You are on page 1of 9

Fission and Fusion

Dr Emily Nurse
2017

1 Fission and Fusion: Principles


As you have learnt in the module on “Nuclear Phenomenology”, the nucleons in nuclei are bound together
with an attractive force (the nuclear strong force), therefore it takes energy to split the nucleus up into
individual nucleons. This energy is known as the binding energy of the nucleus. Since the energy of the
bound nucleus plus the binding energy is equal to the energy of the individual nucleons, the total energy of
the nucleus is the energy of the free nucleons minus the binding energy. And since mass is equal to energy
(times c2 ) we say that the mass of a nucleus is smaller than the mass of the individual nucleons, and the
di↵erence in mass is due to the negative binding energy.
The basic principle behind nuclear fission and fusion is to take advantage of the di↵erence in the binding
energy per nucleon between nuclei with di↵erent nucleon numbers. This is seen in Figure 1, which shows that
the binding energy per nucleon versus the nucleon number increases rapidly for the lowest nucleon numbers
(with some spikes seen for magic nuclei), reaches a maximum at 56 Fe , then slowly decreases again. It can
be seen that splitting one of the heaviest nuclei into two lighter nuclei results in a larger total binding energy,
and since this energy is negative it results in a smaller total energy (or mass). The leftover energy can then be
released as kinetic energy, which can be utilised, essentially unlocking some of the energy inside the nucleus.
This is the basic principle behind nuclear fission. It is also clear that if two of the lightest nuclei can be fused
together to form a heavier nucleus, the total (negative) binding energy will be larger and hence the total
energy will be smaller, releasing some kinetic energy. This is the basic principle behind nuclear fusion.

2 Fission
2.1 Basics
For a nucleon number larger than 100 it can be energetically favourable for massive nuclei to split into lighter
daughter nuclei, with the extra energy carried o↵ as kinetic energy of the fission products. An example is:
235
92
U ! 92
37
Rb + 140
55
Cs + 3n + X MeV, (1)

although it should be noted that this is not the only fission process that 235 92
U undergoes. The decay
products of fission will themselves decay, for example via -decay and the resulting nuclei will often give o↵
additional neutrons. These neutrons are known as delayed neutrons as they will be released much later than
the prompt neutrons of Equation 1. We will see in Section 2.4 that these delayed neutrons are vital for the
safe running of nuclear fission reactors.

1
PHAS3224: Nuclear and Particle Physics 2017

Figure 1: Binding energy per nucleon versus nucleon number.

Let us breakdown the energy released using the Semi-Empirical Mass Formula (SEMF). The di↵erence in
binding energies between the initial and final states is given by the di↵erence in energies for each correction
term of the SEMF.1 We will now go through each correction term in turn (although we ignore the very small
pairing term):
1. Volume term: av (AU ARb ACs ) = 46.7 MeV
⇣ ⌘
2/3 2/3 2/3
2. Surface term: +as AU ARb ACs = 160 MeV
✓ ◆
2 2 2
ZU ZRb ZCs
3. Coulomb term: +ac 1/3 1/3 1/3 = +339 MeV
AU ARb ACs
⇣ ⌘
(ZU AU /2)2 (ZRb ARb /2)2 (ZCs ACs /2)2
4. Asymmetry term: +aa AU ARb ACs = +26.1 MeV

where we have used av = 15.56, as = 17.23, ac = 0.697 and aa = 93.14. So the total di↵erence in energy
released at this stage is 158 MeV. In fact the daughter nuclei also release some energy, due to the subsequent
decays discussed above, so the total energy released is actually ⇡ 200 MeV per fission.
1 The total number of nucleons is the same before and after the decay, so it is only the correction terms we need to consider.

