Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

CHAPTER TWO

2.0 LITERATURE REVIEW

Radioactivity is the spontaneous emission of energy in the form of particles or waves


(electromagnetic radiation), or both (Lange and Carroll, 2008), from the atomic nucleus of
certain elements. It was discovered in 1896 by Henri Becquerel (a French Physicist) who
observed that uranium emitted penetrating rays continuously and without initiation. This
pioneering work was followed by Pierre and Marie Curie shortly afterwards (Lange and Carroll,
2008). They proved that the radioactivity of uranium was an atomic property and not a chemical
one. The pioneering work above led to the discovery that radioactive atoms emit three different
kinds of radiation namely alpha particle, beta particle and gamma rays (gamma rays are similar
to, but more energetic than, the X-rays, an energetic form of electromagnetic radiation
discovered by the German Physicist Wilhelm Conrad Roentgen in 1895) (Lange and Carroll,
2008). Building on the research of Marie Curie and others, it was soon realized that if atoms
emitted such things, they could not be indivisible and unchangeable. Atoms are made up of
smaller particles, and these could be rearranged. In 1900 Ernest Rutherford found that the
radioactivity of the ―emanation‖ (as he called it) from thorium diminished with time. This decay
of radioactivity was a vital clue (Lange and Carroll, 2008). Rutherford, working in Canada with
the Chemist Frederick Soddy, developed a revolutionary hypothesis to explain the process. They
realized that radioactive elements can spontaneously change into other elements, and as they do
so, they emit radiation of one type or another. The spontaneous decay process continues in a
chain of emissions until a stable atom is formed (Lange and Carroll, 2008). They recognized that
the ceaseless emissions pointed to a vast store of energy within atoms. The understanding of how
the atom is built is of great importance in understanding what happens when radioactive atoms
emit radiation. As Rutherford first explained in 1911, each atom is made of a small, massive
nucleus, surrounded by a swarm of light electrons. It is from the nucleus that the radioactivity,
the alpha or beta or gamma rays, shoot out. By 1932 Rutherford's colleagues had found that the
nucleus is built of smaller particles (Shaw et al., 2007), the positively charged protons and the
electrically neutral neutrons. A proton or a neutron each has about the mass of one hydrogen
atom. All atoms of a given element have a given number of protons in their nuclei, called the
atomic number. To balance this charge they have an equal number of electrons swarming around
the nucleus. It is these shells of electrons that give the element its chemical properties (Shaw et
al., 2007). Atoms of a given element can have different numbers of neutrons, and thus different
atomic mass. Soddy named the forms of an element with different atomic masses the isotopes of
the element (Shaw et al., 2007). For example, the lightest element, hydrogen, has the atomic
number 1. Its nucleus normally is made of one proton and no neutrons, and thus its atomic mass
is also 1 (Shaw et al., 2007). Hydrogen on the other hand has isotopes with different atomic
masses. "Heavy" hydrogen, called deuterium, has one proton and one neutron in its nucleus, and
thus its atomic mass is 2. Hydrogen also has a radioactive isotope, tritium. Tritium has one
proton and two neutrons, and thus its atomic mass is 3. The three forms of hydrogen each have
one electron, and thus the same chemical properties (Hutchison and Hutchison, 1997).

Rutherford and Soddy also discovered that every radioactive isotope has a specific half-life. Half
the nuclei in a given quantity of a radioactive isotope will decay in a specific period of time
(Hutchison and Hutchison, 1997). The half-life of uranium-238 for instance is 4.5 billion years,
which means that over that immense period of time half the nuclei in a sample of uranium-238
will decay (in the next 4.5 billion years, half of what is left will decay, leaving one quarter of the
original, and so forth). The isotopes produced by the decay of uranium themselves promptly
decay in a long chain of radiations (Hutchison and Hutchison, 1997). Radium and polonium are
links in this chain. The early work of Marie and Pierre curie led almost immediately to the use of
radioactive materials in medicine. In many circumstances isotopes are more effective and safer
than surgery or chemicals for attacking cancers and certain other diseases. Over the years, many
other uses have been found for radioactivity (Siyal et al., 2022). Until electrical particle
accelerators were invented in the 1930s, scientists used radiation from isotopes to bombard
atoms, uncovering many of the secrets of atomic structure. To this day radioactive isotopes are
still being used as "tracers" to track chemical changes and the processes of life by biologists and
physiologists. Isotopes are crucial even for geology and archeology (Siyal et al., 2022). As soon
as radioactive decay was understood, Pierre Curie realized that it could be used to date materials.
Soon the age of the earth was established by uranium decay at several billion years, far more
than scientists had supposed, and since the 1950s radioactive carbon has been used to determine
the age of plant and animal remains, for example in ancient burials back to 50,000 years ago
(Siyal et al., 2022).
2.1 DECAY MODES AND TYPES OF RADIATION

The four most common modes of radioactive decay are: alpha decay, beta decay, inverse beta
decay (considered as both positron emission and electron capture) (Hameed et al., 2014), and
isomeric transition. Of these decay processes, only alpha decay (fission of a helium-4 nucleus)
changes the atomic mass number (A) of the nucleus, and always decreases it by four. Because of
this, almost any decay will result in a nucleus whose atomic mass number has the same residue
mod 4. This divides the list of nuclides into four classes. All the members of any possible decay
chain must be drawn entirely from one of these classes (Hameed et al., 2014).

Three main decay chains (or families) are observed in nature (Standring et al., 2009). These are
commonly called the thorium series, the radium or uranium series, and the actinium series,
representing three of these four classes, and ending in three different, stable isotopes of lead. The
mass number of every isotope in these chains can be represented as A = 4n, A = 4n + 2, and
A = 4n + 3, respectively. The long-lived starting isotopes of these three isotopes, respectively
thorium-232, uranium-238, and uranium-235, have existed since the formation of the Earth,
ignoring the artificial isotopes and their decays created since the 1940s (Hameed et al., 2014).

2.1.1 ALPHA DECAY

Alpha decay is a fundamental process in nuclear physics, playing a significant role in the natural
decay of certain radioactive elements (Standring et al., 2009). This phenomenon, discovered in
the early 20th century by Ernest Rutherford, forms the basis for understanding radioactive decay
and has profound implications across various scientific disciplines (Standring et al., 2009). At its
core, alpha decay involves the emission of an alpha particle from the nucleus of an unstable
atom. An alpha particle consists of two protons and two neutrons, essentially forming a helium-4
nucleus. This emission process occurs spontaneously in certain radioactive isotopes as they strive
to achieve stability by transforming into nuclei with lower mass and energy (Standring et al.,
2009).

