Vous êtes sur la page 1sur 41

Nuclear & Particle Physics Prof. T.J.

Sumner 2001/2

Nuclear Physics

[Last Update 14/11/2001]

1. Introduction ...........................................................................................................3
1.1 Course Information...........................................................................................3
1.1.1 Book list .....................................................................................................3
1.1.2 Office hours and contact details.................................................................3
1.1.3 Problem sheets ...........................................................................................3
1.2 What is nuclear physics? ..................................................................................3
1.3 Review of atomic and nuclear parameters........................................................4
2. Nuclear Properties.................................................................................................4
2.1 Size ...................................................................................................................4
2.2 Composition .....................................................................................................4
2.3 Mass and charge density...................................................................................5
2.4 Binding energy..................................................................................................5
2.5 Internal Structure ..............................................................................................7
3. Semi-Empirical Mass Formula ............................................................................8
(a) Energy Equation ...............................................................................................8
(i) Volume energy, Ev ............................................................................................8
(ii) Surface tension ..............................................................................................9
(iii) Electrostatic.................................................................................................10
(iv) Asymmetry term..........................................................................................11
(v) Pairing term .................................................................................................13
(b) SEMF summary..............................................................................................14
(c) Magic numbers....................................................................................................14
4. Radio-Active Decay .............................................................................................15
(a) Beta-decay – Q value......................................................................................15
(b) Types of decay................................................................................................17
(i) Beta-decay ......................................................................................................17
(ii) Positron-decay.............................................................................................18
(iii) Electron capture...........................................................................................18
(iv) Nuclear fission.............................................................................................18
(c) Weak decays ...................................................................................................18
(i) Mass parabolae from SEMF ...........................................................................18
(d) Alpha decay ....................................................................................................20
(i) SEMF..............................................................................................................20
(ii) Theory .........................................................................................................21
(iii) Observation .................................................................................................24
(iv) Decay series.................................................................................................25
5. Shell Model ..........................................................................................................27
(a) Magic numbers ...............................................................................................28
(i) Atoms..............................................................................................................28
(ii) Nuclei ..........................................................................................................28
(b) Energy levels and potentials ...........................................................................28
(c) Spin-orbit coupling .........................................................................................31
(d) Nuclear spin....................................................................................................31
(e) Parity...............................................................................................................32
(f) Magnetic Dipole Moments ................................................................................32
6. Formation of Nuclei ............................................................................................33
(a) Big-Bang nucleosynthesis ..............................................................................33
(b) Stellar nucleosynthesis ...................................................................................35
(i) Main Seqeunce ...............................................................................................35
(ii) Red Giants ...................................................................................................36
(c) Supernova explosions .....................................................................................36
(i) s-process .........................................................................................................37
(ii) r-process ......................................................................................................38
7. Nuclear Reactors .................................................................................................38
(a) Fission reactors ...............................................................................................38
(b) Fusion reactors................................................................................................40
1. Introduction
1.1Course Information
These notes on Nuclear Physics constitute about half of the ‘Nuclear and Particle
Physics’ lecture course.

1.1.1 Book list


The main recommended book, which contains most of the required material at the
right level, is Nuclear & Particle Physics , by Williams. Other useful books, which
also cover the material, and more, are Introductory Nuclear Physics , by Krane,
Nuclei and Particles , by E. Segré.

1.1.2 Office hours and contact details


During term time two hours will be kept as formal office hours on Thursdays between
15.00 and 17.00. Other times are possible but need to be agreed individually, subject
to availability.
The office room number is 1111 in the Blackett Laboratory within the Astrophysics
Group. The internal phone number is 47552, and the email address is
t.sumner@ic.ac.uk

Other staff associated with this course are Dr. Jordan Nash, who delivers the Particle
Physics section of this course, and Dr. Paul Dauncey and Prof. Peter Dornan who are
course associates for the whole course. Help and advice can also be sought from
these if required. Their contact details are
Dr. Jordan Nash: room 524, telephone 47804, and email j.nash@ic.ac.uk
Dr. Paul Dauncey: room 524, telephone 47803, and email p.dauncey@ic.ac.uk
Prof. Peter Dornan: room 538, telephone 47882, and email p.dornan@ic.ac.uk

1.1.3 Problem sheets


There will be three problem sheets with answers for this part of the course. The
problem sheets are designed to consolidate some of the lecture material and, in places,
introduce extensions to it. The questions on the problem sheets should not be taken to
be indicative of examination questions in style or complexity. Previous examination
papers are available from the Undergraduate Office.

1.2What is nuclear physics?


Nuclear Physics is the study of how physics needs to be applied within the atomic
nucleus in order to understand how nuclei exist, what their compositions are, what
their properties are and how they behave. The existence of nuclei depends on the
presence of a very short range force, called the strong force. A large range of nuclei
is known to exist with differing compositions. However only certain compositions
seem to be allowed. Properties of nuclei that need to be understood are their size,
charge, mass, angular momentum and magnetic dipole moments. Nuclei are not
necessarily passive objects and ‘behave’ with differing levels of stability, can exhibit
a number of decay processes and can take part in nuclear interactions.
As well as allowing us to understand atomic nuclei, Nuclear Physics shows us how to
exploit the behaviour of nuclei in practical uses, such as power generation in nuclear
reactors, and how to understand the formation of elements within the Universe, with
close connections to cosmology and astrophysics.

1.3Review of atomic and nuclear parameters


Atoms are composed of light negatively charged electrons in orbitals around a
massive central tiny positively charged nucleus. Typically the mass of the nucleus is
mnucleus ≈ 2 * Z * 2000 * me , where Z is the number of electrons and me is the mass of
the electron. The electrons are held in place by electrostatic forces and quantum
mechanical effects. As the nucleus is so small the electrostatic force seen by each
electron is more or less central. Quantum mechanics dictates that the electron
energies fall into discrete ‘levels’. Atomic radii are characterised by the so-called
2
Bohr radius, ao = h ≈ 5.3 × 10 −11 m . Typical energy level values can be
me c 2
estimated from the uncertainty relationship between position and momentum,
2
∆p∆x ≈ h . Putting ∆x ≈ a o gives ε e ≈ h ≈ few eV [1eV = 1.6x10-19 J].
(2a o ) me
2

The particle/wave connection applied to the electron gives a rest mass electron scale
size (wavelength) of λ e ≈ h ≈ 1.3 × 10 −12 m .
me c
Nuclei are composed of a collection of quasi-equal mass particles (nucleons) held
together in a spherical distribution by mutual attraction. Nucleons come in two
varieties, neutrons and protons. The protons provide the positive charge of the
nucleus and the mutual attraction must be strong enough to overcome the Coulomb
repulsion between the protons. The mutual attraction is a two body interaction
between pairs of nucleons and is so short range it does not necessarily extend across
the whole nucleus. Hence the attractive force seen by any particular nucleon is not
central. Thinking about the nucleus as ‘particles in a box’ suggests quantum
mechanical effects will be present. Nuclear radii are characterised by,
ro ≈ 1.3 × 10 −15 m . Typical nucleon ‘energy level’ values can again be estimated from
the uncertainty relationship between position and momentum, this time by putting
2
∆x ≈ ro giving ε n ≈ h ≈ few MeV .
(2ro )2 me

2. Nuclear Properties
2.1 Size
Nuclear radii, R, are seen to increase with the total number of nucleons, A, as
1
R = ro A 3 (1) [R]
This implies that the nuclear volume simply scales with the total number of nucleons.
Each nucleon occupies a similar volume independent of its ‘position’ within the
nucleus.

2.2 Composition
A nucleus with A nucleons, comprised of N neutrons and Z=(A-N) protons has a mass
which will be denoted by M(A,Z). A is the atomic mass number, not to be confused
with Z which is the atomic number, i.e. the number of protons and electrons. The
atomic mass number is simply the sum of the numbers of neutrons and protons,
A= N +Z (2) [R]

2.3 Mass and charge density


The mass density, ρ, of a nucleus is its mass divided by its volume. As neutrons and
protons have equal masses to within 0.14% the mass density is
M 3 Amn 3mn
ρ= ≈ =
V 4πro A 4πro3
3
(3) [R/D]
= constant = 1.82 × 1017 kg/m 3
where mn is the mass of the neutron. This implies the nuclear density does not
depend on the atomic mass number.
Similarly a charge density, ρc, can be defined as
Q 3Ze
ρc = ≈ (4) [R]
V 4πro3 A
where e is the charge on the electron. It will be seen later that stable nuclei have
N~Z~A/2 (figure 1). This means the charge density is also approximately constant.
Q 3Ze 3e
ρc = ≈ ≈ C/m 3 (5) [D]
V 4πro A 8πro3
3

