You are on page 1of 15

Chapter: 7 Nuclear Energy and Forces.

[Grab your reader’s attention with a great quote from the document or use this
space to emphasize a key point. To place this text box anywhere on the page,
just drag it.]
[Cite your source here.]

Nuclear Energy:
Nuclear energy, also called atomic energy, energy that is released in significant amounts in
processes that affect atomic nuclei, the dense cores of atoms. It is distinct from the energy
of other atomic phenomena such as ordinary chemical reactions, which involve only the
orbital electrons of atoms. One method of releasing nuclear energy is by controlled nuclear
fission in devices called reactors, which now operate in many parts of the world for the
production of electricity. Another method for obtaining nuclear energy, controlled nuclear
fusion, holds promise but has not been perfected by 2020. Nuclear energy has been
released explosively by both nuclear fusion and nuclear fission. See also nuclear power.

Figure: 7.1 Diagram showing the difference between nuclear fission and nuclear fusion.
Nuclear fission is used in nuclear reactors to produce energy for electrical power and similar
applications. It also was used to create the atomic bomb. Fusion is used to create
thermonuclear weapons and holds promise to produce electricity.
In nuclear fission the nucleus of an atom, such as that of uranium or plutonium. breaks up
into two lighter nuclei of roughly equal mass. The process may take place spontaneously in
some cases or may be induced by the excitation of the nucleus with a variety of particles
(e.g., neutrons, protons, deuterons, or alpha particles) or with electromagnetic radiation in
the form of gamma rays. In the fission process a large quantity of energy is released,
radioactive products are formed, and several neutrons are emitted. These neutrons can
induce fission in a nearby nucleus of fissionable material and release more neutrons that
can repeat the sequence, causing a chain reaction in which a large number of nuclei
undergo fission and an enormous amount of energy is released. If controlled in a nuclear
reactor, such a chain reaction can provide power for society’s benefit. If uncontrolled, as in
the case of the so-called atomic bomb, it can lead to an explosion of awesome destructive
force.

Figure: 7.2 The Taiwan nuclear power plant, using pressurized-water reactors, in
Lianyungang, Jiangsu province, China.
Nuclear fusion is the process by which nuclear reactions between light elements form
heavier elements. In cases where the interacting nuclei belong to elements with low atomic
numbers (e.g., hydrogen [atomic number 1] or its isotopes deuterium and tritium),
substantial amounts of energy are released. The vast energy potential of nuclear fusion was
first exploited in thermonuclear weapons, or hydrogen bombs, which were developed in the
decade immediately following World War II. The potential peaceful applications of nuclear
fusion, especially in view of the essentially limitless supply of fusion fuel on Earth, have
encouraged an immense effort to harness this process for the production of power.
Although practical fusion reactors have not been built yet, the necessary conditions of
plasma temperature and heat insulation have been largely achieved, suggesting that fusion
energy for electric-power production is now a serious possibility. Commercial fusion
reactors promise an inexhaustible source of electricity for countries worldwide.
Stability of Nuclei:
Not all combinations of neutrons and protons form stable nuclei. In general, light nuclei
(A ˂ 20) contain approximately equal numbers of neutrons and protons, while in heavier
nuclei the proportion of neutrons becomes progressively greater. This is evident from Fig.
7.3, which is a plot of N versus Z for stable nuclides.
The tendency for N to equal Z follows from the existence of nuclear energy levels. Nucleons,
which have spins of 1/2, obey the exclusion principle. As a result, each nuclear energy level
can contain two neutrons of opposite spins and two protons of opposite spins.

Figure: 7.3 Neutron-proton diagram for stable nuclides.


Energy levels in nuclei are filled in sequence, just as energy levels in atoms are, to achieve
configurations of minimum energy and therefore maximum stability. Thus, the boron
isotope 125B has more energy than the carbon isotope 126C because one of its neutrons is in a
higher energy level, and 125B is accordingly unstable (Fig. 7.4).

Figure: 7.4 Simplified energy-level


diagrams of some boron and carbon
isotopes. The exclusion principle
limits the occupancy of each level
to two neutrons of opposite spin and two protons of opposite spin. Stable nuclei have
configurations of minimum energy.