The mechanism of alpha decay can be understood within the framework of nuclear binding
energy. Atomic nuclei are held together by the strong nuclear force, which overcomes the
electromagnetic repulsion between positively charged protons (Vandecasteele, 2004). However,
in some cases, nuclei become energetically unstable due to excess binding energy or unfavorable
proton-neutron ratios. To reach a more stable configuration (Vandecasteele, 2004), these nuclei
undergo radioactive decay, releasing energy in the form of radiation.

Alpha decay typically occurs in heavy, neutron-rich isotopes, where the nucleus is unable to
maintain stability due to excessive mass and energy (Vandecasteele, 2004). As a result, the
nucleus emits an alpha particle, reducing its mass number by four and atomic number by two.
This emission process transforms the original parent nucleus into a different element, known as
the daughter nucleus. For example, the decay of uranium-238 (U-238) produces thorium-234
(Th-234) through the emission of an alpha particle (Villa-Alfageme et al., 2018). The emission
of alpha particles during decay follows specific conservation laws, including conservation of
mass, charge, and momentum. The total mass and charge of the parent nucleus must equal the
combined mass and charge of the daughter nucleus and alpha particle. Additionally, the
momentum of the emitted particles must be conserved, influencing their kinetic energies and
trajectories (Villa-Alfageme et al., 2018).

One characteristic feature of alpha particles is their relatively low penetration power compared to
other forms of radiation, such as beta particles and gamma rays. Due to their large mass and
double positive charge (Villa-Alfageme et al., 2018), alpha particles interact strongly with matter
through electromagnetic forces, losing energy rapidly as they traverse through materials (Villa-
Alfageme et al., 2018). As a result, alpha radiation is easily stopped by a few centimeters of air
or a sheet of paper, making it less hazardous to living organisms unless ingested or inhaled.
Despite their limited penetration, alpha particles can still pose significant health risks if
internalized (Ojovan et al., 2019). When emitted within the body, alpha radiation can cause
severe damage to surrounding tissues, increasing the risk of cancer and other adverse health
effects. Therefore, proper precautions must be taken when handling or working with alpha-
emitting isotopes to minimize exposure and ensure safety (Ojovan et al., 2019).

The rate of alpha decay is governed by the decay constant, which characterizes the probability of
decay for a given radioactive isotope (Ojovan et al., 2019). This decay constant, denoted by λ,
determines the half-life of the isotope, representing the time it takes for half of the radioactive
nuclei to decay. Isotopes with shorter half-lives undergo rapid decay, while those with longer
half-lives decay at a slower rate, contributing to their persistence in the environment (Livingston
and Povinec, 2000).
Alpha decay plays a crucial role in various scientific and technological applications, ranging
from dating geological samples to powering nuclear reactors (Livingston and Povinec, 2000).
One notable application is in radiometric dating, where the decay of alpha-emitting isotopes,
such as uranium and thorium, is used to determine the age of rocks and minerals (Livingston and
Povinec, 2000). By measuring the abundance of parent and daughter isotopes in a sample,
scientists can calculate its age based on the known decay rates. In nuclear power generation,
alpha decay is harnessed to produce energy through the process of nuclear fission (Livingston
and Povinec, 2000). Certain isotopes, such as uranium-235 and plutonium-239, undergo fission
when bombarded with neutrons, releasing a cascade of alpha and beta particles, as well as
gamma rays. This energy release is harnessed to generate electricity in nuclear reactors,
providing a reliable and low-carbon source of power (Livingston and Povinec, 2000).

2.1.1.1 ALPHA PARTICLES

An alpha particle is made up of positively charged helium ion and owing to the charge and
relatively large particle size, they have limited power of penetration, but are highly ionized and
could cause serious internal damage to sensitive parts of the body if radionuclides that emits
alpha particles are ingested (Kaizer et al., 2018). The alpha particle is a stable combination of
two protons and two neutrons, and the only natural source of alpha particles is nuclear decay.
Alpha particles are emitted from the nuclei of many heavy radioactive atomic nuclei during
decay. Alpha particle emission occurs only in elements of high atomic weight. It contains two
neutrons and two protons, and because of this, it has a mass of four units, which makes it an
extremely heavy particle (Kaizer et al., 2018). This indicate that emission of an alpha particle
from a radioisotope results in the formation of another element four mass units lighter and two
atomic numbers lower. An examples of this type of radioactive decay is the transformation of
226-Radium to 222-Radon. Alpha particle has a low penetrating power. For instance, an alpha
particle with an energy as high as 5 MeV has a range (the maximum distance that is necessary to
stop it) in tissue of less than 0.04 millimeter (Kaizer et al., 2018). However, alpha particle is
highly interactive in the vicinity where it is produced because of its large mass and its hefty
charge. When they travel through a material, they lose energy by collision (Kaizer et al., 2018).
2.1.2 BETA DECAY

Beta decay is a fundamental process in nuclear physics, essential for understanding the behavior
of certain radioactive isotopes and their transformation into more stable configurations. This
phenomenon, discovered in the early 20th century, plays a significant role in diverse scientific
disciplines, from particle physics to astrophysics (Jha et al., 2005). At its core, beta decay
involves the transformation of an unstable atomic nucleus through the emission of beta particles.
Beta particles, denoted as β- or β+, are high-energy electrons or positrons ejected from the
nucleus during the decay process (Jha et al., 2005). The emission of these particles alters the
composition of the nucleus, leading to changes in its atomic number and mass number. Beta
decay occurs in isotopes where the neutron-to-proton ratio is unfavorable for stability. In these
cases (Jha et al., 2005), the nucleus undergoes a rearrangement of its constituents to achieve a
more balanced configuration. There are two primary modes of beta decay: beta-minus (β-) decay
and beta-plus (β+) decay (Jha et al., 2005).

In beta-minus decay, a neutron within the nucleus is transformed into a proton, releasing a beta
particle and an antineutrino (IAEA, 2009). This process increases the atomic number of the
nucleus by one while maintaining the same mass number. As a result, the original parent nucleus
is transmuted into a different element, known as the daughter nucleus. For example, the decay of
carbon-14 (C-14) into nitrogen-14 (N-14) involves beta-minus decay (IAEA, 2009). Conversely,
in beta-plus decay, a proton within the nucleus is converted into a neutron, emitting a positron
and a neutrino. This process decreases the atomic number of the nucleus by one while again
preserving the mass number. The parent nucleus transitions into a new element, known as the
daughter nucleus. An example of beta-plus decay is the conversion of potassium-40 (K-40) into
calcium-40 (Ca-40) (Haridasan et al., 2002). The emission of beta particles during decay follows
specific conservation laws, including conservation of energy, momentum, and charge. The total
energy released in the decay process is distributed between the beta particle, antineutrino (in
beta-minus decay), or neutrino (in beta-plus decay) and the daughter nucleus, with the kinetic
energies of the particles varying accordingly (Haridasan et al., 2002).