2.4 Binding energy


Properties of nuclei, such as stability and decay schemes are determined by energy
considerations. For a given value of A the most stable isotopes will have the lowest
overall energy, which will manifest itself as the lowest rest mass energy. The
difference between the rest mass energies of the individual nucleons and the overall
nucleus is accounted for by the interaction energy between the nucleons, which is
holding the nucleus together. This ‘interaction’ energy is referred to as the binding
energy, BE, and is given by
BE = Zm p c 2 + ( A − Z )mn c 2 − M (A, Z )c 2 (6) [R]
where mp is the mass of the proton. As noted earlier the proton and neutron rest
masses are very similar, but not equal, and can be expressed in terms of the atomic
mass unit as:-
mn = 1.00866 amu
| 1 amu = 1.6605 × 10 −27 kg (7) [G]
m p = 1.00728 amu
Figure 2 shows the binding energy per nucleon, which is the more usual way in which
the binding energy is shown.
M (A, Z )c 2
2
BE Zm p c
= + (1 − Z )mn c 2 − (8) [R]
A A A A
In figure 2 the open circles denote actual measured values of the binding energy per
nucleon.
Points to note from figure 2 are:-
• The scale coverage on the vertical axis is an expanded range which only spans
from 7.4<BE/A<8.8 MeV/nucleon for 10<A<240. BE/A is almost a constant!
• The observation data follow a smoothish curve but with some features at
particular values of A i.e. there are obvious bumps at A ~ 95, 140, 205. These
‘features’ point to internal structure ⇒ shell orbitals ⇒ quantum effects ⇒ magic
numbers (see later)
• The behaviour of BE/A shows how the nuclear energy varies with A and hence will
determine which nuclei are stable and which are likely to exhibit radioactive
decay
• The solid dots show predictions of a semi-empirical mass formula for M(A,Z).
This is not rigorously founded physical theory, but is based on a simple starting
model with successively finer corrections. This will be the subject of the next
chapter.

2.5 Internal Structure


Many experiments show clear evidence for regular internal structure with nuclei.
Figures 3a and 3b show two such observations. The first is for proton scattering from
a nucleus (in this case 10B) as a function of proton energy. The rate of scattering
events shows resonant enhancements at specific proton energies. The second shows
γ-ray spectroscopy measurements of 238U nuclei after bombardment with 136Xe
nuclei. The γ-rays are only emitted at certain energies and there is a regularity to the
sequence.
These types of experiment provide clear evidence for detailed energy level structures
for both the neutrons and protons within the nuclei. This suggests the system is
probably quantum mechanical with discrete energy levels. In chapter 5 we shall see
how this is likely to work. The nucleus is actually more complex to deal with
quantum mechanically then atoms and we will not seek rigorous solutions but identify
the main features needed to confront the observational data.
3. Semi-Empirical Mass Formula
The aim is to produce a predictive formula which reproduces the measured values of
BE to better than 5%.
(a) Energy Equation
In general the total nuclear energy, E = ∑ Ei + ∑ m j c 2 (9) [R]
i j

EV + ES + Ee + Ea + Ep + ... + M ( A, Z )c 2 = ( A − Z )mn c 2 + Zm p c 2
↑ ↑ ↑ ↑ ↑ ↑
volume surface electrostatic assymetry pairing term rest
energy energy energy term coupling mass
∝ R3 ∝ R2 ∝ R −1 between
2 −1
∝A ∝A 3
∝A 3
close fermions

The first two terms are from the so-called liquid drop model, which appeals to some
properties of liquids to characterise the main terms in the interaction energy. To these
are then added successive approximations that add in additional effects.

(i) Volume energy, Ev


This is the main term in BE. As noted previously in equation 3 the nuclear density
appears to be independent of size, which implies incompressibility. This is
reminiscent of a liquid. Another similarity with a liquid is that the main force holding
the nucleus together is a very short-range force, which only acts between nearest
neighbours. To a first approximation there is the same amount of energy associated
with each nucleon added, i.e.
BE ∝ A = + a v A
BE (10) [R]
= av = 15.56 MeV
A
A term like this can also be justified by appealing to a fermi gas model for the
neutrons and protons contained in the nucleus. Each of these fermion types can be
treated as separate free gases in a box and their fermi energy, ε f , is given by

ε fn =
(
h 2 3π 2 nn )
2
3
for the neutrons (11a)[G]
2m n

and ε fp =
(
h 2 3π 2 n p )
2
3
for the protons (11b)[G]
2m p
with nn = N and n p = Z being the number densities of neutrons and protons
V V
respectively. To a first approximation it will be seen later than the number of neutrons
is roughly equal to the number of protons, i.e. N ≈ Z ≈ A . Then total energy
2
associated with the two nucleon fermi gases assuming they are both degenerate (i.e.
kT << ε f ) is then

U = (Nε fn + Zε fp ) = A
3 3 (
h 2 3π 2 A
2V
)
2
3

(12)[G]
5 5 2mn
From equation (1) it can be seen that V ∝ A which then implies U ∝ A as required.
Although this is a useful exercise to see how an energy level structure can give a main
total energy term which scales as volume if the effective value of av implied by
equation (12) is evaluated it turns out to be too small. This is partly because a fermi
gas model for the nucleon energy levels is not appropriate, as we shall see later, and
also because the assumption that the effective temperature is such that kT << ε f is
also not valid.
.
Figure 4 shows how the various energy terms combine to predict a form for BE/A
which eventually approximates the data shown in figure 2 to within 5%. The volume
term provides the main component and appears on this figure as a positive constant of
15.56 MeV/nucleon.

(ii) Surface tension


This term is a correction, which allows for the fact that the outer nucleons will be less
tightly bound than those inside the nucleus, whilst the previous volume term assumed
all nucleons were equally tightly bound. As the strong binding force is very short-
range only nearest neighbour interactions are important and the outer nucleons have
fewer nearest neighbours. This term is negative as the overall binding energy is
lowered and the size of the correction scales as the number of nucleons affected,
which is proportional to the surface area. Hence for this term in the binding energy
we have:-
2
BE = −a s A 3

BE −1
= −a s A 3 (13) [R]
A
a s = 17.23 MeV
In figure 4 the effect of this negative correction can be seen.

(iii) Electrostatic
So far the positive charge on the protons has been ignored. These will obviously repel
each other, which lowers the binding energy. Assuming the protons are distributed
uniformly throughout the nucleus the electrostatic energy associated with a small shell
within the nucleus interacting with those protons at smaller radii is

4πr 2 drρ c 4 πr 3 ρ c
dE e = 3
R 4πε o r
(14) [R]
r 4πr ρ
4 2
= c
dr
dr 3ε o
The total electrostatic energy is then
R 4 πr 4 πR 5 2
4
Ee = ∫ ρ c2 dr = ρc
0 3 ε 3 5ε o
o

Substituting for ρ c from equation 4


4 πR 5 Z 2e 2
Ee =
gives
( )
3 5ε o 4 2 π 2 R 6
3 (15)[D]
2 2
3 Z e
=
5 4πε o R
As charge is quantised the integration over radius should rigorously only start once a
single charge has been enclosed. If this is done the total electrostatic energy is
3 Z (Z − 1)e 2
Ee = (16)[D]
5 4πε o R
1 −1
Functionally this means E e ∝ ∝ A 3 and a negative term must be added to the
R
binding energy to account for the electrostatic repulsion which lessens the overall
binding strength. Usually this is written in as:-
Z2
BE = −a c 1
A 3
BE Z2
= −ac 4 (17)[R]
A A 3
a c = 0.7 MeV
Figure 4 illustrates how this correction affects the overall binding energy per nucleon
curve.

(iv) Asymmetry term


This term relates to the imbalance, or asymmetry, between the numbers of neutrons
and protons in a nucleus. It arises because neutrons and protons are both fermions
and each particle must have a different wave function to all other particles of the same
type. This means that as more neutrons (or protons) are added to a nucleus those
neutrons (or protons) must occupy successively higher energy levels. This contributes
to the total energy of the nucleus and it turns out that with two entirely separate
species (neutrons and protons), with their own separate energy level states, that a
minimum overall energy, for a given total number of nucleons, occurs when there are
equal numbers of each species. This can be demonstrated by using the fermi gas
model again. The total energy, U, (assuming kT << ε f ) is given by

3 (3π 2 N
V
)2
3
3 (3π 2 Z
V
)
2
3

U = U p + U n = Nh 2 + Zh 2 (18)[D]
5 2m n 5 2m p

=
( )
3 h 2 3π 2
2
3   N  23 Z  3
N  + Z   
2

5 2m n   V   V  

=
( )
3 h 2 3π 2 3
2
 N 53 + Z 53 
5 2m V 2 3  
n

However at this point substitute equation (1) and the fact that A=N+Z to get
2
3 h2  9π  3  1   5 3
 3   2  N + (A − N ) 3 
5
U=   (19)[D]
 
10 mn  4ro   A 3  
If N is allowed to vary this is a minimum when
2
3 h 2  9π   1
 5 2 3 5
 N − (A − N ) 3  = 0
3
dU
   2
2
=   (20)[D]
dN 10 mn  4ro3   3

 3
A 3
A
which requires N = ( A − N ) i.e. N = . The total energy corresponding to the
2
minimum configuration of N=Z=A/2 is already contained within the main volume
energy term in the binding energy. If N ≠ Z then the internal energy will be higher
than the minimum and this will weaken the binding by an amount
∆U = U p + U n − 2U A (21a)[R]
2

