You are on page 1of 14

Einstein, mass and energy

Einstein’s work on special relativity led him to publish a paper in 1905. This paper suggested that energy
and mass were different ways of expressing the same thing – that energy and mass were
interchangeable and linked using the equation, E=mc 2, where E is energy (J), m is change in mass (kg)
and c is the speed of light, 3 × 108 ms−1. By 1932 his ideas were proved experimentally by Cockcroft and
Walton.

So what does this equation mean? If we use a helium nucleus as an example, the helium nucleus
contains four nucleons – two protons and two neutrons. The mass of the helium nucleus is very slightly
smaller than the mass of its separate nucleons.

As the helium nucleus forms, some mass is converted to energy and released. Calculating the energy
released when an alpha particle is formed is straightforward:

 mass of a proton is 1.6726 × 10−27 kg


 mass of a neutron is 1.6749 × 10−27 kg
 mass of two protons and two neutrons is 6.6950 × 10 −27 kg
 measured mass of a helium nucleus is 6.6337 × 10 −27 kg
 mass difference is 6.13 × 10−29 kg

Using E=mc 2, the energy released when a single alpha particle is formed is 5.5 × 10 −12 J, or 34MeV. You
can see that, in Einstein’s own words, ‘a very small amount of mass may be converted into a very large
amount of energy and vice versa’.

Work must be done to overcome the very strong nuclear forces that bind the nucleons together and pull
a helium nucleus apart. The energy put in to do this creates the extra mass.

A nucleus of Z protons and N neutrons has a mass that is less than the mass of the protons and
neutrons that make it up. This difference in mass is called the mass defect, where mass defect
Δm=Z m p+ N mn −M nucleus measured in kg or atomic mass units (u).

Since mass and energy are interchangeable, we can also express mass as energy. Binding energy is the
energy that corresponds to the mass defect, and is related to the mass defect using
binding energy=mass defect × c 2 .
Binding energy is the energy that would have to be supplied to the nucleus to separate it back into its
constituent protons and neutrons. The binding energy can be expressed in J or in MeV.

Mass-energy Equivalence

The equivalence of mass and energy is part of the theory of relativity that Albert Einstein developed in
1905. This theory correctly predicted that mass and energy are related by the equation

E=mc 2 , where E is energy, m is mass, and c is the speed of light.


The idea that mass and energy are equivalent was proposed by Albert Einstein In 1905 as part of the
theory of relativity with his famous equation E=mc 2 , where E is energy, m is mass, and c is the speed
of light in a vacuum. This equation has two interpretations.

The first is that mass is a form of energy. The interaction of an electron—positron pair illustrates this
idea well — the particles completely destroy each other (annihilation) and the entire mass of the
particles is transformed into two gamma photons.

The second interpretation is that energy has mass. The change in mass ∆ m of an object, or a system, is
related to the change in its energy ∆ E by the equation∆ E=∆ m c 2. A moving ball has kinetic energy,
implying that its mass is greater than its rest mass. The same happens to electrons in particle
accelerators. However, because they can have speeds close to the speed of light, their mass could be a
hundred times greater than their rest mass. Similarly, a decrease in the energy of a system means the
mass of the system must also decrease. For example, the mass of a mug of hot tea decreases as it cools
and loses thermal energy. However, the change in mass is negligibly small (the conversion factor c 2 is
enormous).

For example:

 A sealed torch that radiates 10W of light for 10h (= 36000s) would lose 0.36MJ of energy (=
10W × 36000s). Its mass would therefore decrease by 4.0 ×10 –12 kg (= 0.36MJ/c2), an
insignificant amount compared with the mass of the torch.
 A car of 1000kg mass that speeds up from a standstill to 30ms–1 would gain 450kJ of kinetic
energy so its mass when moving at 30ms–1 would be 5.0 ×10 –12 kg (= 450kJ/c2 ) more than when
it is at rest.
 An unstable nucleus that releases a 5MeV γ photon would lose 8.0 ×10 –13 J of energy. Its mass
would therefore decrease by 8.9 ×10 –30 kg (= 8.0 ×10 –13 J/c2) which is not an insignificant
amount compared with the mass of a nucleus.