If created in a nuclear reaction, a 125B nucleus changes by beta decay into a stable 126C
nucleus in a fraction of a second.
The preceding argument is only part of the story. Protons are positively charged and repel
one another electrically. This repulsion becomes so great in nuclei with more than 10
protons or so that an excess of neutrons, which produce only attractive forces, is required
for stability. Thus, the curve of Fig. 7.3 departs more and more from the N = Z line as Z
increases. Even in light nuclei N may exceed Z, but (except in 11 H and 32 He) is never smaller;
11 11
5 B is stable, for instance, but not 6C .

Sixty percent of stable nuclides have both even Z and even N; these are called “even even”
nuclides. Nearly all the others have either even Z and odd N (even-odd nuclides) or odd Z
and even N (odd-even nuclides), with the numbers of both kinds being about equal. Only
five stable odd-odd nuclides are known: 21 H , 63 Li, 105Be , 147 N , and 180
73Ta . Nuclear abundances

follow a similar pattern of favouring even numbers for Z and N. Only about one in eight of
the atoms of which the earth is composed has a nucleus with an odd number of protons, for
instance.
These observations are consistent with the presence of nuclear energy levels that can each
contain two particles of opposite spin. Nuclei with filled levels have less tendency to pick up
other nucleons than those with partly filled levels and hence were less likely to participate in
the nuclear reactions involved in the formation of the elements.

Nuclear Decay:
Nuclear forces are limited in range, and as a result nucleon interact strongly only with their nearest
neighbours. This effect is referred to as the saturation of nuclear forces. Because the coulomb
repulsion of the protons is appreciable throughout the entire nucleus, there is a limit to the ability of
neutrons to prevent the disruption of a large nucleus. This limit is represented by the bismuth
209
isotope 83Bi , which is the heaviest stable nuclide. All nuclei with Z ˃ 83 and A ˃ 209 spontaneously
transform themselves into lighter ones through the emission of one or more alpha particles, which
4
are 2He nuclei:
A A −4
Z X→ Z−2 Y + 42He (Alpha Decay)

Parent nucleus → Daughter nucleus + Alpha particle.


Since an alpha particle consists of two protons and two neutrons, an alpha decay reduces
the Z and the N of the original nucleus by two each. If the resulting daughter nucleus has
either too small or too large a neutron/proton ratio for stability, it may beta-decay to a
more appropriate configuration. In negative beta decay, a neutron is transformed into a
proton and an electron is emitted:
−¿ ¿

n0 → P+ ¿+ ⅇ ¿
(Beta Decay)
In positive beta decay, a proton becomes a neutron and a positron is emitted:
0 + ¿¿

P+¿→ n +ⅇ ¿
(Positron Emission)

Thus, negative beta decay decreases the proportion of neutrons and positive beta decay
increases it. A process that competes with positron emission is the capture by a nucleus of
an electron from its innermost shell. The electron is absorbed by a nuclear proton which is
thereby transformed into a neutron:
0
−¿→n ¿

P+¿+ⅇ ¿
(Electron Capture)

Binding Energy:
Binding energy is defined as the amount of energy that must be supplied to a nucleus to
separate its nuclear particles (nucleons). It can also be understood as the amount of energy
that would be released if the nucleus was formed from the separate particles. Since 1 u is
equivalent to 931.5 MeV of energy, the binding energy can be calculated by the mass
difference between the nucleus and the sum of those of the free nucleons, including the
mass of electrons associated with protons.

B.E(Z, A) = { Z m p + Z m e + ( A−Z ) mn −M ( Z , N )}c2

Where,
mp = mass of proton (1.0072764 u)
mn = mass of neutron (1.008665 u)
me = mass of electron (0.000548597 u)
mH = mp +me = mass of hydrogen atom = (1.007825 u)

Separation Energy:
The useful and interesting property of the binding energy is the neutron and proton
separation energies. The neutron separation energy Sn is the amount of energy required to
remove a neutron from a nucleus, AZ X (sometimes called the binding energy of the last
neutron). This is equal to the difference in binding energies between AZ X and A −1Z X .

Sn = B.E ( AZ X ) – B.E ( A −1Z X )

Sn = {M ( A −1Z X ) – M ( AZ X ) + mn} c2

Similarly one can define proton separation energy Sp as the energy needed to remove a
proton from a nucleus AZ X (also called the binding energy of the last proton) which convert
to another nuclide, AZ −1
−1Y and can be calculated as follows.