3 3 A
= Zε Zf + Nε Nf − 2 ε fA 2 (21b)[G/D]
5 5 2
From equation (20) it can be seen that
 1  3  5 3  A  3 
2 5

∆U ∝   Z + (A − Z ) − 2  
5
3
 A    2  
(22)[D]
1  13 5 3 3
5
  
∝  A Z + A 3 (A − Z ) 3 − 2 A   
1 5 2 1
A
  2  
A A
Now substitute Z = + δ and N = − δ to get
2 2
1  13  A 3
5 5 5
 3 1  A  3 2 1  
∆U ∝  A  + δ  + A  − δ  − 2 A   3
A
 2  2   2  


A
{
1 13
A (A + 2δ ) 3 + A 3 (A − 2δ ) 3 − 2 A 2
5 1 5
} (23)[D]

1  2  2δ  3 
5 5
2 2δ  3 2
∝  A 1 +  + A 1 −  − 2A 
A  A  A 

δ
Assuming << 1 and expanding to second order in δ gives
A
1   5 2δ 5 2 4δ 2  2 5 2δ 5 2 4δ 2  
∆U ∝  A 2 1 + +  + A 
 3 A + 3 3 2! A 2
1 −  − 2 A 2 
A   3 A 3 3 2! A 2     (24)[D]
δ 2

A
From the basic definition of δ this becomes
δ 2 (N − Z )
2
∆U ∝ ∝ (25)[D]
A A
This now shows the basic form for a correction to the binding energy resulting from
an imbalance or asymmetry between the numbers of neutrons and protons in a
nucleus. Writing this in the usual way gives
BE = −a a
(N − Z )2
A
BE
= −aa
(N − Z )2 (26)[R]
A A2
a c = 23.3 MeV
Figure 4 illustrates the size of this correction as a function of A for the stable nuclei.
This implies the number of neutrons is, in general, not quite equal to the number of
protons for the stable nuclei and the reason is that a nucleus will attempt to find a state
of lowest overall energy and so there is an interplay between the previous Coulomb
term and this asymmetry term. This is dealt with in detail in the next chapter.
(v) Pairing term
The main binding energy term accounts for the short-range strong force interaction
between pairs of nucleons. It assumes all nearest neighbour pairs are equally strongly
bound, such that the overall binding energy scales as the number of nucleons.
However the actual nearest neighbour separation between pairs of nucleons may vary.
Classically the closest a pair can get is to touch each other, which means their centre-
to-centre separation is nucleon diameter. Quantum-mechanically the situation is very
different and nucleon wavefunctions can actually overlap. The spatial overlap can
even be complete, as long as the spin part of the wave function is different –
remember both neutrons and protons are fermions, with S=½, and obey the Pauli
Exclusion principle. This means there can be pairs of nucleons with complete spatial
overlap but with spins opposing. These pairs of nucleons will be very tightly bound
by the strong force. As neutrons and protons have their own separate, and different,
energy levels, this ‘pairing’ effect is most pronounced for pairs of nucleons of the
same type. Pairing tends to be complete for even numbers of neutrons/protons.
Nuclei for which Z and N
are both even will show the strongest binding as all nucleons tend to be in bound
pairs. Those with Z and N both odd will always have at least 1 unpaired neutron and
1 unpaired proton and benefit the least from the pairing effect. Nuclei with A odd will
be intermediate to these other two cases. The main binding energy term includes a
contribution for the intermediate case and the binding energy correction is put in as
−1
BE = δ = + a p A 2
Z & N both even
BE = δ = 0 A odd
−1
BE = δ = −a p A 2
Z & N both odd
BE −3
= +a p A 2 Z & N both even (27)[R/G]
A
BE
=0 A odd
A
BE −3
= −a p A 2 Z & N both odd
A
a p = 12 MeV
Note that the actual form used for this term does vary from one text to another.

One effect of this correction is that, for even values of A, even-even nuclei are more
tightly bound and have the lowest overall energy. This is reflected in the fact that
there are 177 stable even-even nuclei but only 6 stable odd-odd nuclei. There are 121
stable nuclei for which A is odd.
Further evidence for nucleons coupling in opposing spin pairs comes from
measurements of the magnetic moments of nuclei. The overall magnetic moment will
be the net results of adding together the magnetic moments of all the individual
nucleons. There is zero contribution from pairs with equal and opposite spins. Hence
if most nucleons are paired the overall nuclear magnetic moment should be close to
that expected from just a few unpaired nucleons. The neutron and proton have
different magnetic moments with
µ n = −1.913µ N µ = Nuclear magneton
where N (28)[G]
µ p = 2.793µ N = 5.051 × 10 − 27 J T -1
Most measured magnetic moments are very small or zero. Even for the heavier nuclei
the maximum nuclear magnetic moments to be found are below 6 µ N which implies
most nucleons are paired together.

(b) SEMF summary


This chapter has dealt with a number of terms which are required for the nuclear
binding energy to produce a reasonable agreement with measured data. The binding
energy is a measure of how tightly the nucleus is bound. Nuclei will tend to find a
state corresponding to the tightest binding possible, and this can also be thought of as
finding a lowest energy configuration, which in turn implies a lowest mass as the
overall energy can ultimately be expressed as the rest mass energy.
Recall that
M (A, Z ) = Nmn + Zm p − BE (A, Z ) / c 2
Collecting all the terms together for the binding energy produces
BE (A, Z ) = a v A − a s A − a a (N − Z ) A −1 + δ
2 −1
− ac Z ( Z − 1) A
3 3 2
(29a)[R/G]
BE (A, Z )
= av − a s A 3 − a c Z ( Z − 1) A 3 − a a (N − Z ) A − 2 + δA −1
−1 −4 2
(29b)[R/G/D]
A
−1 −1
with δ = a p A 2
,0,−a p A 2
for even - even, A - odd, odd - odd respectively .
Figure 2 shows that this expression agrees with the measured values to within 5% for
all the stable nuclei with A>20. There are some values of A for which systematic
variations do occur and these are the subject of the next section.

(c) Magic numbers


The pairing effect discussed in section 3(a)(v) leads to some nuclei being more stable
than others. If the outer nucleon energy levels are s-shells and are paired this gives
the strongest coupling to the central potential well. This also happens in atoms where
the ‘magic’ atoms are those with closed s-shells which enhances the electrostatic
coupling due to penetration of the electron wavefunctions into the nucleus. These
magic atoms are the hardest to ionise – they are the inert gases – see figure 5.
For nuclei the effect is more complex as there are two separate species (neutrons and
protons) and each of these can have a ‘magic’ configuration. There can be ‘double-
magic’ nuclei. The most significant deviations of the measured binding energy from
the predictions of the SEMF seen in figure 2 occur near these values. This will be
covered more fully later.

4. Radio-Active Decay
A nucleus might be unstable to decay if the net energy of the final products is less
than that of the starting nucleus. The energy difference, Ei – Ef, is defined as the
energy deficit, or Q value. Decay is possible if Q>0 and the decay (or transition) is
not disallowed by any other selection rule (or conservation law). The energy released
in the decay will come out in the kinetic energy of the fragments or in nuclear
excitation followed by relaxation and γ-ray emission.
As an example the case of beta-decay, which changes a neutron into a proton inside
the nucleus will be studied.

(a) Beta-decay – Q value


In the beta decay process a neutron inside the nucleus is converted into a proton. An
electron is ejected from the nucleus together with a neutrino. The Q value is defined
as

Q = M (A, Z )c 2 − M (A, Z + 1)c 2 − me c 2 (30)[R]

Recasting this in terms of binding energies gives

Q = Nm n c 2 + Zm p c 2 − BE (A, Z ) − (N − 1)mn c 2 − (Z + 1)m p c 2 + BE (A, Z + 1) − me c 2


= mn c 2 − m p c 2 − me c 2 − BE (A, Z ) + BE (A, Z + 1) (31)[D]

Using the SEMF the binding energies are

BE (A, Z ) = a v A − a s A − a a (N − Z ) A −1 + δ A, Z
2 −1
− a c Z ( Z − 1) A
3 3 2
(32)[D]
BE (A, Z + 1) = av A − a s A 3 − ac ( Z + 1) ZA − a a (N − 1 − (Z + 1)) A −1 + δ A, Z +1
2 −1 2
3

The difference between these two is


B E (A, Z + 1) − B E ( A, Z ) = −a c Z (Z + 1 − (Z − 1))A
−1
3
( 2 2
)
− a a (N − Z − 2 ) − (N − Z ) A −1 + δ A, Z +1 − δ A, Z

= −2a c ZA
−1
3
− aa (((N − Z ) 2
) 2
)
+ 4 − 4(N − Z ) − (N − Z ) A −1 + δ A, Z +1 − δ A, Z
− 4a a A −1 + 4(N − Z )a a A −1 + δ A, Z +1 − δ A, Z
−1
= −2a c ZA 3 (33)[D]
The pairing energy difference will depend on the starting nucleus. Consider the case
of an odd A nucleus, say odd-even, then it remains odd A after decay but becomes
even-odd. In this case δ A, Z +1 = δ A, Z = 0 and