As we know physicists commonly use the atomic mass unit, u, for calculations involving nuclei and
subatomic particles. For nuclear calculations, it is useful to know the energy equivalent for 1 u:
2 2
E=1u × c =( 1.660539× 10
−27
) ( 2.997925 ×108 ) =1.492418× 10−10J
1 eV
¿ 1.492418 ×10−10 J × =931.4941 MeV
1.602177 ×10−19 J

Thus, 1 u is equivalent to about 931.5 MeV.

Nuclear reactions can involve conversions between mass and energy. The law of conservation of energy
still applies if the conversions are taken into account. For any closed system, the total of the energy and
the energy equivalent of the mass in the system is constant.

Everyday situations

Consider a person with rest mass 70 kg sitting in a stationary car. Now imagine the car travelling at a
steady speed of 15ms-1. The person has gained kinetic energy, an increase in energy ∆ E .The person will
therefore have increased mass. The change in mass ∆ m can be calculated using the mass—energy
equation∆ E=∆ m c 2.

1 1
m v2 ×70 ×15 2
∆E 2 2
∆ m= 2 = 2 =
c c (3.00 ×10¿¿ 8)2=8.8 × 10−14 ≈10−13 kg ¿
This is a minuscule change in mass and is not noticeable.

Annihilation and creation

When a particle and its corresponding antiparticle meet, they destroy each other in a process known as
annihilation. Positrons are the antiparticles of electrons. When they meet, they annihilate each other,
and their entire mass is transformed into energy in the form of two identical gamma photons. This is not
science fiction. It does happen, and medical physicists have exploited this phenomenon in positron
emission tomography (PET). A PET scantier is used to examine the function of organs, including the
brain. Consider an electron—positron pair annihilating each other.

 change in mass ∆ m=2m e ¿


2 2
 energy released ∆ E=∆ m c =2 m e c
2
 minimum energy of two gamma photons = 2 m e c
2
 minimum energy of each gamma photon = m e c

Therefore, the minimum energy of each photon is 8.2 x 10 -14J or about 0.51 MeV. If the Interacting
particles also have kinetic energy, then the energy of each photon would be even greater.

In a pair production, a single photon vanishes and its energy creates a particle and its corresponding
antiparticle.

Forces in the Nucleus

Aside from hydrogen, all nuclei consist of two or more protons and a number of neutrons. Like charges
repel each other, so what keeps these nuclei from flying apart?

Example:

Can gravitational force bind two protons in a nucleus together?

Given

proton mass m=1.76 × 10−27 kg and proton charge q=1.60 ×10−19 C .

Required

Determine if gravitational force can bind two protons in a nucleus together.

Analysis and Solution


Compare the gravitational and electrostatic forces between two protons in a nucleus.

G m1 m2
The magnitude of the gravitational force is F g= .
r2
k q1 q2
The magnitude of the electrostatic force is F e = .
r2
G m1 m2
Fg r2 Gm1 m2
So, = =
Fe k q1q2 kq 1 q 2
2
r
This ratio shows that the relative strength of the two forces does not depend on the distance between
the protons. In order for the gravitational attraction between the protons to overcome the electrostatic
repulsion, the ratio would have to be greater than 1. Substituting the known values into the ratio of the
forces gives

F g ( 6.67 × 10−11 ) (1.67× 10−27 )2 −37


= 9 −19 2
=8.08× 10
F e (8.99 ×10 )(1.60 ×10 )

Paraphrase

The gravitational attraction is vastly weaker than the electrostatic repulsion, so gravity cannot be the
force that holds a nucleus together.

Since gravity is far too weak, there must be some other force that holds the particles in a nucleus
together. Physicists call this force the strong nuclear force, and think that it is a fundamental force of
nature, like gravity and the electrostatic force. The strong nuclear force has a very short range. Although
it is more powerful than the electrostatic force within a nucleus, the strong nuclear force has a negligible
effect on particles that are more than a few femtometres apart. The strong nuclear force acts on both
neutrons and protons, but does not affect electrons.

Binding energy

As discussed earlier, the strong nuclear force acts between neighbouring nucleons within the nucleus. It
has a very short range ≈ 10 −15 m or 1 fm. The stability of many nuclei provides evidence for the strong
nuclear force.

Protons and neutrons are very tightly bound to each other in a nucleus. To separate them, energy must
be supplied to the nucleus. In order to completely dismantle a nucleus into all of its constituent
nucleons work must be done to separate the nucleons and overcome the strong nuclear force acting
between them. This work is known as the nuclear binding energy. The binding energy of a nucleus is
the energy required to separate all of its protons and neutrons and move them infinitely far apart.