One notable property of beta particles is their ability to penetrate matter more deeply than alpha
particles but less so than gamma rays (Haridasan et al., 2002). Beta radiation can be stopped by a
few millimeters of aluminum or several centimeters of wood, making it more hazardous to living
organisms if ingested or inhaled. However, proper shielding and safety measures can mitigate
these risks, ensuring the safe handling of beta-emitting isotopes (Haridasan et al., 2002).

The rate of beta decay is governed by the decay constant, which determines the probability of
decay for a given radioactive isotope (Hameed et al., 2014). This decay constant, denoted by λ,
influences the half-life of the isotope, representing the time it takes for half of the radioactive
nuclei to decay (Hameed et al., 2014). Isotopes with shorter half-lives undergo rapid decay,
while those with longer half-lives decay at a slower rate, affecting their persistence in the
environment. Beta decay has profound implications across various scientific and technological
fields (Hameed et al., 2014). In particle physics, the study of beta decay provides insights into
the fundamental interactions governing the behavior of subatomic particles. Experiments
designed to measure the properties of beta particles and neutrinos have contributed to our
understanding of weak nuclear interactions and the Standard Model of particle physics (Hameed
et al., 2014).

In nuclear medicine, beta-emitting isotopes are used for diagnostic imaging and cancer therapy.
Radioactive tracers labeled with beta-emitting isotopes, such as technetium-99m, are employed
in medical imaging techniques like positron emission tomography (PET) and single-photon
emission computed tomography (SPECT). Additionally, beta-emitting radionuclides, such as
iodine-131, are utilized in targeted radiation therapy to treat various forms of cancer (Guerrero et
al., 2020).

2.1.2.1 BETA PARTICLES

Beta particles are electrons with greater penetrating power than alpha particles but owing to a
lesser ability to ionize (Guerrero et al., 2020), they are not as damaging to living cells as are
alpha particles. Beta particles are of two types namely negatrons (electrons) and positrons. Beta
decay is a spontaneous nuclear process that transforms some unstable radioactive atomic nuclei
into others, and the stability of a nucleus is determined by the number of neutrons and protons
contained in the nucleus. A nucleus that has too many neutrons decays by the emission of an
electron for it to be stable (Guerrero et al., 2020). In this process a force known as weak
interaction force changes a neutron into a proton inside the nucleus, thereby increases the charge
or the number of protons by one. Similarly, a nucleus with an excess number of protons decays
by the emission of a positron by changing a proton into a neutron. The total number of neutrons
and protons in the nucleus does not change in the beta decay process (Guerrero et al., 2020).
Some beta emitters occur in nature, mostly among the heavy elements of the uranium, thorium,
and actinium groups, commonly found in association with crystalline rocks (Guerrero et al.,
2020). Thorium for instance is a principal constituent of some minerals, notably thorite and
monazite (a mixed rare-earth and thorium phosphate). The actinide elements are the fourteen
chemical elements that follow actinium in group IIIB of the periodic table, all of which are
radioactive, because their nuclei are so large that they are unstable and release great amounts of
energy when they undergo spontaneous fission (Guerrero et al., 2020). Generally, the heavy
elements of the uranium, thorium, and actinium groups have an excess of neutrons and hence
decay by the emission of electrons (Guerrero et al., 2020).

2.1.2.2 GAMMA RADIATION

Gamma radiation is a form of electromagnetic radiation, first detected as emissions from natural
radioactive substances such as uranium, radium, and thorium (Fisher et al., 2013). Gamma
radiation does not carry any electric charge or mass but it is a penetrating radiation (Fisher et al.,
2013). Its properties are similar to X-rays but only differ in origin i.e. while the sources of
gamma rays are nuclear processes, those of X rays are atomic. There are several different sources
of gamma radiation. After the emission of an alpha or beta particle from a parent nucleus, the
daughter nucleus formed may have more energy than it would have in its normal state. The
nucleus then de-excites by the emission of gamma rays carrying the excitation energy (Fisher et
al., 2013). Gamma radiation is also produced in a nuclear reaction such as the combination of a
neutron with a proton to form a deuteron. When a particle like an electron combines with its
antiparticle, in this case a positron, they give rise to gamma radiation (Fisher et al., 2013).

Gamma radiation undergoes many diverse interactions with matter at different energy ranges.
Low-energy gamma radiation may be totally absorbed by an atomic electron that is then emitted.
The ejected electron is known as a photoelectron (Eigl et al., 2017), and the process is known as
the photoelectric effect. Gamma radiation can also interact with an atomic electron, sharing its
energy and giving rise to the compton effect, in which the original gamma radiation is scattered
away with reduced energy and the electron is ejected (Eigl et al., 2017). This electron is known
as a compton electron (Eigl et al., 2017). Gamma radiation of sufficiently high energy can also
interact with the electric field of the positively charged nucleus producing an electron and a
positron. This phenomenon is known as pair production (Eigl et al., 2017). When a beam of
gamma radiation passes through matter, its intensity after emergence has diminished, principally
as a result of the above three processes. Very high energy gamma radiation can also cause
nuclear disintegration and can eject a nuclear particle such as a neutron or a proton (Choudri and
Baawain, 2016). Various types of mesons can also be produced by gamma radiation of extremely
high energy in its interaction with atomic nuclei. Gamma radiation is emitted as photons, or
discrete quanta of energy (Choudri and Baawain, 2016).

Fig.1 Penetrative power of alpha radiation (helium nuclei), beta radiation (high-energy
electrons), and gamma rays (photons moving in waves).
2.1.2.3 X-RAY RADIATION

X-rays arises from the electron cloud surrounding the nucleus. They were discovered by
Roentgen in 1895. X-rays are produced in X-ray tube by fast moving electron which is suddenly
stopped by target (Syed et al., 2001).

2.2 UNITS OF RADIATION AND RADIOACTIVITY

Many of the units for measuring radiation and radioactivity are named after the pioneering
scientists of the field—Wilhelm Roentgen (1845–1923) (Choudri and Baawain, 2016), Henri
Becquerel (1852–1908), Marie Curie (1867–1934) and her husband Pierre Curie (1859–1906),
and Ernest Rutherford (1871–1937) (Choudri and Baawain, 2016). The original unit for
measuring the amount of radioactivity was the curie (Ci)–first defined to correspond to
radioactive decay of one gram of radium-226 but more recently defined as:

1 curie = 3.7 × 1010 radioactive decays per second (exactly)

The International System of Units (SI) has replaced the curie with the becquerel (Bq), where:

1 becquerel = 1 radioactive decay per second = 2.703 × 10−11 Ci

Ionizing radiation (radiation with enough energy to ionize [or remove electrons from] other
atoms) is measured using electron-volts, joules and ergs (Calmet, 1989). The electron-volt (eV)
is the energy gained by an electron when it moves from rest through a potential difference of one
volt (e.g. the energy an electron gains as it moves from a negative plate to a positive plate with a
1-V higher potential) (Calmet, 1989). Electron-volts are a useful unit for expressing very small
amounts of energy. One joule (J) is equal to 6.242×1018 electron-volts and is equivalent to the
amount of energy used by a one-watt light bulb lit for one second. The erg is a unit of energy
equal to 6.242×1011 electron-volts or 1×10–7 J (Calmet, 1989).