Q = mn c 2 − m p c 2 − me c 2 − −2ac ZA 3 − 4a a A −1 + 4(N − Z )a a A −1
−1
(34)[D]
Substituting values in units of MeV yields
+ 93.2(N − Z − 1)A −1
−1
Q = 939.6 − 938.3 − 0.511 − 1.4ZA 3

+ 93.2(A − 2Z − 1)A −1
−1
= 0.8 − 1.4ZA 3
(35)[D]
= 94 − 1.4ZA 3 − 93.2(2Z + 1)A −1
−1

As long as Q>0 beta decay can occur. As successive decays occur the value of Z
increases until stability is reached. Stability happens when
− 93.2(2 Z + 1)A −1
−1
0 = 94 − 1.4ZA 3

= 94 − 93.2 A −1 − Z 1.4 A 3 + 186.4 A −1 


−1

 

and hence Z =
(
94 − 93.2 A −1
) (36)[D]
1.4 A − 13 + 186.4 A −1 
 
Evaluating for A=101 and A=202 predicts stable nuclei with Z=43 and Z= 80
respectively. This shows that the values of Z expected for stable nuclei are somewhat
less that A/2. This can be seen clearly in figure 1. Figure 6 illustrates this more
clearly with data for nuclei around A~101. Note this figure does not keep A constant.
Those with a ‘β-’ in the last column are unstable to beta decay. The nucleus with
A=101 and Z=44 is stable. Nuclei with a ‘ε’ in the final column are also unstable, but
this time it is to positron decay which decreases Z and so the stability criterion can be
approached from either side.

Radioactive decay is a means of achieving the lowest energy state for a nucleus.

The higher the value of Q the more energetically favourable the decay is and in
general the more quickly it happens. The penultimate column in figure 6 shows the
decay time constants and illustrated this effect clearly.

Figure 6 Nuclear wall chart for Ruthenium (Z=44)

(b) Types of decay


There are four types of common decay. The first three are weak decays.
(i) Beta-decay
At the quark level this is ddu → duu + e − + ν e
For a specific neutron this is n → p + e− +ν e (37a)[R]

For a nucleus this is A
Z X→ Y + e +ν e
A
Z +1 (37b)[R]
In general Q = (M (A, Z ) − M (A, Z + 1) − me )c 2
(37c)[R]
Beta-decay happens for free neutrons.
(ii) Positron-decay
At the quark level this is duu → ddu + e + + ν e
For a specific proton this is p → n + e+ +ν e (38a)[R]
+
For a nucleus this is A
Z X→ Y + e +ν e
A
Z −1 (38b)[R]
In general Q = (M (A, Z ) − M (A, Z − 1) − me )c 2
(38c)[R]
Beta-decay does not happen for free protons.

(iii) Electron capture


In this mode an atomic electron is captured by the nucleus. Quantum mechanically
this is enhanced by the penetration of some atomic electron wavefunctions into the
nucleus.
For a nucleus this is e − + ZAX → Z −A1Y + ν e (39a)[R]
In general Q = (M (A, Z ) + me − M (A, Z − 1))c 2 (39b)[R]

(iv) Nuclear fission


In this mode the nucleus splits into two nuclear fragments.
A− B
Z X → Z −C Y + C W
A B
For a nucleus this is (40a)[R]
Specific cases are:- B=1 and C=0 - neutron drip
B=1 and C=1 - proton drip
B=4 and C=2 - alpha decay
In general Q = (M (A, Z ) − M ( A − B, Z − C ) − M (B, C ))c 2 (40b)[R]

(c) Weak decays


In this section a more general procedure will be used to predict which nuclei are
stable. The approach is based on looking for a minimum in the rest mass for a
nucleus of fixed A value. Nuclei will decay towards this minimum from both sides.
Nuclear masses will be derived using the SEMF and so-called ‘mass parabolae’ will
be used.

(i) Mass parabolae from SEMF


For a fixed value of A there is only 1 free parameter for a nucleus; either N or Z.
Using Z the mass can be written
M (A, Z ) = (A − Z )mn + Zm p − BE (A, Z ) / c 2 (41)[R]
Substituting for the binding energy and expanding as a power series in Z leaves
M (A, Z ) = C 0 + C1 Z + C 2 Z 2 (42)[R/D]
2
C 0 = Amn c 2 − av A + a s A 3
+ aa A + δ
where C1 = (m p − mn )c 2 − a c A
−1
3
− 4a a (43)[D/G]
−1
C 2 = ac A 3 + 4a a A −1
C1 is < 0 whilst C2 >0 which means there will be a minimum with respect to Z and a
plot of the mass against the value of Z will define a parabola, called the ‘mass
parabola’. The result of plotting this for A=101 is shown in figure 7. The small
points joined by the solid curve are the predicted points from the SEMF. The larger
dots are the measured values. The location of the minimum in the parabola can be
found by differentiating the equation.

Figure 7: Mass parabola for an A odd Figure 8: Mass parabolae for A even
δ=0, A=101)
nucleus (δ δ=±
nucleus (δ ±∆, A=100)

dM
= C1 + 2C 2 Z
dZ
C1
The minimum is then when Z = . For A=101 the minimum is at Zmin=44, as can
2C 2
be seen from figure 7 and the discussion in the previous section. Also shown on
figure 7 are the decay process showing beta-decays if Z < Zmin and positron decay or
electron capture if Z>Zmin. These decays are allowed energetically as long as the mass
difference is greater than the electron rest mass - this is half a division on the vertical
scale of figure 7. If A is even the situation is more complicated because the pairing
correction, δ, is no longer zero. For an even A nucleus successive decays, whether
beta decays or positron decays, changes the type of nucleus as
N odd → even → odd → even →
. The value of δ changes in sign each
Z even → odd → even → odd →
time a decay occurs. This means every other nucleus has the same sign for δ and
there will effectively be two mass parabolae defined by alternate nuclei. This can be
seen in figure 8, which is plotted for A=100. Nuclei on the sides of the parabolae
work their way done towards the valley moving from one parabola to the other. In the
case of A=100, it can be seen from figure 8 that there are actually two end points.
Nuclei decaying from low Z will end up at Z=42 as a decay to Z=43 is not
energetically allowed an this is directly due to the effect of δ changing sign. Nuclei
decaying from high Z will end up at Z=44. A nucleus at Z=43 can decay either way
and which way a specific nucleus goes is purely statistical with probabilities
determined by so-called branching ratios.
The two stable nuclei for A=100 are both even-even as this parabola is lowest in
energy/mass. For an odd-odd nucleus to be stable the mass parabolae must be very
steep. This does happen for a few nuclei. 14N is an example of a stable odd-odd
nucleus.
(d) Alpha decay
The alpha decay reaction is
A− 4
Z X → Z − 2 Y + 2 He
A 4
(44)[R]
The energetics of alpha decay will first be examined using the SEMF before
developing a deeper theoretical foundation, which will allow other properties of this
radioactive decay mode to be studied.
(i) SEMF
The Q value for the decay is given by
Q = [M (A, Z ) − M (A − 4, Z − 2) − M (4,2)]c 2
(45)[D]
= BE (A − 4, Z − 2) + BE (4,2 ) − BE (A, Z )
For alpha decay to be energetically allowed the value of Q must be positive. Q > 0 if
BE (4,2) > BE (A, Z ) − BE (A − 4, Z − 2) (46)[R]
This can obviously be worked out explicitly for any particular nucleus. In general it is
know that it is only the heavier nuclei which tend to show alpha decay. For large A
the form of the binding energy per nucleon is linear with A. This allows a
simplification to the calculation using
∂B
BE (A, Z ) − BE (A − 4, Z − 2) = ∆B ≈ ∆A (47)[D]
∂A
The rate of change of B with A is related to the form of B/A using
( )
∂B
A = 1 ∂B − B which can be rearranged to give ∂B = A
( )
∂B
A +B.
∂A A ∂A A 2
∂A ∂A A
Substituting this gives

BE (4,2) > 4 A

( )
 ∂B
A +
B 
(48)[D]
 ∂A A
 
From figure 2 the rate of change of B/A with A for large A is
( )
∂B
A ≈ −7.7 × 10 −3 MeV/A 2 (49)[G/C]
∂A
and
B 
BE (4,2) > 4  − 7.7 × 10 −3 A (50)[D]
A 
The binding energy of an alpha particle (helium nucleus) is 28.3 MeV [G]. Using this
value
B 
28.3 > 4 − 7.7 × 10 −3 A or
A  (51)[D]
B −3
≤ 7.075 + 7.7 × 10 A
A
is then the criterion for alpha decay to be energetically possible. This criterion
defines a region of parameter space on the binding energy per nucleon plot of figure
2. This shows that decay becomes energetically possible for A>151.
Other points which can be seen by doing an explicit calculation of Q using the SEMF
are that:
• For fixed A, Q increases for large Z
• For fixed Z, Q decreases with increasing A
(ii) Theory
Although the SEMF is very useful in showing what is possible it does not explain
‘how’ the reaction occurs or give any means to calculate other observable properties,
such as life-time (or half-life). These two questions will be used to encourage a
theoretical approach.