A A −1
Sp = B.E ( Z X ) – B.E ( Z −1Y )

Sp = {M ( AZ −1 A
−1Y ) – M ( Z X ) + m ( H)} c
1 2
The Hydrogen mass appears in this equation instead of proton mass, since the atomic mass
is m (1H) = mp + me. The neutron and proton separation energies are analogous to the
ionization energies in atomic physics, in terms of the binding of the outermost valance
nucleons. Just like the atomic ionization energies, the separation energies show evidence for
nuclear shell structure that is like atomic shell structure.

Binding Energy Per Nucleon:


As with many other nuclear properties that will be discussed in the coming sections, we gain
valuable clues to nuclear structure from a systematic study of nuclear binding energy. As the
number of particles in a nucleus increases, the total binding energy also increases. The rate
of increase, however, is not uniform. This lack of uniformity results in a variation in the
amount of binding energy associated with each nucleon within the nucleus. In other wards
since the binding energy increases more or less linearly with atomic mass number A, it is
convenient to show the variation of the average binding energy per nucleon, BE/A as
function of A. Fig. 7.5 shows the average BE/A as plotted versus atomic mass number A.
Fig. 7.5 illustrates that as the atomic mass number increases, the binding energy per
nucleon increases almost linearly for light elements (except for ), then rapidly for A < 60
reaches a maximum value of 8.79 MeV at A = 56 (close to iron) and decreases to about 7.6
MeV for A = 238. The average BE/A of most nuclei is, to within 10%, about 8 MeV.
The general shape of the BE/A curve can be explained using the general properties of
nuclear forces. Very short-range attractive nuclear forces that exist between nucleons hold
the nucleus together. On the other hand, long-range repulsive electrostatic (coulomb)
forces that exist between all the protons in the nucleus are forcing the nucleus apart. So,
the nuclear forces are very short range of the order of the diameter of one nucleon, or they
saturate to pairs of nucleons (two protons and two neutrons) to form α-cluster. This is clear
in light stable nuclei for A < 20 where the sharp rise appears to be off the curve. Those
specific light stable nuclei are 42He , 84 Be , 126C , 168O. This is due to the fact of higher BE/A of 42He
particle (or α-cluster) bounded in the nucleus and the other (A = 4n nuclei for n=1, 2 ...)
stable nuclei for A < 20 compared to their neighbors. In other words, the 4n nuclei for n=1,
2, .. trend to make α-particles.
Figure: 7.5 Binding Energy per Nucleon vs. Mass Number.
As the atomic number and the atomic mass number increase, the repulsive electrostatic
forces within the nucleus increase due to the greater number of protons in the heavy
elements. To overcome this increased repulsion, the proportion of neutrons in the nucleus
must increase to maintain stability. This increase in the neutron-to-proton ratio only
partially compensates for the growing proton proton repulsive force in the heavier, naturally
occurring elements. Because the repulsive forces are increasing, less energy must be
supplied, on the average to remove a nucleon from the nucleus. The BE/A has decreased.
The BE/A of a nucleus is an indication of its degree of stability. Generally, the more stable
nuclides have higher BE/A than the less stable ones. The increase in the BE/A as the atomic
mass number decreases from 260 to 60 is the primary reason for the energy liberation in
the fission process. In addition, the increase in the BE/A as the atomic mass number
increases from 1 to 60 is the reason for the energy liberation in the fusion process, which is
the opposite reaction of fission.
The heaviest nuclei require only a small distortion from a spherical shape (small energy
addition) for the relatively large coulomb forces forcing the two halves of the nucleus apart
to overcome the attractive nuclear forces holding the two halves together. Consequently,
the heaviest nuclei are easily fissionable compared to lighter nuclei.
Mass Spectroscopy:
Binding energies may be calculated if masses are measured accurately. One way of doing
this is by using the techniques of mass spectroscopy. The principle of the method is shown
in Fig. 7.6.
All mass spectroscopes begin with ion source region, which produces a beam of ionized
atoms or molecules. Often a vapor of the material under study is bombarded with electrons
to produce ions; in other cases, the ions can be formed as a result of a spark discharge
between electrodes coated with the material. Ions emerging from the source have broad
range of velocities, as might be expected for a thermal distribution, and many different
masses might be included.

Figure: 7.6 Schematic diagram of a mass spectrometer.