2
PHAS3224: Nuclear and Particle Physics 2017

2.2 Spontaneous and induced fission


Spontaneous fission is when the fission process occurs without external action. In the SEMF we have assumed
that the nucleus is spherical, because this minimises the surface area and hence the energy from the surface
term. If we imagine the process of fission starting with the deformation of a spherical nucleus into a prolate
shape, leading to the eventual splitting into two nuclei, then it is important to consider how the binding
energy changes as the nucleus deforms. As the nucleus stretches into a prolate shape, the surface term will
increase and the Coulomb term will decrease (assuming the volume remains the same). If the change in
Coulomb term is larger than the change in the surface term then the deformed shape will be energetically
favourable and the nucleus is unstable, meaning that spontaneous fission can occur. It turns out that this is
2
2as
only the case for ZA ac ⇡ 49, which is only the case for very heavy nuclei with Z > 116 and A 270.
Fission can also occur in lighter nuclei, but this requires penetrating a potential barrier. Since the energy
of the nucleus increases as it deforms (but then decreases again once it has split) there is a barrier that must
2
be overcome. This is shown in Figure 2, which shows as a solid line the energy of a nucleus with ZA < 2a ac .
s

It can be seen that the energy increases as the nucleus deforms then falls o↵ again as it splits (note also the
dashed line, which shows the energy decreasing through the process in the case of the very heavy unstable
2
2as
nuclei with ZA ac ). The di↵erence in the energies for the spherical nucleus and the maximum of the
curve is known as the activation energy (or fission barrier), which is required in order to induce fission.
Spontaneous fission can occur in this case via a quantum mechanical tunnelling through the barrier, however
the probability is extremely small. For example for 238 92
U the transition rate for spontaneous fission is
3 ⇥ 10 24 s 1 compared to about 5 ⇥ 10 18 s 1 for ↵ decay.

Figure 2: Energy of a nucleus during the fission process.

Another possibility is to supply the energy needed to overcome the barrier. This is known as induced
fission and is achieved by supplying a flow of neutrons. Since neutrons are electrically neutral they can
approach the nuclei and be attracted by the strong nuclear force. For heavy nuclei, such as uranium, the
activation energy is only about 6 MeV. This energy can be supplied with a flow of low energy neutrons that
induce neutron capture reactions. They push the nucleus into an excited state above the fission barrier and
it then splits up. When a nucleus absorbs a neutron some energy is released due to the binding energy of
that neutron, if this energy is as large as the activation energy then fission can be induced. The binding
energy of the last added neutron is not the same as the average binding energy per nucleon (this is explained
by the shell model). For uranium it is a little less, and it depends on how adding another neutron a↵ects the

3
PHAS3224: Nuclear and Particle Physics 2017

pairing term. For example, the capture of a neutron by 235 92


U (which leads to 236 92
U ) converts an even-odd
nucleus to a more tightly bound even-even nucleus due to the larger pairing term. Conversely, the capture of
a neutron by 23892
U (which leads to 23992
U ) converts an even–even nucleus to a less tightly bound even–odd
one. The binding energy of the last neutron is 6.5 MeV for 236 92
U and 4.8 MeV for 239 92
U (the di↵erence
between them is due to the di↵erence in the change in the pairing term). The activation energy for fission in
235
92
U is 5 MeV (less than the energy supplied by neutron capture) and that in 238 92
U is 6 MeV (more than
the energy supplied by neutron capture). For this reason fission can be induced in 235 92
U by slow neutrons,
in fact the kinetic energy of the neutrons can be zero as a neutron capture will always provide enough energy
to induce fission. Conversely, for fission to be induced in 238 92
U neutrons with at least 1.2 MeV of kinetic
energy (fast neutrons) are required. Materials such as 23592
U that require only slow neutrons to induce fission
are known as fissile materials, other examples include 233 92
U , 239
94
Pu and 24194
Pu . Materials such as 23892
U
232
that require fast neutrons to induce fission are known as non-fissile materials. Other examples include 90 Th
, 240
94
Pu and 24294
Pu .