2.3 HALF-LIVES
All unstable atoms will undergo radioactive decay at some point, but this decay is a random
event; it is impossible to predict at what point in time any given atom will decay. However, for a
very large number of atoms, the number of nuclei that will decay in a given period of time is
predictable (Buesseler et al., 2017). The proportion of atoms decaying in a given period of time
remains constant, (i.e. the number of atoms of a given radioisotope remaining within a sample
reduces exponentially over time). This radioactive decay is expressed in terms of half-life
(Buesseler et al., 2017).

The half-life of a radioisotope is defined as the time taken for half of the radioactive nuclei
within a sample of that isotope to decay or the time taken for the activity (the number of decays
per unit of time) to halve (Buesseler et al., 2017). Each radioisotope has a specific half-life,
which may be anywhere from microseconds to hundreds of years, or even longer. In fact, the
half-life of some radioisotopes is so long that they have remained in their current state since
before the Earth was formed, and some isotopes have half-lives that are longer than the age of
the universe. For example, bismuth-209 has recently been found to have a half-life of 1.9×1019
years, whereas the universe is estimated to have an age of only 1.38×1010 years. Knowing the
decay rate of radioisotopes has a number of practical applications (Buesseler et al., 2012). For
example, decay rates can allow us to determine how long an environment, plant, or animal
contaminated by radioactive waste will remain hazardous and can allow us to determine the age
of various materials including archaeological artifacts, sediment, etc (Buesseler et al., 2012).

2.4 SOURCES OF RADIOACTIVITY

Radioactive substances occur naturally across the whole biosphere, and life has evolved in this
radioactive environment (Buesseler et al., 2012). The natural background levels provide a
reference for acceptable levels and are important to understand before we attempt to measure
anthropogenic increases. Radioactive elements may be found in differing concentrations around
the world as a result of natural and anthropogenic processes. Till date, around 3000 natural and
artificial radioisotopes have been identified (Buesseler et al., 2012).
2.4.1 NATURAL RADIOACTIVITY AND RADIATION

Natural radiation comprises cosmic radiation and the radiation arising from the decay of
naturally occurring radionuclides (Belahbib et al., 2021). The natural radionuclides include the
primordial radioactive elements in the earth's crust, their radioactive decay products, and
radionuclides produced by cosmic-radiation interactions. Primordial radionuclides have half-
lives comparable with the age of the earth (Belahbib et al., 2021). Cosmogenic radionuclides are
produced continuously by bombardment of stable nuclides by cosmic rays, primarily in the
atmosphere. Humans are exposed to natural radiation from external sources, which include
radionuclides in the earth and cosmic radiation, and by internal radiation from radionuclides
incorporated into the body. The main routes of radionuclide intake are ingestion of food and
water and inhalation (Belahbib et al., 2021). A particular category of exposure to internal
radiation, in which the bronchial epithelium is irradiated by alpha particles from the short-lived
progeny of radon, constitutes a major fraction of the exposure from natural sources. In most
places on the earth, natural radiation from external sources varies within about a factor of 4; but
in some localities, the variation is greater because of abnormally high or low soil concentrations
of radioactive minerals (Ojovan et al., 2019). Cosmic radiation alone varies by about a factor of
2 over the range of elevation that encompasses most of the world's population (0-2,000 m) and to
a much smaller degree with latitude because of the variation in the earth's magnetic field (Ojovan
et al., 2019).

Radioisotopes of naturally occurring elements comprise Naturally Occurring Radioactive


Materials (NORM) and are ubiquitous in the environment (Ojovan et al., 2019), occurring in
soil, sand, clay, rocks, air, water, and the tissues of plants and animals. These radioisotopes
undergo radioactive decay that results in one or more types of radiation. Cosmic rays from the
sun and outer space are referred to as ionizing radiation and constantly bombard the Earth. Most
naturally occurring radioactive substances (predominantly radium and radon) are the result of
uranium and thorium decay (Batlle et al., 2018). They may be mobilized, redistributed and
concentrated by human activities such as fossil fuel mining and burning and fertilizer mining.
When NORM are concentrated, or the potential for exposure has been enhanced, due to human
activities they are termed Technologically Enhanced Naturally Occurring Radioactive Materials
(TENORM) (Ojovan et al., 2019). Naturally occurring radioisotopes have been used in the
environmental sciences for over 150 years and enable the study of processes from a cellular level
to broad oceanic scales. They can be applied both in field and laboratory studies. The evolving
field of radioecology has had a strong focus on marine research since the 1970s, and new
applications are expanding the research scope (Cresswell et al., 2020). Studies using naturally
occurring radioisotopes are useful for understanding the chronological formation of the Earth,
sedimentology, contaminant behavior, and nutrient transport through food chains, global element
cycles, and defining natural and anthropogenic sources of nutrients, industry compliance,
ecotoxicology, and remediation success, among others (Batlle et al., 2018).

2.4.1.1 COSMIC RADIATION

Cosmic radiation, also known as cosmic rays, refers to high-energy particles that originate from
outer space and continuously bombard the Earth (Batlle et al., 2018). These particles, primarily
protons but also including helium nuclei and heavier atomic nuclei, travel at nearly the speed of
light and can have significant impacts on both the environment and human health. Understanding
cosmic radiation involves exploring its sources, the mechanisms of its journey to Earth, its
interactions with the Earth's atmosphere (Batlle et al., 2018), and its effects on living organisms
and technology. Cosmic radiation can be categorized into two main types based on their origins:
galactic cosmic rays (GCRs) and solar cosmic rays (Batlle et al., 2018).

Galactic Cosmic Rays GCRs are high-energy particles that originate outside the solar system,
primarily from supernova explosions, which are powerful enough to accelerate particles to near-
light speeds. These particles come from remnants of massive stars that exploded, spreading
elements throughout the galaxy. They are a continuous source of radiation, uniformly distributed
in the galaxy (Batlle, 2011).

Solar cosmic rays, also known as solar energetic particles (SEPs), are emitted by the sun,
especially during solar flares and coronal mass ejections (CMEs) (Aruga and Wakamatsu, 2018).
These events can accelerate particles to high energies (Batlle, 2011), which then travel through
space and occasionally reach Earth. Solar cosmic rays are more sporadic and depend on the solar
activity cycle, which peaks approximately every 11 years (Batlle, 2011).