A source of ions of charge q, containing various isotopes passes through a second region
called velocity selector where there are uniform electric (E) and magnetic (B 1) fields at right
angles. The electric field will exert a force.
FE = qE
in one direction and the magnetic field will exert a force,
FB = qυB1
in the opposite direction, where υ is the speed of the ions. By balancing these forces, ions of
a specific speed υ = E/B1 can be selected and allowed to pass through a collimating slit. Ions
with other velocities (shown as dashed lines) are deflected. The beam is then allowed to
continue through a third region momentum selector where a second uniform magnetic field
B2 will be exerted on the beam, then the beam will be bent into a circular path of radius r,
given by:

m  qB2r
and since q, B2 and v are fixed, particles with a fixed ratio q/m will bend in a path with a
unique radius. Hence, isotopes may be separated and focused onto a detector (e.g. a
photographic plate) with a proportional radii r.
q E
=
m B1 B2 r

In the common case where B1= B2= B.


q E
= 2
m B r
q B2 r
or m=
E
In practice, to achieve high accuracy, the device is used to measure mass differences rather
than absolute values of mass. We could calibrate for one particular mass, and then
determine all masses by relative measurements. The fixed point on the atomic mass scale is
12
C, which is taken to be exactly 12.000000 u.
To determine the mass of another atom, such as 1H, it would need to make considerable
changes in E, B1 and B2, and it is perhaps questionable whether the calibration would be
valid to one part in 106 over such a range. It would be preferable to measure the smaller
difference between two nearly equal masses.
For example, let as set the apparatus for mass 128 and measure the difference between the
molecular masses of C9H20 (nonane) and C10H8 (naphthalene). This difference is measured to
be ∆m = 0.09390032 ± 0.00000012 u. Neglecting corrections for the difference in the
molecular binding energies of the two molecules, which is of the order of 10-9 u, the mass
difference can be calculated as follows:
∆m = M(C9H20) – M(C10H8) = 12M(1H) – M(12C)

Thus m(1H) = 1/12 [ M(12C) + ∆m]


= 1.00000000 + 1/12 ∆m
= 1.00782503 ± 0.00000001 u
Then set the apparatus for mass 28 and determine the difference between C2H4 and N2:
∆m = M(C2H4) – M(N2)
= 2M(12C) +4m(1H) – 2M(14N)
= 0.025152196 ± 0.00000003 u
This system of measuring small differences between close lying masses is known as the mass
doublet method.
In addition, the conventional mass spectroscopy cannot be used to find the masses of very
short-lived nuclei and in these cases, the masses are determined from kinematical analysis
of nuclear reactions.
Energy Levels of Atoms:
The electrons that circle the nucleus move in well-defined orbits. Some of these electrons
are more tightly bound in the atom than others. For example, only 7.38 eV is required to
remove the outermost electron from a lead atom, while 88,000 eV is required to remove
the innermost electron. The process of removing an electron from an atom is called
ionization, and the energy required to remove the electron is called the ionization energy.
In a neutral atom (number of electrons = Z), it is possible for the electrons to be in a variety
of different orbits, each with a different energy level. The state of lowest energy is the one
in which the atom is normally found and is called the ground state. When the atom
possesses more energy than its ground state energy, it is said to be in an excited state.
An atom cannot stay in the excited state for an indefinite period. An excited atom will either
eventually transfer to a lower-energy excited state or directly to its ground state. This is
performed by emitting a discrete bundle of electromagnetic energy called an x-ray. The
energy of the x-ray will be equal to the difference between the energy levels of the atom
and will typically range from several eV to 100,000 eV in magnitude.
Energy Levels of The Nucleus:
The nucleons in the nucleus of an atom, like the electrons that circle the nucleus, exist in
shells that correspond to energy states. The difference between them is that in contrast to
electrons, the nucleons have no center to orbit around it, as will be clarified in Chapter-7.
The energy shells of the nucleons in the nucleus are less defined and less understood than
those of the electrons of the atom. There is a state of lowest energy (the ground state) and
discrete possible excited states for a nucleus. Where the discrete energy states for the
electrons of an atom are measured in eV or keV, the energy levels of the nucleus are
considerably greater and typically measured in MeV.
A nucleus that is in the excited state will not remain at that energy level for an indefinite
period. Like the electrons in an excited atom, the nucleons in an excited nucleus will make a
transition towards their lowest energy configuration, and in doing so, emit a discrete bundle
of electromagnetic radiation called a gamma ray (γ-ray). The only differences between x-
rays and γ-rays are their energy levels and whether they are emitted from the electron shell
or from the nucleus.
The ground state and the excited states of a nucleus can be depicted in a nuclear energy-
level diagram. The nuclear energy-level diagram consists of a stack of horizontal bars, one
bar for each of the excited states of the nucleus. The vertical distance between the bar
representing an excited state and the bar representing the ground state is proportional to
the energy level of the excited state with respect to the ground state. This difference in
energy between the ground state and the excited state is called the excitation energy of the
excited state. The ground state of a nuclide has zero excitation energy. The bars for the
excited states are labelled with their respective energy levels. Fig. 2. 6 is the energy level
diagram for Nickel-60.
Figure: 7.7 The energy level diagram for nickel-60.