2.3 Nucleus–neutron cross sections


Since neutrons are used to induce fission it is useful to look at the nucleus–neutron cross section for uranium,
the most commonly used element in fission reactors. Figure 3 shows the total and fission nucleus–neutron
cross sections for 235 92
U and for 238 92
U . The total cross section includes contributions from elastic and
inelastic scattering, and from neutron absorption. If the neutron is absorbed an excited state is formed that
will either emit one or more particles, e.g. neutrons, protons, ↵ particles; de-excite by emitting photons;
or undergo fission. There are resonant regions in the neutron absorption cross section where the compound
nucleus produced after the neutron absorption is in a region of excitation where its energy levels are well
separated.
There are a number of things to note about Figure 3. Firstly we can see that the fission cross section for
238
92
U is only relevant if the neutron has a kinetic energy above the activation energy of 1.2 MeV, whereas
for fissile 235
92
U the fission cross section is largest for small energy neutrons. In fact the fission cross section
is 84% of the total cross section at 0.1 eV. For 238 92
U the total cross section is dominated by elastic and
inelastic scattering and shows little dependence on energy except in the resonant region between 1 eV and
1 keV, where is becomes likely that a neutron will be captured into a resonance with radiative decay of the
excited state.
For 23592
U the total cross section at low energies is dominated by fission, with a sizeable contribution from
radiative capture of the neutron with the formation of the excited state 236 92
U plus one or more photons.
There is a clear resonance region, where in this case the resonance will decay via fission. At higher energies
the total cross section is dominated by elastic scattering and inelastic excitation of the nucleus.

2.4 Fission chain reactions


We have seen that neutrons can induce fission and also that fission produces neutrons. This leads to the idea
of a chain reaction in which one fission reaction leads to one or more fission reactions due to the produced
neutrons. If the probability that a given neutron induces fission is q and each fission reaction produces an
average of n neutrons, then each neutron will lead to (nq 1) additional neutrons in a time tp , where tp is the
average time before absorption of the neutron occurs. If there are N (t) neutrons present at time t, then at
N (t+ t) N (t)
time t+ t there will be N (t)+N (t) (nq 1) t/tp neutrons present and therefore dN dt = limt!0 =
dN (nq 1)dt R N (t) dN
R tt (nq 1)dt
N (t) (nq 1) /tp giving N (t) = tp . Integrating between t = 0 and t = t gives: N (0) N (t) = 0 tp

4
PHAS3224: Nuclear and Particle Physics 2017

(a) (b)

Figure 3: Nucleus–neutron cross sections versus the neutron kinetic energy for (a) uranium-235 and (b)
uranium-238.

(nq 1)t
giving ln [N (t)] ln [N (0)] = tp leading to:

N (t) = N (0) e(nq 1)t/tp


. (2)

This leads to three possible scenarios, depending on the exact value of nq:

1. For nq < 1, N (t) decreases exponentially, the process is said to be subcritical and the reactions will
soon die out.
2. For nq = 1, N (t) remains constant, the process is said to be critical and the conditions are right for a
sustained, controlled reaction as is required in a nuclear power plant.

3. For nq > 1, N (t) increases exponentially and the process is said to be supercritical. The energy will
grow very rapidly, leading to an explosion, exactly what is needed in a nuclear fission bomb.

2.5 Nuclear fission bombs


Before starting this section I would like to note that the topic of nuclear fission bombs is obviously extremely
sensitive. They have had devastating e↵ects on peoples lives and the teaching of the topic in no way condones
their use, it is simply here to demonstrate how nuclear fission has been utilised and show the physics behind
the weapons.
First we consider the conditions required to create a supercritical reaction. For simplicity we consider
pure 23592
U , which has an average n ⇡ 2.5 and tp ⇡ 10 8 seconds (note that the process in Equation 1
is only one way that fission can occur). For supercritical conditions we require q > 0.4. One of the key
determining factors in the design of a bomb is the size of the metal. If it is small enough that neutrons are
likely to reach the edge of the volume before tp , then the reaction will die out. A neutron with 2 MeV kinetic
energy (which is the average energy of a neutron coming from fission) will move approximately 20 cm in a
straight line in 10 8 seconds. Due to the relatively small kinetic energy compared to the neutron rest mass
(mn = 939.57 MeV/c2 ) we can estimate this using the classical expression for kinetic energy: 12 mn v 2 , so that
2
we have: 2 MeV = 12 ⇥ 939.57 MeV/c2 ⇥ v 2 giving: 0.0043 = vc2 which gives v = 2 ⇥ 107 ms 1 and hence

