Z increase.
The fact that the binding energy per nucleon B=A decreases beyond A ≈ 60 means that decays
in which the nucleus breaks into two smaller pieces may be able to release energy, and thus
become possible. Of such decays, alpha-instability is the most common, because the decay leads
to ejection of a tightly bound 42 He nucleus, making the 28.3 MeV binding energy of this nucleus
available. All nuclei of A > 165 can release energy by ejecting an alpha-particle, but the rate is
often so slow as to be insignicant even over the age of the universe.
Attempting to understand the curve of binding energy leads to the semi-empirical mass formula.
We require to include a constant term in estimating B/A. Since (to lowest order) B ∝ A, then
B = av A, (53)
where av = a constant of order of 8 MeV. That is, the contribution to the binding energy from
the volume. If every nucleon attracts all others, then the binding energy would be proportional
2 NUCLEAR PROPERTIES 34
to A(A − 1) ∼ A2 . But, since B varies linearly with A, this suggest that each nucleon attracts
only its nearest neighbours, and not all of the other nucleons.
From scattering theory, we know that the nuclear density is roughly constant. Thus, each nucleon
has about the same number of neighbours, and so contributes roughly the same amount to the
binding energy. However, the nucleon on the nuclear surface is surrounded by fewer neighbours
and thus less tightly bound than those in the central region. Such nucleons on the surface do
not contribute to B quite as much as those in the centre. In this, (53) is an approximation, thus
we subtract from B a term which is proportional to the nuclear surface area, where the surface
area is proportional to R2 or to A2/3 . Thus, the surface nucleons contribute to B a term of the
form - As A2/3 .
The binding energy formula must also include the Coulomb repulsion of the protons, which
tends to make the nucleus less tightly bound . Since each proton repels every other protons,
the Coulomb repulsion term is proportional to Z(Z − 1), and assuming a uniformly charged
2
sphere, we obtain - 32 4πεe◦ R◦ Z(Z − 1)/A1/3 . The negative sign implies a reduction in B. The
constants give 0.72 MeV with R◦ = 1.2 fm. We can generally use a Coulomb constant ac to give
1
-ac Z(Z − 1)A− 3 .
From the study of the distribution of stable and radioactive isotopes, we note that stable nuclei
have Z ≈ A/2. So, the binding energy must take into account stable nuclei, especially for light
nuclei. The possible form (symmetry term) is asym (A − 2Z)2 /A, tends to make the nucleus
symmetric in protons and neutrons. However, the term becomes less important for heavy nuclei,
because the rapid increase in the Coulomb repulsion requires additional neutrons for nuclear
stability.
Finally, we also have to include a term that accounts for the tendency of like nucleons to couple
pairwise to stable configurations. When we have an odd number of nucleons (e.g. odd Z and
even N, and vice versa), this term does not contribute to B. However, when both Z and N are
odd, we gain binding energy by converting one of the odd protons into a neutron (or vice versa)
so that it can form a pair with its formerly odd partner. There are only few nuclei in nature
with odd N and Z, i.e. 2 H, 6 Li, 10
B and 14
N, but there are 167 nuclei with even N and Z. The
pairing energy δ is usually expressed as +ap A−3/4 for N and Z even, and zero for A(= Z + N)
2 NUCLEAR PROPERTIES 35
odd.
There are therefore five terms contributing to the binding energy. Thus, we have
The constants must be adjusted to give the best agreement with the experimental curve shown
above. Using the formula for B, we then have the semi-empirical mass formula;
For constant A, the semi-empirical mass formula represents a parabola when M is plotted Vs
Z, centred about the point where (57) reaches a minimum, (where ∂M/∂Z = 0), i.e.
With ac = 0.72 MeV, asym = 23 MeV, it follows that the first two terms of the numerator are
negligible. Thus,
A 1
Zmin ≈ .
2 1 + 41 A2/3 ac /asym
And for small A, Zmin ≈ A/2, but for large A, Zmin < A/2.
that we have also the lower cases l, s and j. To the extent that the nuclear potential is central,
l, s and j will be constants of the motion.
In quantum mechanical sense, we can label every nucleon with the quantum numbers l, s and j.
The total angular momentum of a nucleus containing A nucleons would then be the vector sum
of the angular momenta of all the nucleons. It is usually called “Nuclear Spin”and is represented
~
by the symbol I.
For many applications, the nucleus behaves as if it were a single entity with an intrinsic angular
~ In an ordinary magnetic field, we can observe the nuclear Zeeman effect as the
momentum I.
state I splits up into its 2I + 1 individual substates, m = −I, −I + 1, · · · , I − 1, I. The substates
are equally spaced.