The Decay Mechanism


To understand how a composite particle can be emitted from a nucleus it is necessary
to think about the potential wells seen by nucleons ‘trying to escape’.

~6 MeV ~6 MeV

The neutrons, on the left, see a potential well created by the strong force. The
topmost occupied energy level, shown here with two neutrons in it, will typically be
about 6 MeV below zero. For the protons, on the right, there is also the Coulomb
interaction. Paradoxically, once inside the nucleus, this acts as an additional
impediment to escape as the immediate potential barrier is higher. Of course once a
proton is moved far enough away to get onto the downward slope it then is repelled
away. For any of these nucleons to escape they would need to tunnel through the
barrier. For an alpha particle to escape two neutrons and two protons would need to
simultaneously tunnel out. This is highly unlikely. However, what is thought to
happen is that there is a finite chance of two protons and two neutrons occasionally
getting close enough together within the nucleus to form a sub nucleus within the
larger nucleus. This can release a certain amount of energy equivalent to the effective
binding energy of this sub-nucleus and the newly formed alpha particle finds itself in
a shallower energy level as shown below. The alpha is in a level with energy Qα


~6 MeV
r=b

and is still classically forbidden to escape. However quantum tunnelling of this sub-
nucleus as an entity is much more likely. Once the particle wavefunction has
extended out to r = b Coulomb repulsion will help repel the particle away. The
probability of quantum mechanical tunnelling out to r = b can be estimated using a
rectangular potential barrier approximation.
U

ψ~eikr ψ~αeKr
+Be-ikr ψ~Ceikr
+βe-Kr

nucleus barrier free

In the regions where U = 0, (i.e. inside the nucleus and beyond the barrier) the
wavenumber k is given by hk = 2 ME (52a)[R] where M is the mass of the particle.
Inside the barrier hK = 2 M (U − E ) (52b)[R]. A standard exercise in quantum
mechanics is then to match the wavefunction and its derivative at all boundaries to get
−1 −1
 k2   K2 
C = 4e 1 + 2  1 + 2
− 2 Kt

2

 K   k 
 K 2  k 2 
= 4e − 2 Kt  2 
2  2

2 
(53)[D]
 K + k  k + K 
k 2K 2
= 4e − 2 Kt
(k 2
+ K2 ) 2

where t is the barrier thickness. If U ≈ 2 E ⇒ k ≈ K and C ≈ e −2 Kt (54)[R/D/G].


2

A closer approximation to the 1/r Coulomb potential barrier shape can be obtained by
using a sequence of rectangular barriers of equal width, dr, but successively smaller
height.

R r=b r

If the thin rectangular barrier sections are numbered sequentially out from the nucleus
the overall transmission through the potential barrier is simply the product of the
transmission probabilities through each of the thin slabs, i.e.
T = C = e −2 K1dr e −2 K 2 dr e −2 K 3dr ...
2

−2 ∑ K i dr − 2 K (r )dr
b

=e i
⇒ e ∫R (54)[D]
2 b 
− ∫R 2 M (V (r )− E )dr 
=e h 
≡ e −G

where G is the Gamov factor.


For a Coulomb barrier the potential is
2 Ze 2
V (r ) ≈ (55)[R]
4πε o r
The integration ends when r = b at which point the Coulomb barrier potential is equal
to the alpha particle kinetic energy:-
2Ze 2
V (b ) = E = (56)[D]
4πε ob
Evaluating the integral eventually gives (see problem sheet):-
Z
G ≈ 3.96 − 2.97 RZ (57)[G]
E
where E is the alpha particle energy in MeV
R is the nuclear radius in fm
Z is the atomic number
The value of G depends sensitively on E. This means the transmission factor which
varies as T = e − G is incredibly sensitive to the energy, E.
For example, consider the radioactive nucleus 238 92 U . It has a radius of 10fm and emits
alpha particles of energy ~4.2MeV.
G = 180 − 95 = 85 and
Hence (58)[C]
T = e −85 = 1.8 × 10 −37
If the alpha energy were a factor of ~2 higher at 9MeV, say, the numbers would be
G = 30 ⇒ T = 1.3 × 10 −13 . (59)[C]

In this example the transmission factor, T, has changed by a factor of 1024 for a
factor of 2 change in alpha particle energy! This is a startling prediction of this
theory and it will be seen how this stands up to observational scrutiny in the next
section.

Decay Half-life
If there are a certain number of radioactive nuclei to start with the rate at which they
decay is given by
N = N o e − λt
dN (60)[R]
= − λN
dt
The half-life, τ 1 , is defined as the time over which half the nuclei decay, i.e.
2
− λτ 1
e 2 = 0.5
λτ 1 = ln (0.5) (61)[R/D]
2

ln (0.5)
τ1 = −
2 λ
A theoretical estimate of λ can be obtained from the preceding theory. λ is the rate of
decay and this will depend on the probability of quasi-alpha particles existing within
the nucleus, the rate at which these quasi-alpha particles then collide with the
potential barrier walls, and the probability of a barrier tunnelling at each collision.
The probability of alpha particles existing within the nucleus can be characterised by
using a number giving the mean number of alpha particles averaged over a long time
period, Nα. Hence
λ ≈ N α fT (62)[R]
where f is the frequency of wall collision and T is the barrier transmission factor due
to tunnelling. In general Nα will be <<1. The frequency of wall collision will be
roughly the distance between wall collisions divided by the alpha particle speed. The
barrier transmission is given from the Gamov factor in the preceding theory. Putting
these together and taking logarithms gives
v
ln λ = ln N α + ln α − G
R
2 Eα 1
= ln N α + ln −G (63)[D]
mα R
2 1 12 −1
= ln N α + ln Eα − 3.96 ZEα 2 + 2.97 RZ
mα R
The largest term in this is the third one and hence the main trend should be
−1
ln λ ∝ ZE 2 (64)[R/D].
How well this agrees with observation will be seen in the next section.

(iii) Observation
Measurements of radioactive alpha particle decay rates exist for a number of heavy
elements with values of A between 210 and 240. The logarithm of these is plotted in
figure 9 as a function of Z . According to the theory this plot should show a

linear relationship. The solid symbols are the measurements and the dashed line
shows the best linear fit to the data. The fit is a good match over a range of half-
lives ranging from 10µ µs up to the age of the Universe!! The theory works over a
23
factor of 10 in half-life.

Figure 9: Logλ against Z/E1/2 for Figure 10: Log τ1/2 against Qα for the isotopes of
various nuclei with a linear trend various elements. The form of these curves is well
line superimposed. matched to the Geiger-Nuttall rule.
The understanding of alpha particle decay using quantum theory and tunnelling came
about through the work of Gamov, Condon and Gurnley. Before quantum mechanics
was available an empirical relationship between the decay rate and the alpha particle
range, Rα, in air had been established. This was known as the Geiger-Nuttall rule and
took the form:-
1
log10 = B log10 Rα − C (65)[G]
τ1
2

If Rα is in cm and τ in seconds then B = 57.5. The value of C depends on the element


and equals 41 for uranium. The range in air, Rα, is proportional to the alpha particle
energy, Eα, and the energy of the alpha is close to the Q value as all the excess energy
is released as kinetic energy of the lightest product. Figure 10 shows plots for various
elements with radioactive isotopes and the systematic variation of half-life with Q
value is clearly seen. The curves connecting isotopes of the same element are of the
form given by the Geiger-Nuttall rule.

Figure 11: Qα against A over a more extended range than figure 10.

Figure 11 shows a more extensive plot covering a wider range of A. The general
trend of decreasing Q as A increases for fixed Z is clearly seen, although there is also
some change in this behaviour around A=210. The reason for this will become
apparent in the next chapter.

(iv) Decay series


The decay of the heavy transuranic elements tends to produce chains of successive
decays as the nucleus works its way to the top of the stable nucleus locus ( see figure
1). For example 238U decays to 234Th by alpha decay. 234Th is unstable to beta decay
and then decays to 234Pa. Likewise 234Pa is unstable to beta decay and decays to 234U.
The decays, of various types, continue until a stable nucleus is reached, which in this
case is 206Pb. The overall sequence is called a 'decay series' or 'chain'. Figure 12
shows the whole series for 238U.