Nuclear Forces and Interactions:


There is a dominant short-range part of the nucleon nucleon interaction potential, which is
central and provides the overall shell-model potential. A part, whose range is much smaller
than the nuclear radius, tends to make the nucleus spherical and to pair up with nucleons. A
second part, whose range is of the order of the nuclear radius, tends to distort the nucleus.
There is a spin-orbit interaction and spin-spin interaction. The force is charge independent
(Coulomb interaction excluded).
One of the aims of nuclear physics is to calculate the energies and quantum numbers of
nuclear bound states. In atomic physics, one can do this starting from first principles. In fact,
Coulomb’s law (or more generally the equations of electromagnetism) determines the
interactions between electrons and nuclei. Spin corrections and relativistic effects can be
calculated perturbatively to a very good accuracy because of the smallness of the fine
structure constant α = e2 /(4πεħc) ≈ 1/137. Together with the Pauli exclusion principle which
leads to the shell structure of electron orbitals, these facts imply that one can calculate
numerically spectra of complex atoms despite the difficulties of the many-body problem.
Unfortunately, none of these holds in nuclear physics. Forces between nucleons are neither
simple nor fully understood. One of the reasons for this is that the interactions between
nucleons are “residuals” of the fundamental interactions between quarks inside the
nucleons. In that sense, nuclear interactions are like Van der Waals forces between atoms or
molecules, which are also residual or “screened” Coulomb interactions. For these reasons,
forces between nucleons are described by semi-phenomenological forms, e.g., the potential
proposed by Yukawa in 1939, which are only partly deduced from fundamental principles.
The subject of nucleon–nucleon potentials is very complex, and we will give only a
qualitative discussion. Our basic guidelines are the following facts:
1. Protons and neutrons are spin 1/2 fermions and, therefore, obey the Pauli Exclusion
Principle which states, in a closed system no two identical particles can be in the
same energy state. More details will be clarified in the following sections.
2. Nuclear forces are attractive and strong, since binding energies are roughly a 106
times the corresponding atomic energies. They are, however, short range forces (a
few fm). The combination of strength and short range makes 2 nucleon systems only
marginally bound but creates a rich spectrum of many-nucleon states.
3. Nuclear forces are "charge independent". They are blind to the electric charge of
nucleons. If one were to “turn off” Coulomb interactions, the nuclear proton–proton
potential would be the same as the neutron–neutron potential. A simple example is
given by the binding energies of isobars, such as tritium and helium-3:
BE (3H) = 8.492MeV > BE (3He) = 7.728MeV. If the difference ∆BE = 0.764MeV is
attributed to the Coulomb interaction between the two protons in 3He, ∆BE = e2 <
1/r12 > /4πε, one obtains a very reasonable value for the mean radius of the system:
R ≈ 2 fm (this can be calculated or measured by other means). We shall come back to
this question in a more quantitative way when we discuss isospin.
4. Nuclear forces saturate. As we will see in the next section, this results in the volumes
and binding energies of nuclei being additive and, in first approximation,
proportional to the mass number A. This is a remarkable fact since it is reminiscent
of a classical property and not normally present in quantum systems. It appears as if
each nucleon interacts with a given fixed number of neighbours, whatever the
nucleus is.
Reversing the order of inference, physicists could have derived the form of Coulomb’s law
from the spectrum of bound states of the hydrogen atom. This is not possible in nuclear
physics because there is only one two-nucleon bound state, the deuteron. In the next
subsection, we will find that there is much to be learned from this fact, but it will not be
enough to derive the nucleon–nucleon potential in all its details. To do this, we will need to
attack the more difficult problem of nucleon–nucleon scattering.