Suppose we could reverse the process and construct a nucleus from a group of individual nucleons. We
would expect there to be energy released as the strong nuclear force pulls them together. This would be
equal to the nuclear binding energy needed to separate them. This implies that energy is needed to
deconstruct a nucleus from nucleons and is given out when we construct a nucleus.

Mass defect and Binding energy

Deuterium is an isotope of hydrogen. A nucleus of deuterium consists of one proton and one neutron.
Now imagine separating these two nucleons. All nucleons are bound together by the strong nuclear
force, so they can only be separated by doing work to overcome that force. External energy has to be
supplied to make this happen. According to Einstein's mass—energy equation, energy and mass are
equivalent, therefore the total mass of the separated nucleons must be greater than the mass the
deuterium nucleus.

Is this really true? We can use a mass spectrometer to determine the mass of particles accurately. In
terms of unified atomic mass units u (1.661 × 10-27kg), a deuterium nucleus has mass 2.013553u, a
proton has mass 1.007276u, and a neutron has mass 1.008665 u. The total mass of the separated proton
and neutron is indeed more than the mass of the deuterium nucleus. The difference is 0.002388u, which
is equivalent to an energy of about 3.5 x10 -13J or 2.2 MeV. In simple terms, this means that a minimum
energy of 2.2 MeV is needed to completely separate the nucleons of a deuterium nucleus. Suppose we
could reverse the process and construct a deuterium nucleus from a proton and a neutron. This time, an
energy of 2.2 MeV would be released — most likely in the form of a photon.

In the example of the deuterium nucleus above, the difference in mass of 0.002388u is known as the
mass defect of the deuterium nucleus.

The mass defect of a nucleus is defined as the difference between the mass of the completely separated
nucleons and the mass of the nucleus.

In other words, the mass of a nucleus is less than the total mass of its individual protons and neutrons.
This difference in mass is called mass defect.

The energy difference of 2.2 MeV for the deuterium nucleus is known as its binding energy. The binding
energy of a nucleus is defined as the minimum energy required to completely separate a nucleus into its
constituent protons and neutrons.

To calculate the binding energy of a nucleus, we can use Einstein's mass—energy equation.

binding energy of nucleus=mass defect of nucleus × c 2


The binding energy is not the same for all nuclei.

A uranium-235 has 92 protons and 143 neutrons, and the external energy required to split this nucleus
into its constituent protons and neutrons to be much greater than 2.2 MeV — there are many more
strong nuclear bonds to be broken.

The following formula uses atomic masses to calculate the mass defect for a nucleus:

∆ m=Z m proton + N mneutron −mnuclide


where Z is the proton number, and N is the neutron number.

And for binding energy

Bindingenergy ( BE ) =∆ m c 2=(Z m proton + N m neutron −mnuclide ) ×c 2


The binding energy per nucleon

It is defined as the total binding energy of a nucleus divided by A, the total number of nucleons. We
calculated above that the binding energy of 21 H is 2.2 MeV, so its binding energy per nucleon is
2.2MeV/2 = 1.1 MeV.

The binding energy per nucleon is the average energy needed to separate a nucleus into its individual
nucleons. A greater binding energy per nucleon means a more tightly bound nucleus in other words a
nucleus is more stable if it has a greater BE per nucleon.

Figure is a graph of BE per nucleon against nucleon number A. The shape of the graph helps us to
understand processes such as natural radioactive decay, fission, and fusion. The last two processes are
covered in greater depth in the next two topics. From the graph you can see that:

The main features of the graph are:


1
 The binding energy per nucleon for hydrogen, 1 H , is zero because there is only one particle in
the nucleus.
 The curve rises sharply for low values of A.
 For nuclei with A < 56, the BE per nucleon increases as A increases.
 For nuclei with A> 56, the BE per nucleon decreases as A increases.
56
 The nucleus of iron-56 ( 26 Fe ) has the greatest BE per nucleon — it is the most stable isotope in
nature. (and 62
28 ¿ as well)
4 12 16
 There are peaks at the position of the nuclei 2He , 6C and 8O , which makes these nuclei
unusually stable compared to their immediate neighbours.
 The curve drops gently from the peak at A = 56 (After A=62) and onwards.
 In a fusion process, two low A number nuclei join together to produce a higher A number
nucleus. The newly formed nucleus has much greater binding energy than the initial nuclei and
therefore energy is released. Fusion is the process by which the Sun and other stars produce
their energy. Thanks to fusion, we have life on Earth.
 In a fission process, a high A number nucleus splits into two lower, 4 number nuclei. Energy is
released because the two nuclei produced have higher binding energy than the parent nucleus.
All fission reactors use this process to produce energy.