Figure 12: Decay series for 238U

The main features of the 238U decay series are:-


• The decay half-life of 238U is extremely long. It is 6.5×109 years, which is a
significant fraction of the age of the Universe. Uranium is produced in supernova
explosions (see later) and its long lifetime ensures that it is still around long after
the supernova ejecta has dispersed and mixed into its surroundings. Uranium is
found naturally on Earth and this will have come from ancient supernovae.
• The decay series involves 15 steps with successive alpha and beta decays ending
up with 206
82 Pb (see figure 12)

• 238
U has the longest life-time of all the nuclei in the series (apart from 206Pb,
which does not decay of course). All the ‘daughter’ nuclei have short enough life-
times that once a parent nucleus has decayed the whole series proceeds quickly
enough that there is no bottle-neck in the chain. This means that for a sample
containing a large number of uranium nuclei, which has reached equilibrium:-
! the ‘activity’, or number of decays per second, of each daughter type is the
same.
! the amount of each daughter nucleus is proportional to its lifetime.
• Some decays leave an excited nucleus (see next chapter) and this leads to γ-ray
emission as the nucleus relaxes to its ground state. Figure 13 shows a detailed
example of this for the decay of 226Ra. About 5.4% of the alpha decays leave the
resulting 222Rn nucleus in an excited state. This decays in about 0.3ns emitting a
γ-ray of energy 186keV.
• The end of the chain is a particular stable nucleus.
• For uranium there are four possible decay series depending which isotope the
92 U → 82 Pb 92 U → 83 Bi
238 206 237 209

chain starts from. These are 239 .


92 U → 207
82 Pb 236
92 U → 208
82 Pb
It is intriguing that all the decay series for U end up on a stable nucleus with A~208
and Z~82.

Q. Why is this?
A. The stable nuclei must be so tightly bound that it is not energetically favourable for
another decay to occur.

Q. Why are these nuclei so tightly bound?


A. Because they are ‘magic’ or even ‘double-magic’

Q. Why are they magic nuclei?


A. Because the nucleus is a quantum system, which has separate energy level
structures for the neutrons and protons. Magic behaviour occurs when there are filled
shells. This is the topic of the next chapter.

Figure 13: Detail for the alpha decay step involving 226Ra showing possible branches
for alpha decay leaving excited nuclear states with subsequent gamma-ray emission.

5. Shell Model
The Shell Model is an attempt to understand the existence of features in nuclear
properties which show systematic divergence from the SEMF predictions at particular
values of N and/or Z.
The assumption is that there must be quantum mechanical energy levels within the
nucleus for both neutrons and protons and that the ‘magic’ numbers occur where the
binding energy is enhanced through detail in the energy level structure and
occupancy. If this is the correct reason then it should be possible to recover the magic
numbers from a predicted energy level diagram.
(a) Magic numbers
Magic numbers are also displayed by quantum system with energy level structures. In
general there are special characteristics associated with systems in which complete
energy level shells are occupied. A brief reminder about atomic systems will be given
first.
(i) Atoms
For atoms we know that closed shell effects can give tighter binding of atomic
electrons, leading to increased ionisation potentials at Z = 2, 10, 18, 36, 54, 80, 86
….(see figure 5). These are noble gases with filled outer p-shells causing the outer
electrons to ‘see’ more clearly the Coulomb potential.
(ii) Nuclei
The are a number of effects which are manifestations of the magic number behaviour
of nuclei. These are:-
• BE varies smoothly from the SEMF prediction near the magic numbers
• It is more difficult to remove single neutrons from nuclei with magic numbers of
neutrons.
• It is more difficult to remove single protons from nuclei with magic numbers of
protons.
• Elements with magic numbers of protons tend to have more isotopes
• Nuclei with N magic tend to have more isotomes (range of Z values)
• The abundances of the elements are enhanced for magic number nuclei - see
figure 14a
• Magic nuclei interact less readily with other particles - e.g. it is more difficult to
induce neutron capture as more energy is needed to change to nuclear state away
from a preferred magic on - see figure 14b

The magic numbers at which these magic nuclei occur are

N and/or Z = 2, 8, 20, 28, 50, 82, 126, … (66)[G]

208
Note that 82 Pb is actually double magic !!!

In the next section various energy level schemes will be studied to try to understand
the magic number sequence.

(b) Energy levels and potentials


Previously the neutrons and protons were considered as separate gases of fermions
contained in a box - the nucleus. In this section the energy level structure will be
looked at more closely together with the effect of using different forms for the
retaining potential.

For a general quantum mechanical approach a solution to the relevant Shrodinger


equation is needed.
h2 2
− ∇ Ψ + V (r )Ψ = EΨ (67)[R]
2m
where it is assumed the wavefunction can be separated into
Ψ = R(r )Υlm (θ , φ ) (68)[R]
l is the orbital angular momentum, m is the magnetic quantum number ( − l ≤ m ≤ l )
Also L2 = l (l + 1)h 2 and LZ = mh (69)[R]

Figure 14a: Abundances of elements measured Figure 14b: Cross-section for neutron
in solar system. Peaks are seen around magic capture by nuclei shows lower ease of
number nuclei capture by magic nuclei.

The radial equation can be written using R(r ) = U (r )


1
(70)[R] as
r
h ∂ U (r ) 
2 2
l (l + 1)h 
2
− + V (r ) + U (r ) = EU (r ) (71)[D]
2m ∂r 2
 2mr 2 
The solutions U nl (r ) are eigenfunctions which depend on n the principal quantum
number, and l the orbital angular momentum.

A similar notation will be used for nuclear levels as that used for atomic levels, i.e.
ns np nd nf
↑ ↑ ↑ ↑
l = 0 l =1 l = 2 l = 3
Each level can accommodate 2(2l + 1) neutrons/protons (72)[R].

The big question now is how to represent V (r ).


A number of options have been tried including:-
V
• A Coulomb potential with V (r ) = − o (73a)[R]
r
⇒ atomic type energy levels with l < n
• A Simple Harmonic Oscillator with V (r ) = µω 2 r
1
(74b)[R]
2
• An Infinite Square Well with V (r ) = −∞ for r < R
• A Finite Square Well with V (r ) = −Vo for r < R
Vo
• A rounded square well (Saxon-Woods) with V (r ) = − (74c)[G]
r −R
1 + exp 
 a 
None of these reproduce the magic numbers. Figure 15 shows the energy level
structures produced by each of the potential forms. On the far right of the diagram
can be seen the numbers of fermions required to fill each of the levels and the largest
effects characteristic of magic numbers would be seen where there are the biggest
gaps above the levels. If these are worked out for the first four level schemes (putting
2(2l + 1) fermions into each level) in figure 15 it will be seen that the predicted magic
number sequence is 2, 8, 20, 34, 40, 58, 92, 112, 138, 168 … Numbers up to 20 are
correct, but beyond that there is no correspondence. A more radical change to the
form of the potential is needed. Looking beyond the form of the main attractive
potential, the next level modification to the Hamiltonian, prompted by experience
r r
with atomic systems, would be to add a spin-orbit coupling term involving L ⋅ S .

Figure 15: Energy level diagrams for various assumed central potentials. The magic
numbers are only successfully predicted when a spin-orbit term is included.
(c) Spin-orbit coupling
For atomic systems a significant addition to the Hamiltonian beyond the simple
electrostatic potential is a spin-orbit coupling term. As the nuclear magic numbers do
not seem to be predicted correctly with a simple nuclear potential the next level
modification to the nuclear Hamiltonian worth trying is a nucleon spin-orbit coupling
term. This is often added in to the potential using:-
{ } r r
V (r ) = V (r ) 1 + Wls L ⋅ S (75)[R]
The mechanics of this is dealt with in the companion ‘Particle Physics’ half of the
course currently given by Dr. Jordan Nash.
r r r
Spin-orbit coupling creates a new angular momentum vector J = L + S (76)[R].
According to the rules of quantum mechanics the coupling can produce a new vector
with a range of possible values of j, from l+s to l-s in integer steps. For neutrons and
protons s=1/2, and so there are only two possible new angular momentum options of
l+½ and l-½. This means that each of the energy levels in the first 4 potential well
options in figure 15 splits into two levels. This can be seen in the last ‘column’ in
figure 15. A level of degeneracy in l-states has been lifted. The new levels are
labelled nl j and each state has an possible occupancy of (2 j + 1) corresponding to the
allowed values of m j . From figure 15 it can be seen that the l+½ level is always
r r
lower than the l-½ level; i.e. states with L ⋅ S > 0 are more tightly bound in deeper
energy levels.

The predicted magic numbers after adding in the spin-orbit interaction are shown on
the far right of figure 15. There is a good match between numbers just below the
largest gaps between the levels and the observed magic number sequence. Armed
with this new energy level structure it is now possible to predict other nuclear
properties, such as nuclear spin, parity, and magnetic moment.