2 NUCLEAR PROPERTIES 36
If we could apply and incredibly strong magnetic field that would break the coupling between
the nucleons, we would see each individual j splitting into 2j + 1 substates. However, no fields
of sufficient strength to break the coupling of the nucleons can be produced. We therefore
observe the behaviour of I as if the nucleus were only a single spinning particle. Thus, the spin
(total angular momentum) I~ and the corresponding spin quantum number I are used to describe
nuclear states.
Normally, I is used to denote the nuclear spin and j to represent the total angular momentum
of a single nucleon. Often, a single valence particle determines all of the nuclear properties, so
that I = j. In other cases, it may be necessary to consider two valence particles in which
I~ = ~j1 + ~j2 ,
and several different resultant values of I may be possible. At times, the odd particles and the
remaining core of nucleons each contributes to the angular momentum, so
I~ = ~jparticle + ~jcore .
The restriction on the allowed value of I comes from consideration of the possible z-components
of the total angular momentum of the nucleons. Each j must be half-integral ( 21 , 23 , 25 , · · · ), and
thus its only possible z-components are likely half-integral (± 21 ~, ± 32 ~, ± 52 ~, · · · ).
If we have an even number of nucleons, there will be an even number of half-integral components,
with the result that the z-component of the total I can only take on integral values. This requires
that I itself be an integer. If there are odd number of nucleons, the total z-component must be
half-integral and so must the total I. The rules are thus:
The measured values of nuclear spin can tell a great deal about the nuclear structure. For
example, of the known stable and radioactive even-Z and even-N nuclei, all have spin-0 ground
states. This is evidence for the nuclear pairing force, i.e. the nucleons couple together in spin-0
pairs, giving a total I of zero. However, the ground state spin of an odd-A nucleus must be
equal to the j of the odd proton or neutron.
2 NUCLEAR PROPERTIES 37
We may get more information about the nuclear energy levels in a nucleus by solving the
Schrödinger equation for the well in three dimensions. For simplicity we assume again an
infinitely deep well, so that the wave function of a proton or neutron is completely confined
within a region of radius R around the origin of coordinates, and goes to zero on the boundary.
It is assumed that the mean potential in which each nucleon moves is spherically symmetric, and
it is known that the solution is separable. As before, we have Ψ(r, θ, φ) = ul (r)Ylm (θ, φ), with
the spherical harmonics of definite angular momentum quantum numbers l and ml describing
the angular variation, and the radial wave function ul (r) satisfying
~2 1 d 2 ~ l(l + 1)
− 2
(rul ) + ul = Eul .
2mn r dr 2mn r 2
The boundary conditions are that ul (0) is finite and ul (R) = 0. For l = 0 it is easy to verify
that the solution is
sinkr
us (r) = ,
kr
where E = ~2 k 2 /2mn . To satisfy the boundary condition at R, k must satisfy k = kn,s = nπ/R.
This condition defines a series of l = 0 (s-state) energy levels of energies
~2 nπ 2
E(n, s) = .
2mn R2
~2 1 d 2 ~ 1
− (ru p ) + up = Eup ,
2mn r dr 2 mn r 2
sinkr coskr
up (r) = − .
(kr)2 kr
Again the boundary condition at r = R requires up (R) = 0. This condition is satisfied for a
series of k values; the first few zeros of up (R) are at kn,pR = 4.49, 7.73, 10.90, . . . For each kn,p ,
there is a corresponding energy level E(n, p), interleaved with the energy levels of the s-states.
In general, the solutions of the Schroedinger equation for arbitrary l are the spherical Bessel
functions,
l
l 1 d sinkr
jl (kr) = (−kr) .
k 2 r dr kr
2 NUCLEAR PROPERTIES 38
For each l, acceptable values of k are those which make jl (kR) = 0. These zeros of the solution
define a series of energy levels, interleaved with the levels of other values of l. In nuclear physics,
we number these levels sequentially for each l, so the lowest few energy levels are 1s, 1p, 1d, 2s,
1f, 2p, 1g, 2d, 1h, 3s, etc. With degeneracy with respect to ml and spin, there are 2(2l + 1)
states of the same energy in each level (i.e. this number of protons and of neutrons may occupy
one energy level). Thus, for this simple model, we find a series of energy levels, each of which
can be occupied by only a limited number of protons and neutrons, and each of which has
definite angular momentum. The situation is very reminiscent of atomic orbitals. This model
can be improved in obvious ways, by making the well finite in depth rather than infinite, and by
softening the outer boundary somewhat. It is found that these improvements lower the energy
levels (relative to the ground state) somewhat, compared to the simple well with infinite walls,
but do not change the order of single-nucleon levels much.
Alongside nuclear spin, parity is also used to label nuclear states. The parity can take either +
(even) or - (odd) values. If the wave function of every nucleon were known, one would determine
nuclear parity by multiplying together the parities of each of the A nucleons, ending with a result
π either + or -, i.e.
π = π1 π2 · · · πA .