 Energy is released in natural radioactive decay. Figure above can be used to show that in cases
of spontaneous decay the total binding energy of the parent nucleus is less than the binding
energy of the daughter nucleus and the alpha particle. The difference is the energy released in
the decay as kinetic energy

 Most nuclei have a binding energy per nucleon between 7 and 9 MeV. The short range of the
force implies that a given nucleon can interact with its immediate neighbours only and not with
all of the nucleons in the nucleus. So for large nuclei (i.e. roughly A > 20) any one nucleon is
surrounded by the same number of immediate neighbours and so the energy needed to remove
that nucleon from the nucleus is the same. Thus, the short range nature of the nuclear force
explains why the binding energy per nucleon is roughly constant above a certain value of A.

Nuclear reactions:

Some nuclei can release energy from nuclear fission or nuclear fusion. Almost all nuclear reactions that
occur naturally result in nuclei that are more stable. This increases the binding energy per nucleon
compared with the original nuclei. The mass difference between the original nuclei and the nuclei of the
products corresponds to the amount of energy released. Figure shows that, to increase the binding
energy per nucleon, lighter elements tend to fuse and heavier elements tend to undergo radioactive
decay or fission.
In all nuclear reactions, total proton number Z and mass number A are conserved, and the reaction often
results in more than one product.

Nuclear Reactions and the Transmutation of Elements


When a nucleus undergoes α or β decay, the daughter nucleus formed is that of a different
element from the parent. The transformation of one element into another, called
transmutation. This also occurs by means of nuclear reactions. A nuclear reaction is said to occur
whenever an incident nucleus, particle, or photon causes a change to occur in a target nucleus.
Ernest Rutherford was the first to report seeing a nuclear reaction. In 1919 he observed that
some of the α particles passing through nitrogen gas were absorbed and protons emitted. He
concluded that nitrogen nuclei had been transformed into oxygen nuclei via the reaction
4
2 He + 147 N → O+ 178¿ 11 H ¿

4 1
where 2He is an α particle, and 1 H is a proton.
In nuclear reaction, both the electric charge and the nucleon number are conserved.
Energy and momentum are also conserved in nuclear reactions. These conservation laws can be
used to determine whether a given reaction can occur or not. For example, if the total mass of
the products is less than the total mass of the initial particles, this decrease in mass (
2
∆ E=∆ m c ¿ is converted to kinetic energy of the outgoing particles. But if the total mass of
the products is greater than the total mass of the initial reactants, the reaction requires energy.
The reaction will then not occur unless the bombarding particle has sufficient kinetic energy.
Consider a nuclear reaction of the general form
a+ X=Y + b
where a is a projectile particle (or small nucleus) that strikes nucleus X , producing nucleus Y
and particle b (typically, proton, neutron, α , γ ). We define the reaction energy, or Q-value, in
terms of the masses involved, as
Q=(M ¿ ¿ a+ M x −M Y −M b ) c 2 ¿
When Q is positive, the total mass decreases and the total kinetic energy increases. Such a
reaction is called an exoergic reaction. When Q is negative, the mass increases and the kinetic
energy decreases, and the reaction is called an endoergic reaction. The terms exothermal and
endothermal, borrowed from chemistry, are also used. In an endoergic reaction the reaction
cannot occur at all unless the initial kinetic energy is at least as great as |Q|.That is, there is a
threshold energy, the minimum kinetic energy to make an endoergic reaction go. KE is greater
after the reaction than before. If Q is negative (Q < 0), the reaction is said to be endothermic or
endoergic.

Nuclear fission
Nuclear fission is a decay process in which an unstable nucleus splits into two fragments of
comparable mass.
If we are able to take a large nucleus and split it into two smaller ones, the binding energy per
nucleon will increase as we move from the right-hand side to the centre of figure of a graph of
BE per nucleon against nucleon number A. This means that energy must be given out in the form
of the kinetic energy of the fission products.
When a nucleus with A > 120 splits into smaller nuclei, they have greater binding energy per
nucleon. This fission reaction gives off energy equal to the difference between the binding
energy of the original nucleus and the total binding energy of the products.
For both nuclear fission and fusion, the energy released, ∆ E , is
∆ E=Eb −E b =(net change ∈the mass defect )×c 2
f i

where Eb is the total binding energy of the original nucleus or nuclei, and Eb is the total binding
i f

energy of the product(s).