When cosmic rays enter the Earth's atmosphere, they collide with atomic nuclei in the air,
creating a cascade of secondary particles in a process known as an air shower (Aruga and
Wakamatsu, 2018). These secondary particles include protons, neutrons, muons, electrons, and
various types of mesons. Most of the primary cosmic rays and their secondary particles are
absorbed by the atmosphere, which acts as a protective shield, but some manage to reach the
Earth's surface (Aruga and Wakamatsu, 2018).

2.4.1.2 TERRESTRIAL RADIATION

Terrestrial radiation refers to the natural radiation emitted by radioactive materials present in the
Earth's crust (Aruga and Wakamatsu, 2018). This type of radiation contributes significantly to
the natural background radiation that all living organisms are exposed to daily (Aruga and
Wakamatsu, 2018). Understanding terrestrial radiation involves examining its sources,
distribution, effects on human health, and ways to mitigate its impact. The primary sources of
terrestrial radiation are naturally occurring radioactive materials (NORM) found in the soil,
rocks, water, and even in the atmosphere. The most significant contributors are isotopes of
uranium, thorium, and potassium (Jadiyappa, 2018).

Uranium-238 is a heavy metal found in various minerals and rocks, particularly granite and
phosphate rocks. It decays through a series of radioactive progeny, eventually leading to stable
lead-206. Each decay step emits alpha, beta, or gamma radiation, contributing to the overall
radiation dose (Jadiyappa, 2018). Thorium-232, like uranium, is a naturally occurring radioactive
element found in many rocks and soils (Jadiyappa, 2018). It also decays through a series of
radioactive progeny, producing radiation at each step until it becomes stable lead-208.
Potassium-40 is a naturally occurring isotope of potassium, present in significant quantities in the
Earth's crust (Jadiyappa, 2018). It decays by both beta emission and electron capture,
contributing to background radiation. Potassium-40 is also found in many foods, and thus inside
the human body. Radon-222 is a radioactive gas produced by the decay of uranium-238. It can
seep out of the ground and accumulate in buildings, especially in confined areas like basements.
Radon is a significant health concern because its decay products can be inhaled, posing a risk for
lung cancer (Singh, 2011).

2.4.2 ARTIFICIAL OR MAN-MADE RADIATION

In addition to natural background radiation, human beings are exposed to man-made radiation
obtained from nuclear installations (Syed et al., 2001), nuclear explosions, and nuclear fuel
cycle, radioactive waste releases from nuclear reactor operations, and accidents and other
industrial, medical, and agricultural uses of radioisotopes. The most significant sources of
exposure, which gives the largest contribution to the public is from medical diagnostic X-rays,
nuclear medicine, and nuclear therapy. This is also generated from consumer products such as
combustible fluids (gas and coal), TV, luminous watches and dials, and electron tubes. The
public are exposed to the radiation from the nuclear fuel cycle, which includes the entire
sequence from mining and milling of uranium, the actual production of power at a nuclear power
plant, and residual fallout from nuclear weapon testing and accident. The public are not exposed
to all the sources of radiation, for example, patients who are treated with the medical irradiation
or the workers of nuclear industry may receive higher radiation exposure than the public (Syed et
al., 2001).

2.5 RADIONUCLIDE

A radionuclide (radioactive nuclide, radioisotope or radioactive isotope) is a nuclide that has


excess numbers of either neutrons or protons, giving it excess nuclear energy, and making it
unstable (Spinks and Woods, 1990). This excess energy can be used in one of three ways:
emitted from the nucleus as gamma radiation; transferred to one of its electrons to release it as a
conversion electron; or used to create and emit a new particle (alpha particle or beta particle)
from the nucleus. During those processes, the radionuclide is said to undergo radioactive decay
(Spinks and Woods, 1990). These emissions are considered ionizing radiation because they are
energetic enough to liberate an electron from another atom. The radioactive decay can produce a
stable nuclide or will sometimes produce a new unstable radionuclide which may undergo
further decay. Radioactive decay is a random process at the level of single atoms: it is impossible
to predict when one particular atom will decay (Petrucci et al., 2002). However, for a collection
of atoms of a single nuclide the decay rate, and thus the half-life (t1/2) for that collection, can be
calculated from their measured decay constants. The range of the half-lives of radioactive atoms
has no known limits and spans a time range of over 55 orders of magnitude. Radionuclides occur
naturally or are artificially produced in nuclear reactors, cyclotrons, particle accelerators or
radionuclide generators (Petrucci et al., 2002). There are about 730 radionuclides with half-lives
longer than 60 minutes. Thirty-two of those are primordial radionuclides that were created before
the Earth was formed. At least another 60 radionuclides are detectable in nature, either as
daughters of primordial radionuclides or as radionuclides produced through natural production
on Earth by cosmic radiation (Petrucci et al., 2002). More than 2400 radionuclides have half-
lives less than 60 minutes. Most of those are only produced artificially, and have very short half-
lives. For comparison, there are about 251 stable nuclides. All chemical elements can exist as
radionuclides (Best et al., 2013). Even the lightest element, hydrogen, has a well-known
radionuclide, tritium. Elements heavier than lead, and the elements technetium and promethium,
exist only as radionuclides. Unplanned exposure to radionuclides generally has a harmful effect
on living organisms including humans, although low levels of exposure occur naturally without
harm. The degree of harm will depend on the nature and extent of the radiation produced, the
amount and nature of exposure (close contact, inhalation or ingestion), and the biochemical
properties of the element; with increased risk of cancer the most usual consequence. However,
radionuclides with suitable properties are used in nuclear medicine for both diagnosis and
treatment (Best et al., 2013). An imaging tracer made with radionuclides is called a radioactive
tracer. A pharmaceutical drug made with radionuclides is called a radiopharmaceutical (Best et
al., 2013).

On Earth, naturally occurring radionuclides fall into three categories: primordial radionuclides,
secondary radionuclides, and cosmogenic radionuclides (Loveland et al., 2006). Radionuclides
are produced in stellar nucleo-synthesis and supernova explosions along with stable nuclides.
Most decay quickly but can still be observed astronomically and can play a part in understanding
astronomic processes. Primordial radionuclides, such as uranium and thorium, exist in the
present time because their half-lives are so long (>100 million years) that they have not yet
completely decayed (Severijns et al., 2006). Some radionuclides have half-lives so long (many
times the age of the universe) that decay has only recently been detected, and for most practical
purposes they can be considered stable, most notably bismuth-209: detection of this decay meant
that bismuth was no longer considered stable. It is possible decay may be observed in other
nuclides, adding to this list of primordial radionuclides (Severijns et al., 2006).