However, in practice, no such procedure is possible because, generally, we cannot assign a definite
wave function of known parity to every nucleon. Like the spin, parity is regarded as an overall
property of the whole nucleus. It can be directly measured using a variety of techniques of
nuclear decays and reactions. The parity is denoted by a + or - superscript to the nuclear spin
1− 5+
as I π . For example, 0+ , 2− , 2
, 2 , and so on. There is no direct theoretical relationship
between I and π. For any value of I, it is possible to have either π = + or π = -.
Quantum states can have some kind of space symmetry. Consider the parity operator P, which
reflects space coordinates through the coordinate origin:
Pf (r) = f (−r).
It is Hermitian operator, and hence has real eigenvalues, so it may be observable. We can easily
find its eigenvalues
PΨα (r) = αΨα (r),
2 NUCLEAR PROPERTIES 39
since
PPf (r) = Pf (−r) == f (r),
so central force wave functions have the parity of l: even for even l and odd for odd l.
Much of the knowledge of nuclear structure comes from studying the weaker electromagnetic
interaction, and not the strong nuclear interaction. The strong nuclear interaction establishes the
distribution and motion of nucleons in the nucleus, which can be probed with the electromagnetic
interaction. We use electromagnetic fields that have less effects on the motion of nucleons than
the strong force, thus the measurements do not distort the object under study.
Any distribution of electric charges and currents produces electric and magnetic fields that vary
with distance. We assign the charge and current distribution an electromagnetic multipole
moment associated with each characteristic spatial dependence, namely;
The 1
r2
electric field arising from the net charge - the zeroth or monoploe moment.
The 1
r3
electric field arising from the first or dipole moment.
The 1
r4
electric field arising from the second or quadrupole moment, and so on.
The magnetic multipole moments behave similarly save monopole moments, which either do not
1
exist or are exceedingly rare, hence the monopole moment field (∝ r2
) does not contribute.
2 NUCLEAR PROPERTIES 40
The simplest distributions of charges and currents give only the lowest order multiple fields. On
the other hand, a spherical charge distribution gives only monopole (Coulomb) field, and the
higher order fields vanish. It is therefore necessary to measure (or calculate) only the lowest
order mutipole moments to characterise the electromagnetic properties of the nucleus.
Another restriction on the multipole moments comes from the symmetry of the nucleus. Each
electromagnetic multipole moment has a parity determined by the multipole operator when
r −→ −r. As a hint, in the nuclear regime, we can treat the multipole moments in the operator
form (recall quatum mechanics) and calculate the expectation values for various nuclear states.
The parity of electric moments is (−1)L , where L is the order of the moments, i.e. L = 0
(monopole), L = 1 (dipole), L = 2 (quadrupole), and so on. For magnetic moments, the parity
is (−1)L+1 .
When we compute the expectation value of a moment, we evaluate an integral of the form
R ∗
ψ Ôψ, where Ô is the electromagnetic operator. If Ô has odd parity, then the integrand is an
odd function of the coordinates and must vanish identically. Thus, all odd-parity static multipole
moments must vanish, i.e. electric dipole, magnetic quadrupole, electric octupole (L = 3), and
so on.
We study nuclear structure in part through the properties of nuclear excited states. Like atomic
excited states, the nuclear excited states are unstable and decay rapidly to ground state. We
can obtain nuclear excited states by moving nucleons to higher order energy states, hence the
orbits of individual nucleons can be revealed. Spectroscopy partly aims at observing the possible
excited states and to measure their properties.
Among the properties to measure are energy of excitation, life and models of decay, spin and
parity, magnetic dipole moment, and electric quadrupole moment. It is only through experiments
and calculations involving use of powerful computers that we can be able to obtain detailed
interpretations of nuclear structure.
2 NUCLEAR PROPERTIES 41
Many nuclei are observed to be unstable; they spontaneously change into other nuclei by emission
of one or more particles. The two common kinds of decay of naturally occurring nuclei are:
137 −
55 Cs −→137
56 Ba + e + ν¯e
235
92 U −→231
90 Th + α
Such decays occur when the nucleus in question can go to a state of lower energy by the decay,
and enough energy is available from the transition to supply what is needed for the decay (e.g
to create an electron and a neutrino).
In beta-decay, no nucleons are lost by the nucleus, so A stays constant, while Z → Z ± 1 and
N → N ∓ 1. The binding energy formula is accurate enough to predict correctly (in almost all
cases) which nuclei are stable against beta-decay.Since the ν is (nearly) massless, beta-decay is
possible if
mn (Z, A) > mn (Z + 1, A) + me .
Adding Zme to each side, this may be expressed using atomic masses as
If Z > Zmin (the actual Z of lowest energy), decay can occur by positron emission if
mn (Z, A) > mn (Z − 1, A) + me , or
In an atom, another process that competes with positron emission is electron capture or K-
capture. The nucleus absorbs an electron from its cloud (usually from the K shell), converting
a proton to a neutron and emitting a neutrino.