5
PHAS3224: Nuclear and Particle Physics 2017

the distance travelled is 20 cm. However, the neutron will not travel in a stright line because an average of
6 collisions is expected before inducing fission, and if we assume that the direction of the neutron changes
randomly after each collision, the average distance travelled is actaully 7 cm. However, not all these neutrons
will induce fission, as some will escape the material and some will be captured in nuclei without inducing
fission. This means that a larger sphere is required and it turns out that a radius of 9 cm, corresponding to
a critical mass of 50 kg, is required to ensure a supercritical reaction. A bomb is usually made by starting
with a subcritical mass of material then using a mechanism to increase its size when the bomb is to go o↵.

2.6 Nuclear fission power reactors


In order to produce power from nuclear fission, a critical reaction is required, with nq = 1. As we will see,
any small deviations above unity would lead to a huge increase in the power output in a very short time.
We will consider a thermal reactor, which uses uranium as the fuel and low-energy neutrons to establish
a chain reaction. Natural uranium contains only 0.7% 235 92
U and 99.3% 238 92
U , so a neutron is much more
238
likely to interact with a nucleus of 92 U . A 2 MeV neutron has little chance of inducing fission in natural
uranium. It is more likely to scatter inelastically with a 23892
U nucleus, losing energy in the process (see
Figure 3(b)). After a couple of such interactions the neutron energy will be below the threshold of 1.2 MeV
for inducing fission in 238
92
U . Such a neutron is much more likely to be captured into one of the 238 92
U
resonances, with the emission of photons, before inducing fission in one of the rare 23592
U nuclei. There are
two possible ways to ensure a critical reaction in uranium:
235
1. Enrich the uranium so that it contains 2-3% 92
U , making it more likely that a neutron will induce
fission in a 235
92
U nucleus.
2. Surround the natural uranium fuel in a large volume of moderator material, which has the job of slowing
down the fast neutrons produced in fission via elastic collisions. The slow neutrons are not energetic
enough to be absorbed into a 238
92
U resonance. It will therefore be more likely to eventually meet and
induce fission in a 235
92
U nucleus, the cross section for which is high at low energies (see Figure 3).
Usually heavy water (D2 O) or graphite are used as moderator materials.
Vital to the safe running of a nuclear fission reactor is that it operates with precisely nq = 1. This is
achieved with control rods that are mechanically inserted into the reactor whenever the reaction needs to be
reduced (and removed if the reaction needs to be increased). They are usually made of cadmium, which has
a very high absorption cross section for neutrons. The problem with this mechanism is that tiny increases to
nq, from e.g. fluctuating temperatures, would very rapidly lead to an explosive reaction. In order to see this
we consider Equation 2. Typically tp ⇡ 10 3 seconds in natural uranium (note it is much longer than the
equivalent time in pure 23592
U due to the smaller cross sections), so for a modest 0.1% increase in nq such
that nq = 1.001, the reactor flux would increase by e60 ⇡ 1026 in only one minute, which is clearly a highly
unstable system. Mechanical insertion of the control rods is not possible on the timescales required to control
such rapid increases in flux. It is therefore necessary to have a much smaller rate of increase in a safe nuclear
reactor. The key to this is to consider the a↵ect of delayed neutrons, which appear on average ⇡ 13 seconds
after the fission. These delayed neutrons come from the fission of decay products that have undergone a
series of decays. Taking account of delayed neutrons we need to replace nq with (nprompt + ndelayed ), where
nprompt is the number of prompt neutrons per fission and ndelayed is the number of delayed neutrons per
fission. To ensure that reactions remain constant we have the requirement (nprompt + ndelayed ) q = 1, where
ndelayed ⇡ 0.02, giving a roughly 1% correction to nprompt . The idea is to keep nprompt q far enough below 1

6
PHAS3224: Nuclear and Particle Physics 2017

that small variations in q cannot lead to an unstable system. However (nprompt + ndelayed ) q is kept very close
to 1, and it is the time scale of the delayed neutrons (⇡ 13 seconds) that dictate the growth of the neutron
flux in the event of a deviation from criticality. These timescales are manageable for mechanical control of
the reactor via the insertion of neutron absorbing control rods.