Secondary radionuclides are radiogenic isotopes derived from the decay of primordial
radionuclides. They have shorter half-lives than primordial radionuclides. They arise in the decay
chain of the primordial isotopes thorium-232, uranium-238, and uranium235. Examples include
the natural isotopes of polonium and radium. Cosmogenic isotopes, such as carbon14, are present
because they are continually being formed in the atmosphere due to cosmic rays (Severijns et al.,
2006). Many of these radionuclides exist only in trace amounts in nature, including all
cosmogenic nuclides. Secondary radionuclides will occur in proportion to their half-lives, so
short-lived ones will be very rare. For example, polonium can be found in uranium ores at about
0.1 mg per metric ton (1 part in 1010) (Severijns et al., 2006). Further radionuclides may occur in
nature in virtually undetectable amounts as a result of rare events such as spontaneous fission or
uncommon cosmic ray interactions (Severijns et al., 2006).

Radionuclides are produced as an unavoidable result of nuclear fission and thermonuclear


explosions (Luig et al., 2011). The process of nuclear fission creates a wide range of fission
products, most of which are radionuclides. Further radionuclides can be created from irradiation
of the nuclear fuel (creating a range of actinides) and of the surrounding structures, yielding
activation products (Luig et al., 2011). The complex mixture of radionuclides with different
chemistries and radioactivity makes handling nuclear waste and dealing with nuclear fallout
particularly problematic. Synthetic radionuclides are deliberately synthesized using nuclear
reactors, particle accelerators or radionuclide generators (Luig et al., 2011). As well as being
extracted from nuclear waste, radioisotopes can be produced deliberately with nuclear reactors,
exploiting the high flux of neutrons present. These neutrons activate elements placed within the
reactor. A typical product from a nuclear reactor is iridium-192. The elements that have a large
propensity to take up the neutrons in the reactor are said to have a high neutron cross-section.
Particle accelerators such as cyclotrons accelerate particles to bombard a target to produce
radionuclides (Wahl, 2007). Cyclotrons accelerate protons at a target to produce positron-
emitting radionuclides, e.g. fluorine-18. Radionuclide generators contain a parent radionuclide
that decays to produce a radioactive daughter. The parent is usually produced in a nuclear
reactor. A typical example is the technetium-99m generator used in nuclear medicine. The parent
produced in the reactor is molybdenum-99 (Wahl, 2007).

2.5.1 URANIUM-238

The uranium decay series, also known as the uranium-radium series, is a sequence of radioactive
decay processes that uranium-238 undergoes until it reaches a stable isotope of lead-206. This
decay chain is significant both scientifically and historically, as it includes several important
radionuclides and radiation types (Wahl, 2007). Understanding the uranium decay series is
crucial for various applications, including radiometric dating, nuclear energy, and radiation
protection (Mohammed et al., 2000).

Uranium-238 (U-238) is the most abundant isotope of uranium, constituting over 99% of natural
uranium (Mohammed et al., 2000). It has a half-life of about 4.5 billion years, making it one of
the longest-lived radionuclides. U-238 is not fissile but can capture a neutron to become
plutonium-239, which is fissile (Mohammed et al., 2000). The decay of U-238 starts a series of
transformations involving the emission of alpha and beta particles, as well as gamma radiation.
The uranium-238 decay series consists of 14 steps, starting from U-238 and ending in stable
lead-206 (Pb-206) (Mohammed et al., 2000). Each step involves the transformation of one
radionuclide into another through radioactive decay. The key steps and intermediates in the
decay series are as follows:

1. Uranium-238 (U-238)

- Decays by alpha emission.

- Half-life: 4.5 billion years.

- Transforms into thorium-234 (Mohammed et al., 2000).

2. Thorium-234 (Th-234)

- Decays by beta emission.

- Half-life: 24.1 days.

- Transforms into protactinium-234 (Mohammed et al., 2000).


3. Protactinium-234 (Pa-234)

- Decays by beta emission.

- Half-life: 1.17 minutes.

- Transforms into uranium-234 (Mohammed et al., 2000).

4. Uranium-234 (U-234)

- Decays by alpha emission.

- Half-life: 245,500 years.

- Transforms into thorium-230 (Mohammed et al., 2000).

5. Thorium-230 (Th-230)

- Decays by alpha emission.

- Half-life: 75,380 years.

- Transforms into radium-226 (Mohammed et al., 2000).

6. Radium-226 (Ra-226)

- Decays by alpha emission.

- Half-life: 1,600 years.

- Transforms into radon-222 (Mohammed et al., 2000).


7. Radon-222 (Rn-222)

- Decays by alpha emission.

- Half-life: 3.8 days.

- Transforms into polonium-218 (Mohammed et al., 2000).

8. Polonium-218 (Po-218)

- Decays by alpha emission.

- Half-life: 3.1 minutes.

- Transforms into lead-214 (Mohammed et al., 2000).

9. Lead-214 (Pb-214)

- Decays by beta emission.

- Half-life: 26.8 minutes.

- Transforms into bismuth-214 (Mohammed et al., 2000).

10. Bismuth-214 (Bi-214)

- Decays by beta emission.

- Half-life: 19.7 minutes.

- Transforms into polonium-214 (Mohammed et al., 2000).


11. Polonium-214 (Po-214)

- Decays by alpha emission.

- Half-life: 164 microseconds.

- Transforms into lead-210 (Mohammed et al., 2000).

12. Lead-210 (Pb-210)

- Decays by beta emission.

- Half-life: 22.3 years.

- Transforms into bismuth-210 (Mohammed et al., 2000).

13. Bismuth-210 (Bi-210)

- Decays by beta emission.

- Half-life: 5 days.

- Transforms into polonium-210 (Mohammed et al., 2000).

14. Polonium-210 (Po-210)

- Decays by alpha emission.

- Half-life: 138.4 days.

- Transforms into stable lead-206 (Mohammed et al., 2000).


15. Lead-206 (Pb-206)

- Stable isotope.

- End product of the uranium-238 decay series (Mohammed et al., 2000).

2.5.1.1 APPLICATIONS AND SIGNIFICANCE OF URANIUM-238

1. RADIOMETRIC DATING

One of the primary applications of the uranium decay series is in radiometric dating,
particularly uranium-lead dating (Israelsson et al., 1973). By measuring the ratios of
uranium-238 to lead-206, geologists can determine the age of rocks and minerals. This
method has been instrumental in dating the Earth’s crust and understanding geological time
scales (Israelsson et al., 1973).

2. NUCLEAR ENERGY

Understanding the decay series is also crucial in the context of nuclear energy. Radium-226,
a decay product (Israelsson et al., 1973), was historically significant as a source of radon gas
in early nuclear research and medical applications. Additionally, knowing the decay products
and their properties helps in managing radioactive waste and designing safe nuclear reactors
(Israelsson et al., 1973).