(d) Nuclear spin


For a given number of neutrons/protons the level structure can be used to predict the
values of n, l, and mj for each particle. The overall properties of the are then obtained
by summing over all the particles. In the case of nuclear spin, this is simplified by
noting that in general the particles tend to form pairs with opposing spins, and so the
overall spin is caused by nucleons which are not paired; this usually means only the
outermost nucleons are relevant.

even, even nuclei


all have spin zero

Neutrons Protons
A - odd nuclei
spin = j of add
nucleon, which is
always half-integer

n or p p or n

odd, odd nuclei


↑↑ or ↑↓ or ↓↓ or ↓↑
Jtotal = Jn + Jp
e.g. 14N has Z=N=7
Jn = ½h & Jp = ½h
⇒Jtotal = 0(✖) or 1(✔)
n or p p or n

(e) Parity
[This topic is covered in more depth in the ‘Particle Physics’ half of the course by Dr
Jordan Nash].
The overall wavefunction for the nucleus, ψ, involves all the nucleons, and the parity
of the nucleus is then the eignevalue, λ, of the parity operator, P, which inverts the
cartesian coordinate system such that x → − x, y → − y, and z → − z .
Pψ = λψ such that ψ (r ) = λψ (− r ) and λ = ±1 (77)[R]
For a single nucleon with orbital angular momentum, l
ψ (− r ) = (− 1) ψ (r )
l
(78)[R]
For a nucleus with A nucleons the overall parity of the nucleus, P, is the product of
the parities of all the individual nucleons.
A
P = ∏ (− 1) i
l
(79)[R]
i =1

(f) Magnetic Dipole Moments


The only contribution to the overall magnetic moment of a nucleus is from unpaired
nucleons. This means even-even nuclei all have zero nuclear magnetic moment.

For the unpaired nucleons there are two contributions to the individual magnetic
moment, µ, of that nucleon. These are from its spin, s, and its orbital angular
momentum, l. These combine to give:-
µ = g s s + gl l (80)[R]
 eh 
Nuclear magnetic moments are quoted in units of the nuclear magneton ≡  .
 2m 
 p 
Neutrons and protons have different intrinsic magnetic moments and in units of the
nuclear magneton:-
protons have g s = 5.5856 and gl = 1
(81)[G]
neutrons have g s = −3.8261 and g l = 0 (no charge!)
For unpaired nucleons their value of j must be used to work out their magnetic
moment using
µ = gj j (82)[R]
 j ( j + 1) − l (l + 1) + s (s + 1)   j ( j + 1) + l (l + 1) − s (s + 1) 
with g j = g s   + gl   (83)[G]
 2 j ( j + 1)   2 j ( j + 1) 

6. Formation of Nuclei
Virtually all the nuclei in existence today have been created in an astrophysical
context. There three main scenarios:-
• The lightest nuclei were mainly created during the very early Universe
• Nuclei up to the most stable ones, Ni & Fe, were made by fusion processes inside
stars during their normal lifetimes, during which the energy released is responsible
for producing their starlight. Some of these nuclei are then released into the
environment when stars explode or eject their envelopes.
• The heaviest nuclei are made in catastrophic events creating high density neutron-
rich conditions.

(a) Big-Bang nucleosynthesis


Big-Bang cosmology assumes the Universe started out as a very high density and high
temperature event. The Universe then expanded adiabatically, cooling as it did so
according to
−1
TUNIV ≈ 1.5 × 1010 t 2 K (84)[R]
During the very early Universe the density is so high that radiation and matter are in
thermal equilibrium and the effective thermal energy of the material in the Universe is
−1
EUNIV ≈ kTUNIV ≈ 1.29t 2
MeV (85)[R]

Whilst the density is high enough for a thermal equilibrium between the radiation
field and matter to exist there will be on-going particle-antiparticle formation going
on as long as the temperature/energy is high enough to allow pairs to be produced
from the radiation field. This is true when:-
For nucleon pair - production EUNIV ≥ 1.86 GeV ⇒ T ≥ 2.1 × 1013 K ⇒ t < 0.5 µs
For electron pair - production EUNIV ≥ 1.86 GeV ⇒ T ≥ 1.2 × 1010 K ⇒ t < 1.67 s
Once nucleon production has come to an end at ~0.5 µs after the start of the Big-Bang
the Universe then contains mainly neutrons and protons in the ratio
Nn  (mn − m p )c 2 
≈ exp−   (86)[D/G]
Np  kTUNIV 
 
This is very close to 1, i.e. N n ≈ N p
After this there is very little further nucleon production and a series of predictable
interactions will now take place.
To start with there will be a number of immediate reversible processes, which directly
affect the neutron to proton ratio. These are:-
p + e− ↔ n + υe
p + υe ↔ n + e+
n ↔ p + e− + υe
The first more complex nucleus (deuterium) to form come from p + n → d + γ , which
releases 2.225 MeV of energy for each deuterium nucleus formed. However in the
very early stages this process is negated by a dissociation induced by γ-ray absorption
giving d + γ → p + n . Until the ambient γ-ray flux or energy becomes low enough,
as the Universe expands and becomes cooler and more dilute, the deuterium
abundance will be kept down. This continues to be the case until about 250s when
TUNIV ≈ 9.5 × 10 8 K and the black-body γ-rays no longer have enough energy to
destroy the deuterium. However what happens now is that, as soon as the deuterium
is allowed to survive long enough, a whole chain of processes takes place leading the
4
He production. These processes actually suppress the deuterium build up even more
effectively but open the way for heavier nuclei to be produced. Figure 16 shows
graphically how the mass fraction of various nuclei are though to evolve with time.

Figure 16: Formation of light nuclei during the first three minutes after the Big-Bang.

The reactions which have been included in the calculations leading to figure 16 are:-
p + n → d +γ
d + n→ 3 H + γ
d + p → 3 He + γ
d + d →3H + p
d + d → 3 He + n
3
H + p→ 4 He
3
He + n→ 4 He + γ
There is very little production of nuclei beyond A=4 partly because there are no stable
nuclei with A=5 or A=8 to act as stepping stones. However there are still a few
heavier nuclei produced via:-
3
He+ 4 He→ 7 Be + γ
3
H + 4He→ 7 Li + γ
7
Be + n→ 7 Li + p
The main result of Big-Bang nucleosynthesis, which lasts for about the first 3 minutes
of the Universe, is that the surviving matter content of the Universe is
4
He ~ 23%
H ~ 77%
D ~ 10 − 4
3
He ~ 2 × 10 −5
7
Li ~ 2 × 10 −10

(b) Stellar nucleosynthesis


The first stars to form did so from a gas containing 77% hydrogen and 23% helium.
(i) Main Seqeunce
During the main part of their lifetime stars derive their energy by converting hydrogen
into helium. Our Sun is in this phase at the moment. This requires temperatures
around 107 K and the starting interaction is
p + p → d + e+ + υe (87)[R]
This reaction is enabled by quantum mechanical tunnelling through the Coulomb
barrier. The probability of this happening is given by a Gamov factor in a similar way
to that seen earlier, with
T = e −G (88)[R]
2m r
and G = παZ A Z B c (89)[G]
E
In this case ZA=ZB=1 for collisions between two protons, and mr is the reduced mass
for the collision. Typical timescales for the p-p starting interaction is 6×109 years,
which ensures the Sun burns ‘gently’, but the large numbers of protons available
enable enough interactions to provide the energy output of the Sun. Once the first
interaction has taken place a number of further interactions take place leading to
helium production:-
d + p → 3 He + γ
3
He+ 3He→ 4 He + p + p
or 3He+ 4 He→ 7 Be + γ
7
Be + e − → 7 Li + υ e
7
Li + p → 4 He+ 4He
The overall result is 4 p → 4 He + 2e + + 2υ e + 24.7 MeV (90)[R]

For heavier stars with core temperatures ~5×107 K there are other interactions which
can make use of existing heavier nuclei as catalysts. These heavier nuclei exist in
second generation stars in which ashes from previous stars have been incorporated.
Carbon, nitrogen and oxygen are the most effective and give rise to
12
C + p →13 N + γ
13
N →13 C + e + + υ e
13
C + p →14 N + γ
14
N + p →15 O + γ
15
O→15 N + e + + υ e
15
N + p→12 C + 4He
The overall result is again 4 p → 4 He + 2e + + 2υ e and this has used 12C as a catalyst.
(ii) Red Giants
Eventually main sequence stars will use up all the available hydrogen in the core and
when this happen they undergo gravitational collapse and the core becomes hotter. At
this stage the main material in the core is helium. If the temperature gets high enough
(~108 K) then quantum mechanical tunnelling can happen in helium-helium collisions
leading to:-

4
He+ 4He→ 8 Be Q = −0.091 MeV
4
He+ 8Be→12 C Q = 7.37 MeV
4
He+ 12C →16 O + γ Q = 7.16 MeV
4
He+ O→ Ne + γ
16 20
Q = 4.73 MeV
4
He+ Ne→ Mg + γ
20 24
Q = 9.31 MeV

When the helium in the core is exhausted and energy production stops, the core will
collapse again and if the temperature subsequently rises to ~109 K fusion can start in
the next generation of nuclei, such as:-

12
C + 12 C → 20 Ne+ 4 He
12
C + 12 C → 23 Na + p
16
O + 16 O→ 24 Mg + 4 He+ 4 He
16
O + 16O→ 32 S + γ

For the most massive stars this succession of fusion, core collapse, more fusion, more
collapse can continue until the most stable nuclei have been produced. The most
56 56
stable nuclei are those with A=56, i.e. 26 Fe, 28 Ni, 2756 Co . Beyond this there is no
further energy release and collapse would continue with no further fusion.