Since the binding energies correspond to the mass defects for the nuclei, the energy released
corresponds to the decrease in the total mass defect. This change in the total mass defect
equals the change in the total mass. Thus, the energy released corresponds to the mass that the
reaction transforms into energy:
∆ E=( m f −m i ) ×c 2
where m i is the total mass of the original nucleus or nuclei, and m f is the total mass of the
product(s).
Often, fission results from a free neutron colliding with a large nucleus. The nucleus absorbs the
neutron, forming a highly unstable isotope that breaks up almost instantly.
When a neutron is absorbed by a nucleus of uranium-235, uranium momentarily turns into
uranium-236. It then splits into lighter nuclei plus neutrons. One possibility is:
1
n + 235 236 141 92 1
0 92 U → 92U → 56 Ba + 36 Kr + 3 0 n

The energy released can be calculated as follows:


the total mass of the original nuclei(m i )=m 235
U
92
+m n
1
0

= 235.043 930 u+1.008 665 u


= 236.052 595u

total mass of the products ( mf )=m 141


Ba
+m 92
Kr
3m n
1
56 36 0

¿ 140.914 412 u+91.926 156 u+3 (1.008 665u)


¿ 235.866 563 u
m i−m f =236.052 595u−235.866 563u=0.186 032 u

Now, use mass-energy equivalence to calculate the energy released.

1 u is equivalent to 931.5 MeV, so

∆ E=173.3 MeV
Note that the fission process is an induced process and begins when a neutron collides with a
nucleus of uranium-235.
Fission of a uranium nucleus, triggered by neutron bombardment, releases other neutrons that
can trigger more fissions, suggesting the possibility of a chain reaction.
The chain reaction may be made to proceed slowly and in a controlled manner in a nuclear
reactor or explosively in a bomb. The energy release in a nuclear chain reaction is enormous, far
greater than that in any chemical reaction.

Nuclear fusion
The process of building up nuclei by bringing together individual protons and neutrons, or
building larger nuclei by combining small nuclei is called nuclear fusion.

The joining together (or fusing) of small nuclei to give larger ones releases energy. This is
because the total nuclear binding energy of the fused nuclei is larger than the sum of total
nuclear binding energies of the component nuclei. The difference in the binding energies
between the fusing nuclei and the nucleus produced is emitted as the kinetic energy of the
fusion products.

The only way to make nuclei fuse is to bring them close together, to within a few 10 -15m in, so
that the short-range strong nuclear force can attract them into a larger nucleus. All nuclei have a
positive charge, so they will repel each other. The repulsive electrostatic force between nuclei is
enormous at small separations. At low temperatures, the nuclei cannot get close enough to
trigger fusion. However, at higher temperatures, they move faster and can get close enough to
absorb each other through the strong nuclear force.
The conditions for fusion are just right in the core of our stars like our Sun. The temperature is
close to 1.4 x 107K and the density is 1.5 x 105kg m-3. The enormous density ensures a high
number of fusion reactions per second.
Examples of fusion
There are many different types of fusion reactions that can take place within stars - they all
release energy. Fusion reactions often occur in cycles or sequences. One such cycle is shown
here.
2
 Two protons fuse together to produce a deuterium nucleus ( H ¿ 1¿ , a positron, and a neutrino.
This reaction produces about 2.2 MeV of energy.
1
1 p + 11 p → H + 21¿ +01e + ν ¿
 We can easily explain this energy using the graph of binding energy per nucleon against nucleon
number. The two single protons have zero binding energy and the deuterium nucleus has a
binding energy of1.1 ×2=2.2 MeV . The difference in the binding energies, of 2.2 MeV, is the
energy released in this fusion reaction.
 The deuterium nucleus from the first reaction fuses with a proton. A helium=3 nucleus is formed
and 5.5 MeV of energy is released.
2
1 H + 11 p → 32 He + γ
• The helium-3 from the second reaction combines with another helium-3 nucleus. A helium-4
nucleus is formed, together with two protons and 12.9MeV of energy.
3
2 He + 32 He → 42He +2 11 p
Total

4 11H → 42 He +2 +10β +2 ν + γ

The whole cycle is repeated again with the two protons. This cycle is known as the proton-
proton cycle or the hydrogen-burning cycle, and is one of the main production routes for helium
in stars. The proton-proton cycle occurs around 9 x 10 37 times each second inside the Sun.