2.5.2 CARBON-14 (C-14)

Carbon-14 (C-14), also known as radiocarbon, is a radioactive isotope of carbon with a half-life
of approximately 5,730 years. This isotope is widely known for its application in radiocarbon
dating, a method used to determine the age of ancient biological materials (Shahbazi-Gahrouei,
2003). Carbon-14 is a naturally occurring isotope that plays a significant role in various scientific
fields including archaeology, geology, and environmental science. Carbon-14 is formed in the
upper atmosphere through the interaction of cosmic rays with nitrogen-14 (N-14). When high-
energy cosmic rays collide with atmospheric nitrogen, a neutron is produced. This neutron
subsequently collides with a nitrogen-14 atom, resulting in the formation of carbon-14 and a
proton (Shahbazi-Gahrouei, 2003). Carbon-14 then combines with oxygen to form carbon
dioxide (CO₂), which is absorbed by plants during photosynthesis. As animals consume plants,
they too incorporate C-14 into their bodies, ensuring that all living organisms maintain a constant
level of this isotope. Once the organism dies, it stops absorbing C-14, and the isotope begins to
decay at a known rate (Shahbazi-Gahrouei, 2003). Carbon-14 decays through beta emission,
transforming into nitrogen-14 with the release of a beta particle (an electron) and an antineutrino.
The decay process is characterized by a half-life of 5,730 years, which is the time it takes for half
of the C-14 in a sample to decay. This relatively long half-life makes C-14 ideal for dating
materials that are up to about 50,000 years old (Kwan-Hoong, 2003).

Radiocarbon dating, developed by Willard Libby in the late 1940s, revolutionized the field of
archaeology and our understanding of historical timelines. The technique measures the amount
of C-14 remaining in a sample to estimate its age. By comparing the ratio of C-14 to stable
carbon isotopes (C-12 and C-13) in a sample, scientists can determine how long it has been since
the organism died (Kwan-Hoong, 2003). Radiocarbon dating has been instrumental in dating
ancient artifacts, such as the Dead Sea Scrolls, the Shroud of Turin, and numerous prehistoric
fossils, providing insight into historical and pre-historical chronologies. In the biomedical field,
C-14 is used as a tracer in metabolic studies. By incorporating C-14 into molecules, researchers
can track the biochemical pathways and understand metabolic processes in living organisms.
This application is crucial in pharmacology for understanding drug metabolism and in
biochemistry for studying cellular processes (Kwan-Hoong, 2003).

2.5.3 AMERICIUM-241

Americium-241 (Am-241) is a synthetic radioisotope that has found significant applications in


various fields due to its unique properties. Discovered in 1944 by Glenn T. Seaborg and his team
during the Manhattan Project (Chuvilin et al., 2012). Am-241 is a byproduct of plutonium-241
decay and is one of the most commonly used radioisotopes in both industrial and scientific
applications. Americium-241 has a half-life of approximately 432.2 years, decaying primarily by
alpha emission to neptunium-237 (Chuvilin et al., 2012). It also emits gamma rays, which make
it useful for a variety of applications. Am-241 is produced in nuclear reactors through the
neutron bombardment of plutonium-239 (Pu-239). When Pu-239 captures a neutron, it becomes
Pu-240, which can further capture another neutron to become Pu-241. Pu-241 then decays via
beta emission to form Am-241. The long half-life and the type of radiation emitted make Am-
241 a stable and reliable source of radiation, which is essential for its various applications
(Chuvilin et al., 2012).

One of the most well-known uses of Am-241 is in household smoke detectors. The isotope is
used in ionization smoke detectors, which rely on the ionization of air molecules. The alpha
particles emitted by Am-241 ionize the air within a detection chamber, creating a small but
steady electric current (Dawson et al., 1993). When smoke enters the chamber, it disrupts this
current, triggering the alarm. This application leverages the alpha radiation's ability to ionize air
molecules without posing significant health risks due to its short range in air (Dawson et al.,
1993).

Am-241 is used in industrial gauging devices, particularly in measuring the thickness of


materials and the density of certain substances. These devices utilize the gamma radiation
emitted by Am-241 (Dawson et al., 1993). For instance, in the paper and metal industries,
gauging devices help ensure uniform thickness of products. The gamma rays pass through the
material, and the amount of radiation that emerges on the other side is measured. This reading is
inversely proportional to the thickness of the material, allowing for precise control during
manufacturing (Dawson et al., 1993). Another important application of Am-241 is in
combination with beryllium to create neutron sources. When alpha particles emitted by Am-241
strike beryllium atoms, neutrons are released. These neutron sources are used in various
scientific and industrial applications, including neutron radiography, well logging in the oil and
gas industry, and as a source for initiating nuclear reactions in research reactors. Although not as
commonly used in medical treatments due to its alpha radiation, Am-241 still finds some
applications in medical and scientific research. Its properties make it useful in certain types of
radiography and as a calibration source for gamma spectroscopy equipment. Additionally, the
study of americium compounds and their behavior contributes to a better understanding of
actinide chemistry and nuclear waste management (Sandhya Rani et al., 2014).

The use of Am-241, like any radioactive material, comes with safety and environmental
concerns. Alpha particles, while not deeply penetrating (Sandhya Rani et al., 2014), can cause
significant biological damage if the material is ingested or inhaled. Therefore, handling Am-241
requires stringent safety protocols to prevent contamination and exposure. In smoke detectors,
the amount of Am-241 is minimal and encased in a way that poses little risk to users. Disposal of
Am-241-containing devices also needs careful consideration. Industrial gauges and neutron
sources contain larger amounts of Am-241 and must be handled as radioactive waste (Sandhya
Rani et al., 2014). Regulatory bodies such as the Environmental Protection Agency (EPA) and
the Nuclear Regulatory Commission (NRC) provide guidelines for the safe disposal and
management of such radioactive materials (Sandhya Rani et al., 2014).

2.6 APPLICATIONS OF RADIOISOTOPES

Radioisotopes can be either natural or artificial (Singh, 2011). The natural ones are those that
occur naturally in the environment and are created through various natural processes, such as the
decay of primordial isotopes, cosmic ray interactions and radioactive decay chains (Singh, 2011).
Artificial radioisotopes, on the other hand, are those created in laboratories or nuclear reactors
through nuclear reactions. They do not occur naturally in the environment (Singh, 2011). After
more than 100 years since the discovery of radioactivity, radioisotopes are successfully used in
many aspects of human life, such as energy, medicine, agriculture, archaeology, or various
industries (Singh, 2011).