3 Fusion
3.1 Basics
Fusion is the opposite process to fission, where two very light nuclei fuse to form a heavier nucleus, releasing
energy due to the fact that the (negative) binding energy per nucleon is larger for the fused nucleus. The
energy released in fusion is less than that in fission, but the light nuclei are much more abundant in nature,
making it an attractive potential for power generation.

3.2 Coulomb barrier


The practical problem with fusion comes from the fact that two positively electrically charged nuclei will
repel each other due to the Coulomb repulsion, stopping them getting close enough for the strong nuclear
force to take over allowing them to fuse. The Coulomb potential between two nuclei gives the amount of
energy required to overcome this Coulomb barrier and is given by:
0
1 ZZ e2
VC = ,
4⇡✏0 R + R0
0 0 1
where Z and Z are the two atomic numbers and R and R are their e↵ective radii. If we take R = 1.2A 3 fm
(from the “Nuclear Phenomenology” module) then we obtain:
0 0
e2 h̄cZZ ZZ
VC = h i = 1.198 h i MeV,
4⇡✏0 h̄c 1.2 A1/3 + (A )
0 1/3
fm A1/3 + (A0 )
1/3

2
where we have used ↵ = 4⇡✏e0 h̄c = 137
1
and h̄c = 0.197 GeV fm (see the “Basic Ideas in Particle Physics”
0 0
notes). For simplicity we set A ⇡ A ⇡ 2Z ⇡ 2Z , then VC ⇡ 0.15A5/3 MeV and for A ⇡ 8, VC ⇡ 4.8 MeV.
This would be the amount of energy required to overcome the Coulomb barrier, and would be smaller for
lighter nuclei.
In order to overcome this barrier a confined mixture of nuclei can be heated enough that the kinetic
energy supplies the required energy. Using the relationship between particle energy and temperature in a
medium, E = kB T , with Boltzmann’s constant kB = 8.6 ⇥ 10 5 eV K 1 , we obtain T ⇡ 5.610 K for a particle
energy of 4.8 MeV. This is a very high temperature! Actually, fusion will occur at lower temperatures than
this for two reasons:
1. Quantum tunnelling: similar to ↵-decay, discussed in the “Nuclear Phenomenology” module, the prob-
G
ability to penetrate an
penergy barrier via quantum tunnelling is e , where G is the energy dependent
Gamow factor: G = EG /E, where EG increases as the barrier increases and E is the energy of the
particles.
2. A collection of nuclei with a given mean energy will have a Maxwellian distribution of energies about
the mean and the distribution of the energies will have the form e E/kB T .

7
PHAS3224: Nuclear and Particle Physics 2017

So the e↵ect from quantum mechanical tunnelling increases with energy and that from the energy distribution
of the particles decreases with energies. When these two e↵ects are combined we end up with a reaction rate
that peaks at a certain energy and there is a small range of energies within which fusion takes place at a
non-negligible rate (see the lecture slides for a distribution of the reaction rate versus energy).