(c) Supernova explosions


After the red-giant phase is completed and all fusion processes have run their course a
star will collapse until it becomes one of three types of dead star:-
• A white dwarf - this is the fate of stars with masses < 1.5M". The force of
gravity is balanced by repulsion between electrons arising out of the Pauli
exclusion principle - a degeneracy pressure.
• A neutron star - these are slightly more massive than white dwarfs and electron
degeneracy pressure is not sufficient to combat gravity. The star collapses further,
electrons are ‘squashed’ out of existence through recombination with protons to
form neutrons, and then neutron degeneracy pressure stops the collapse.
• A black hole - these have masses so high that neutron degeneracy pressure can not
stop gravitational collapse, which continues unopposed until the object is so small
and compact that even photons can not escape from the surface.

The process of collapse results in very high densities being reached and neutron rich
environments forming. Sometimes the collapse is catastrophic and results in a
subsequent explosion probably caused by radiation pressure from intense neutrino
production. From the point of view of production of nuclei beyond A=56 the critical
point is that neutron rich environments can develop in dense regions. If this happens
there are two possible routes to heavy nuclei being formed. Both of these involve
neutron capture by existing nuclei. If an explosion does happen this is a mechanism
for then dispersing the elements into the outside environment. The Earth and
everything on it, including us, are formed from the remnants of stellar processing and
explosion.

(i) s-process
In moderate density situations the time between successive neutron captures for a
specific nucleus is long enough that if an unstable nucleus is formed at any point it
has time to undergo beta-decay before the next capture. This effectively allows the
nucleus to siddle along the line of stable nuclei, gradually increasing both A and Z.
This is called the ‘slow’ or s-process. Figure 17 shows which nuclei are likely to be
formed during the s-process and this can be compared to the line of stable nuclei in
figure 1.

Figure 17: Build up of heavier nuclei during supernova explosions. Two processes are
shown – the s-process (slow enough that beta decay happens between successive neutron
captures) and the r-process (too rapid for beta decay to keep up with neutron capture).
(ii) r-process
Much more rapid neutron capture is likely to occur in very high density environments,
such as during the collapse leading to explosion and formation of a neutron star or
black hole. In this case it may be possible to capture several neutrons before beta-
decay has a chance to happen and restore stability.
56
Fe + 5n→ 61Fe
e.g.
61
Fe→ 61Co + e − + υ 6 minutes
Typical combinations of nuclei formed by the ‘rapid’ or r-process are shown in figure
17. At positions corresponding to magic numbers it becomes more difficult to capture
the next neutron and this causes a pause during which decay can occur. This explains
the discontinuities at the magic numbers.

7. Nuclear Reactors
A useful power reactor needs to be able to produce large amounts of power, over long
periods of time, in a controllable fashion. Energy can be released in nuclear
interactions which result in nuclear masses moving closer to the peak, at A=56, in the
binding energy per nucleon curve of figure 2. This can be done in two; either by
combining light nuclei into larger ones or by splitting very large nuclei into smaller
ones. These correspond to fusion and fission respectively.

(a) Fission reactors


Fission causes large nuclei to split into two almost equal sized fragments. From
figure 2 it can be seen that about 1 MeV/nucleon can be released by doing this, so for
uranium, for example, some 200 MeV per fission is potentially available. The fission
process can be thought of as a progressive deformation of the nucleus into two halves.
As the halves begin to separate they first see an attractive potential pulling them back
together. However if they can separate far enough they get over this attractive barrier
and then ‘fall’ apart. The height of the barrier opposing the break up into two roughly
equal sized fragments is called the fission barrier energy, Ef, and can be approximated
(
by E f ≈ 0.455 49 − Z
2
)A
MeV . This is about 6.1 MeV for 238U. The height of this
barrier determines how likely it is for spontaneous fission to occur. To get useful
amounts of energy release requires sufficient amounts of the material plus a high
enough fission probability. The straight forward spontaneous fission rates are not
high enough. However in some nuclei it is possible to enhance the fission rate by
using neutron irradiation. Moreover there are some of these nuclei which also
produce free neutrons during the fission process. This enables a chain-reaction where
a neutron causes a fission which releases more neutrons which cause more fissions …
Nuclei which exhibit this process are called ‘fissile’ nuclei. The neutrons which cause
fission have typically thermal energies (~1/40 eV). The neutrons emitted during
fission have much higher energies, up to 1 MeV. An example is 235U. The following
bullets show the considerations needed for a sustainable reactor process using this
isotope of uranium
• 235U produces 2-3 neutrons with energies of 0.1 to 1 MeV during fission.
• At least 1 of these neutrons must induce another 235U nucleus to fission. However
various other things can happen to the neutrons preventing them from causing
fission in another 235U nucleus. These other processes include capture by 238U and
235
U without fission. The relative likelihood of these other processes and the
useful fission one are shown in figure 18. Only for low energy neutrons
92 U (n, f ) .
(<200meV) does the absorption by 235U to produce fission dominate ( 235

Figure 18: Neutron capture probabilities in uranium isotopes together with subsequent
decay products (f – fission, γ - gamma-ray emission) as a function of incident neutron
energy

This is compounded by the fact that the natural abundance of U is 99.3% 238U and
only 0.7% 235U. The only way an efficient process can be achieved is by deliberately
reducing the energy of the neutrons, En, released by fission of 235U quickly enough.
This is done using a ‘moderator’ in which elastic neutron scattering to transfer kinetic
energy to recoiling nuclei, and hence reduce En. In addition it is possible to enhance
the fraction of 235U to some extent by ‘enrichment’.

The most efficient moderator would be one in which as much energy was lost by the
neutron at each scattering. This implies a moderator with nuclei of same mass as the
neutron; i.e. one containing lots of single protons, such as water. However protons
actually capture neutrons quite efficiently via p + n → d + γ . The next best is to use
deuterium ⇒ heavy water.

The conditions for stable operation can be derived by noting that this would result in
an equilibrium number of thermal neutrons, N ; i.e.
= 0 = (kN − N )
dN 1
(91)[R]
dt τ
N kN
where is the rate at fissions are being induced by thermal neutrons, and is the
τ τ
rate at which new thermal neutrons are being produced. There are several effects
which contribute to the value of k. These are:-
k = υpfPf ε (92)[G]
where υ is the number of fast neutron released per fission (~2-3), p is the probability
of thermalisation without absorption, f is the probability of subsequent interaction
with a U nucleus (~0.9), Pf is the probability that a fission is induced, ε is a
correction for other fission routes (~1.03). For stable operation k = 1 and once viable
design has been this found allowing this (figure 19) control can be maintained using
adjustable neutron absorbing material to change the thermal neutron flux by changing
p .

Figure 19: Possible configurations for fission reactors showing the major components

(b) Fusion reactors


The biggest problem for man-made nuclear fusion reactors is to overcome the
Coulomb barrier and bring two light nuclei close enough together to get tunnelling to
occur with subsequent fusion. There are several possible interactions which have
been tried, such as:-
pp{ p + p →12 H + e + + γ Q = 0.42 MeV
 12 H + 12 H → 24 He + γ Q = 23.8 MeV

DD  12 H + 12H → 23 He + n Q = 3.3 MeV
2
 1 H + 1 H →1 H + p Q = 4.0 MeV
2 3

DT { 1 H + 1 H → 2 He + n
2 3 4
Q = 17.6 MeV
A popular choice is the DT process as this is enhanced by a resonant state in an
intermediate 25 He nucleus. This reduces the typical temperature required to
~ ~ 200 × 10 6 K . The critical requirement for useful power generation is a
combination of temperature, density (or pressure) and time, called the fusion product.
Since work began in the 1950s this ‘fusion product’ has increased by about 7 orders
of magnitude. It is currently about 1 order of magnitude below that needed. The
main techniques in use are:-
• High-power laser implosion of pellets of D and T
• Magnetic confinement of plasma heated by high currents - tokamak.
• Z-pinch - plasma in which a high current produces a torroidal magnetic field
which compresses the plasma - ICSTM MAGPIE type device.
Figure 20 shows designs for the next generation of a tokamak-type device called
START.

Figure 20: Cut-away diagram of proposed fusion reactor configuration using magnetic
confinement in a ‘spherical tokamak’.
The energy released in fusion reactors comes out in the form of kinetic energy of the
products. For DT this is in neutrons. One way of extracting this energy from the
neutrons is to use the following reactions
n+ 7 Li → 4 He+ 3H + n (has 2.46 MeV less energy - transfered to helium and tritium nuclei)
n+ 7 Li → 4 He+ 3H (plus 4.8 MeV - transfered to helium and tritium nuclei)
Once the energy has been transferred to a charged particle (alpha or tritium nucleus) it
is readily transferred to other forms by electromagnetic interactions. One option for
START is to use a lithium ‘blanket’ (see figure 20) which then gets hot.

Much more information on fusion can be found on http::/www.fusion.org.uk

Vous aimerez peut-être aussi