2.6.1 RADIOISOTOPES IN MEDICINE

The use of radioisotopes in medicine has revolutionized the diagnosis and treatment of many
diseases and conditions (Siyal et al., 2022), saving the lives of millions of patients. Over 90% of
the procedures involving nuclear medicine are for diagnosis, where special radionuclides with
short half-life are used to complement the traditional imaging methods based on X-rays. The
patients are administered a small amount of these radionuclides which emit gamma radiation that
is detected and used to create images of the body's organs and tissues. Radioactivity can be used
for treatment through a technique called radiation therapy, or radiotherapy. This technique
involves the use of radionuclides that emit high-energy radiation, such as gamma rays, X-rays, or
charged particles to cancer cells, in order to kill them and shrink tumours. The radiation damages
the DNA of the cancer cells, which leads to their death (Siyal et al., 2022). This can be done
either externally, when the radiation source is placed outside the body and only a limited area of
the body is irradiated, or internally, where radioactive nuclides is administered to the patient and
deliver a toxic level of radiation to the targeted sites (Siyal et al., 2022). Even though
radiotherapy can have some side effects, such as fatigue, nausea, or skin irritation, it is
considered very well tolerated compared to more traditional treatments, due to their minimum
damage to healthy tissue and is considered a powerful tool in the fight against cancer (Siyal et
al., 2022).

2.6.2 RADIOISOTOPES IN AGRICULTURE

Radioactivity has several important applications in agriculture, improving the quality and
productivity of agricultural products (Siyal et al., 2022). Radioisotopes are very useful in the
effective use of the fertilizers, which may result in lower costs and minimal damage to the
environment. The working principle is quite simple, the nutrients taken by the plant are labelled
with a radioactive tracer such as Nitrogen – 15 or phosphorus – 32 and the movement of these
species through the plant and the soil can be monitored (Siyal et al., 2022). Radioisotopes were
also used to induce mutations, by exposing seeds to radiation, thus leading to new and desirable
treats, such as disease resistance, earlier ripening, and ultimately, higher yield. Radioactive
nuclei are also used in insect pest management (Farkas and Mohácsi-Farkas, 2011). Agricultural
productivity is at risk due to insect pests, which not only cause a reduction in crop yields, but
also spread diseases to cultivated crops. To monitor the persistence of pesticide residues in food
items, soil, groundwater, and the environment, radiolabel pesticides have been employed. These
studies have been beneficial in identifying and reducing the adverse effects of insecticides and
pesticides. Another great use of radioactivity was recorded in food processing and preservation.
Between 25 and 30 percent of the world's food production is lost due to spoilage caused by
microbes and pests, and these losses are even higher in developing countries (Farkas and
Mohácsi-Farkas, 2011). However, this loss of food can be prevented by implementing effective
food preservation methods. Radiation is one method that can be used to eliminate microbes in
food and control insect and parasite infestation in harvested food, which prevents different types
of wastage and spoilage (Farkas and Mohácsi-Farkas, 2011). The agents responsible for spoilage,
such as microbes and insects, are eliminated by irradiation from packaged food, and the
packaging materials are impermeable to bacteria and insects, preventing recontamination. Also,
it can replace or significantly reduce the use of hazardous food additives and fumigants, which
are dangerous for both consumers and food processing workers (Farkas and Mohácsi-Farkas,
2011).

2.6.3 RADIOISOTOPES IN ARCHAEOLOGY

Radioactivity has various applications in archaeology, primarily in the field of radiocarbon


dating (Taylor, 1995). Radiocarbon dating is a technique used to determine the age of organic
materials based on the decay of the radioactive isotope carbon-14. When plants and animals are
alive, they absorb carbon14 from the atmosphere. Once they die, the carbon-14 decays at a
known rate, and by measuring the remaining amount of carbon14 in a sample, scientists can
calculate the time elapsed since the organism died (Taylor, 1995).

2.6.4 RADIOISOTOPES IN INDUSTRY

The main use of radioisotopes in industry is in smoke detecting. Smoke detectors use a small
quantity of radioactive material, such as Americium-241, which emits alpha particles. These
particles ionize the air, creating a small electrical current. When smoke is the air, it disrupts the
flow of ions and reduces the current (Taylor, 1995). Radioactive isotopes can be used as
radiotracers, tracking and monitoring industrial processes. They can be added to a material or a
process stream, and their movement and behaviour can be detected and measured (Pant, 2022).
The main industries that radiotracers are used in include:

 Oil and gas industry: flow rates of oil and gas in pipelines, leaks and blockages or the
performance of oil wells can be monitored with radiotracers
 Mining industry: Radiotracers are used to monitor the movement of ores and minerals
through processing plants, and to determine the effectiveness of mineral extraction
processes (Pant, 2022).
 Chemical industry: The mixing and dispersion of chemicals in industrial reactors and
performance of catalysts can be monitored with radiotracers (Pant, 2022).
 Environmental industry: The path of contaminants in the environment and the
effectiveness of environmental remediation processes can be monitor with radiotracers
(Pant, 2022).
2.6.5 RADIOISOTOPES IN CONSUMER PRODUCTS

A variety of consumer products contain radioactive isotopes, either as a working part of the
product or as naturally occurring radioactive material (Shaw et al., 2007). The most commonly
used radioisotopes are:

 Tritium (H-3): used in exit signs, watches and other products that require long-lasting,
low-level illuminations (Shaw et al., 2007).
 Americium (Am-241): used in smoke detectors.
 Radium (Ra-226): used in older watches or clocks, to make dials and numbers visible in
the dark (Shaw et al., 2007).
 Thorium (Th-232): used in the production of lenses for cameras. However, due to its
radioactivity, the use of Th has declined in recent years.
 Uranium: used as a colorant in glass or ceramics before the adverse effects of its
radioactivity were understood (Shaw et al., 2007).

2.6.6 RADIOISOTOPES IN POWER SYSTEMS

Radioisotope power systems generate energy from the natural decay of radionuclides. A
radioisotope power system typically consists of two main parts, namely, a radioisotope heat
source and an energy conversion system. During the decay process, heat is produced and is
partially transformed into electricity (Lange and Carroll, 2008). These power systems are known
for being rugged, compact, and highly reliable. They can also be produced and used safely, with
minimal risk both to the operating personnel and the public, as well as having minimal impact on
the environment. The most commonly used isotopes for this purpose are plutonium-238 (Pu-238)
and strontium-90 (Sr-90). Both isotopes emit alpha particles, which can be converted to heat
when they collide with a solid material. Radioisotope power systems are commonly used in
space probes and satellites, where they can provide a reliable source of power for many years
without the need for refueling or maintenance. They are also used in remote or harsh
environments on Earth, such as in buoys, lighthouses, and research stations in Antarctica (Lange
and Carroll, 2008).

You might also like