Meson Theory of Nuclear Forces/ Yukawa’s Theory:


As we saw how a molecule is held together by the exchange of electrons between adjacent
atoms. Is it possible that a similar mechanism operates inside a nucleus, with its component
nucleons being held together by the exchange of particles of some kind among them?
The first approach to this question was made in 1932 by Heisenberg, who suggested that
electrons and positrons shift back and forth between nucleons. A neutron, for instance,
might emit an electron and become a proton, while a proton absorbing the electron would
become a neutron. However, calculations based on beta-decay data showed that the forces
resulting from electron and positron exchange by nucleons would be too small by the huge
factor of 1014 to be significant in nuclear structure.
The Japanese physicist Hideki Yukawa was more successful with his 1935 proposal that
particles intermediate in mass between electrons and nucleons are responsible for nuclear
−¿¿

forces. Today these particles are called pions. Pions may be charged ( π +¿, π ¿ ) or neutral ( π 0
) and are members of a class of elementary particles collectively called mesons. The word
pion is a contraction of the original name meson.
According to Yukawa’s theory, every nucleon continually emits and reabsorbs pions. If
another nucleon is nearby, an emitted pion may shift across to it instead of returning to its
parent nucleon. The associated transfer of momentum is equivalent to the action of a force.
Nuclear forces are repulsive at very short range as well as being attractive at greater
nucleon-nucleon distances; otherwise the nucleons in a nucleus would mesh. One of the
strengths of the meson theory of such forces is that it can account for both these properties.
Although there is no simple way to explain how this comes about, a rough analogy may
make it less mysterious.
Let us imagine two boys exchanging basketballs (Fig. 7.8). If they throw the balls at each
other, the boys move backward, and when they catch the balls thrown at them,

Figure: 7.8 Attractive and repulsive forces can both arise from particle exchange.

their backward momentum increases. Thus, this method of exchanging basketballs has the
same effect as a repulsive force between the boys. If the boys snatch the basketballs from
each other’s hands, however, the result will be equivalent to an attractive force acting
between them.
A fundamental problem presents itself at this point. If nucleons constantly emit and absorb
pions, why are neutrons and protons never found with other than their usual masses? The
answer is based upon the uncertainty principle. The laws of physics refer to measurable
quantities only, and the uncertainty principle limits the accuracy with which certain
combinations of measurements can be made. The emission of a pion by a nucleon which
does not change in mass—a clear violation of the law of conservation of energy—can take
place provided that the nucleon reabsorbs it or absorbs another pion emitted by a
neighbouring nucleon so soon afterward that even in principle it is impossible to determine
whether or not any mass change has actually been involved.
From the uncertainty principle in the form,

ΔEΔt ≥
2
an event in which an amount of energy ΔE is not conserved is not prohibited so long as the
duration of the event does not exceed ℏ ∕ 2 ΔE. This condition lets us estimate the pion
mass.
Let us assume that a pion travels between nucleons at a speed of v c (actually v
m
˂ c, of course); that the emission of a pion of mass π represents a temporary energy
discrepancy of ΔE m π c2 (this neglects the pion’s kinetic energy); and that ΔE Δt
ℏ. Nuclear forces have a maximum range r of about 1.7 fm, and the time t
needed for the pion to travel this far (Fig. 7.9) is
r r
Δt =
ν c

Figure: 7.9 The uncertainty principle


permits the creation, transfer, and
disappearance of a pion to occur
without violating conservation of
energy provided that the sequence
takes place fast enough. Here a
positive pion emitted by a proton is
absorbed by a neutron; as a result,
the proton becomes a neutron and
the neutron becomes a proton.

We therefore have,
ΔEΔt ℏ

( mπ C2 ) ( Cr ) h
h

rC
which gives a value for m π of

1.05× 10−34 J . s
mπ 10 8 m 2 ×10−28 kg
(1.7 ×10−15 m)(3× )
s
This rough figure is about 220 times the rest mass me of the electron.

Discovery of the Pions:


A dozen years after Yukawa’s proposal, particles with the properties he had predicted were
discovered. The rest mass of charged pions is 273 me and that of neutral pions is 264 me,
not far from the above estimate.
Two factors contributed to the belated discovery of the free pion. First, enough energy must
be supplied to a nucleon so that its emission of a pion conserves energy. Thus at least mc 2
of energy, about 140 MeV, is required. To furnish a stationary nucleon with this much
energy in a collision, the incident particle must have considerably more kinetic energy than
mc 2 in order that momentum as well as energy be conserved. Particles with kinetic
energies of several hundred MeV are therefore required to produce free pions, and such
particles are found in nature only in the diffuse stream of cosmic radiation that bombards
the earth. Hence the discovery of the pion had to await the development of sufficiently
sensitive and precise methods of investigating cosmic-ray.

Exercises:

You might also like