Numericals:

1. Find the total energy released in the fusion reactions in the proton–proton cycle.

mass of proton=1.007 825 u , mass of helium=4.002 603 u


Solution:
The net nuclear result of the proton–proton cycle is to fuse four protons to form an alpha
particle. Study the reactions above for the proton–proton cycle to be sure you understand how
four protons become an alpha particle.

the initial mass of the system=4 ( 1.007 825 u )=4.031300 u


the change∈mass of the system=4.031300 u−4.002 603 u=0.028 697 u
MeV
energy released ( E )=0.028 697 u × 931.494 =26.7 MeV
u
This energy is shared among the alpha particle and other particles such as positrons, gamma
rays, and neutrinos.

2. The Sun radiates about 4 × 1026 W and has a mass of 1.99 ×1030 kg. Astronomers estimate that
the Sun can convert only the innermost 10% of its hydrogen into helium. 73% of the mass of the
sun is hydrogen. Estimate how long the Sun can continue to shine at its present intensity.

Given Power=4 × 1026 W m Sun =1.99 ×1030 kg


Hydrogen available for conversion=10 % of total hydrogen
¿ find timethe Sun will take ¿ convert 10 % of its hydrogeninto helium (t)
Solution:
The fusion of four hydrogen atoms produces26.71 MeV . To find the rate at which helium nuclei
are produced, divide the Sun’s power by the energy released during the formation of each
helium nucleus:
Power of sun
Rate of helium production=
energy released per helium atom
26
4 × 10 W 4 × 1026 J / s
¿ = =9.36× 1037 atoms/ s
26.71 MeV /atom MeV 1.60 ×10 J −13
26.71
atom (1 MeV )
Since 4 hydrogen atoms are needed for each helium produced, multiply by 4 to find the rate at
which the Sun converts hydrogen atoms into helium.
Then, convert this rate to mass per second by multiplying it by the mass of a hydrogen atom:
kg
Rate of hydrogen conversion=4 (9.36 ×1037 atoms /s)×(1.67× 10−27 )
atoms
¿ 6.25 ×1011 kg /s
Hydrogen makes up 73% of the mass of the Sun, but only 10% of this hydrogen can be converted
into helium.
The lifespan of the Sun approximately equals the amount of hydrogen that can be converted
divided by the conversion rate.
amount of hydrogenavailable 1.99 ×1030 kg ×73 % × 10 % 17 9
t= = 11
=2.32 ×10 ∨about 7 ×10 years
rate of conversion 6.25× 10 kg / s
The Sun can continue to produce energy at its present rate for about 7 billion years.
235
1. Calculate binding energy of 92U .
235
Given : mass of U = 235.004393, mass of proton=1.007276u and mass of neutron=1.008665u
92
Solution:
The uranium nucleus has 2 protons∧(235−92)=143neutrons
Therefore total mass of neucleons=( 92 ×1.007276 ) + ( 143 ×1.008665 ) =236.908487u
mass defect for Uranium=236.908487 u−235.004393u=1.904094 u

mass defect ∈ kg=1.904094 u × 1.66× 10−27=3.16 × 10−27 kg

B . E .=∆ m× c 2=3.16 ×10−27 ×(3.0 ×108 )2=2.84 × 10−10 J

2.84 ×10−10
BE∈MeV = =1777.9 ≈ 1780 MeV
1.60 ×10−13
Pair production

The opposite process to annihilation is called pair production. In this process, a photon with enough
energy can interact with a large nucleus and be converted directly into a particle – antiparticle pair. The
rest energy of an electron (and therefore a positron) is 0.51 MeV, or 8.2 × 10 −14 J. In order to create this
particle – antiparticle pair, the photon must have enough energy to create both particles, in other
words, (2 × 0.51) = 1.02 MeV, or 16.4 × 10 −14 J. The wavelength of the photon needed to do this can be
calculated by:

hc
E=
λ

hc 6.6 ×10−34 × 3.0× 108 −12


λ= = =1.2× 10 m
E 16.4 ×10 ¿14

You might also like