3.3 Stellar fusion


The energy of the sun comes from nuclear fusion reactions, and mostly from the proton–proton chain. The
temperature of the sun is 107 K and all the material is fully ionised and is referred to as a plasma. The
nuclei of hydrogen atoms (which are just protons) are the starting point for the process, which has the three
following steps:
1. Two hydrogen nuclei (protons) fuse to form a deuterium nucleus:
1
1
H + 11 H ! 21 H + e+ + ⌫e + 0.42 MeV.
1
2. The deuterium fuses with more hydrogen to produce helium and a photon: 1
H + 21 H ! 32 He + +
5.49 MeV.
3. Two helium nuclei (from the above two steps occuring twice) fuse to form an ↵ particle plus some extra
hydrogen nuclei: 32 He + 32 He ! 42 He + 11 H + 11 H + 12.86 MeV.
In each step the energy released is the di↵erence in binding energies between the initial and final states. The
relevant binding energies are: B 21 H = 2.224 MeV, B 32 He = 7.718 MeV and B 42 He = 28.3 MeV. For
the first step the di↵erences in the masses of the particles in the initial and final states must also be taken
into account: me+ = 0.511MeV/c2 , mn = 939.566 MeV/c2 and mp = 938.272 MeV/c2 . You can convince
yourselves that accounting for these gives the energy di↵erences seen above. The extra energy is in the kinetic
energy of the particles. The first step is slow as it involves the weak interaction and it therefore sets the
scale for the long lifetime of the sun. For a complete chain the first two steps occur twice providing the two
helium nuclei for the third step. The total energy released is: 2 ⇥ (0.42 + 5.49) + 12.86 = 24.68 MeV and the
overall interction is given by:

4 1
1
H ! 42 He + 2e+ + 2⌫e + 2 + 24.68 MeV.

The produced positrons will annhilate with electrons in the plasma providing an extra 1.02 MeV of energy
per positron so the total energy is 26.72 MeV. However, each neutrino will carry o↵ an average of 0.26 MeV of
energy, which will be lost in space. The rest of the energy is transported to the surface of the sun and emitted
as photons or ejected high-energy particles. The total energy radidated from the sun from the proton–proton
chain is therefore 6.55 MeV per proton. The proton–proton chain is the dominate fusion process powering
the sun, but other processes involving the fusion of heavier nuclei also occur producing additional energy and
changing the composition of elements in the sun to include the heavier elements.

3.4 Fusion reactors


Due to the abundance of light nuclei the idea of utilising the energy from fusion as a source of power is
a very attractive one. The proton–proton interactions that power the sun are very slow (due to the weak
interaction) so cannot be used. Possible candidates for fusion reactor processes are deuterium–deuterium
fusion and tritium–dueterium fusion. The former proceeds via: 21 H + 21 H ! 32 He + n + 3.27 MeV or 21 H + 21 H !
3
1
H + p + 4.03 MeV and the latter via: 21 H + 31 H ! 42 He + n + 17.62 MeV, which gives o↵ more energy and also

8
PHAS3224: Nuclear and Particle Physics 2017

has the advantage of having a larger cross section than deuterium–deuterium fusion. In all these reactions the
energy produced comes from the di↵erence in binding energies before and after the reaction. The problem
with tritium–deuterium fusion is the low abundance of natural tritium, which has a half-life of only 17.7
years. By contrast deuterium is found in huge quantities in sea water.
Nonetheless, due to its advantages, tritium–deuterium fusion is considered as a candidate for fusion reac-
tors. The cross section for this process is reasonable when the nuclei have an energy of 20 keV, corresponding
2⇥104
to a temperature of T = kEB = 8.6⇥10 5 = 2 ⇥ 10
8
K. The main problem with practical fusion is how to
contain plasmas at such high temperatures. Any material container would vaporize at these temperatures.
The solution is to contain the plasma using either:
1. magnetic confinement where the charged particles in the plasma follow a helix path as they curve
round a magnetic field with a direction that points in a circle around a donut shape or
2. intertial confinement where pulsed lasers bombard small pellets of a tritium–deuterium mixture in
many directions at the same time at very high energies.

Fusion research aims to try and make the ratio of energy output to energy input (required to reach such high
temperatures) greater than one so that fusion is feasible for a source of power. This is known as the Lawson
criteria. This ratio is larger for longer confinement times and higher particle densities. A value greater than
one has not yet been reached but research is still ongoing.

You might also like