You are on page 1of 18

PHYSICS

MR. FATIH SEREN

HISTORY
OF THE ATOM
___

Chiosa Andrei

History of the atom


Matter is composed of indivisible building blocks. This idea was recorded as early as the 5th
century BCE by Leucippus and Democritus. The Greeks called these particles atomos, meaning
indivisible, and the modern word “atom” is derived from this term.

Dalton’s Model

Part 1: All matter is made of atoms.

Dalton hypothesized that the law of conservation of mass and the law of definite proportions
could be explained using the idea of atoms. He proposed that all matter is made of tiny
2

indivisible particles called atoms, which he imagined as "solid, massy, hard, impenetrable,
movable particle(s)".

It is important to note that since Dalton did not have the necessary instruments to see or
otherwise experiment on individual atoms, he did not have any insight into whether they might
have any internal structure. We might visualize Dalton's atom as a piece in a molecular
modeling kit, where different elements are spheres of different sizes and colors. While this is a
handy model for some applications, we now know that atoms are far from being solid spheres.

Part 2: All atoms of a given element are identical in mass and properties.

Dalton proposed that every single atom of an element, such as gold, is the same as every other
atom of that element. He also noted that the atoms of one element differ from the atoms of all
other elements. Today, we still know this to be mostly true. A sodium atom is different from a
carbon atom. Elements may share some similar boiling points, melting points, and
electronegativities, but no two elements have the same exact set of properties.

Part 3: Compounds are combinations of two or more different types of atoms.

In the third part of Dalton's atomic theory, he proposed that compounds are combinations of
two or more different types of atoms. An example of such a compound is table salt. Table salt is
a combination of two separate elements with unique physical and chemical properties. The first,
sodium, is a highly reactive metal. The second, chlorine, is a toxic gas. When they react, the
atoms combine in a 1:1 ratio to form white crystals of NaCl.

Part 4: A chemical reaction is a rearrangement of atoms.

In the fourth and final part of Dalton's atomic theory, he suggested that chemical reactions
don't destroy or create atoms. They merely rearranged the atoms. Using our salt example again,
when sodium combines with chlorine to make salt, both the sodium and chlorine atoms still
exist. They simply rearrange to form a new compound.

Thomson’s Model

Postulates of Thomson’s atomic model


3

Postulate 1: An atom consists of a positively charged sphere with electrons embedded in it.
Postulate 2: An atom as a whole is electrically neutral because the negative and positive charges
are equal in magnitude

Thomson atomic model is compared to watermelon. Where he considered:

Watermelon seeds as negatively charged particles

The red part of the watermelon as positively charged

Limitations of Thomson’s atomic model

It failed to explain the stability of an atom because his model of atom failed to explain how a
positive charge holds the negatively charged electrons in an atom. Therefore, This theory also
failed to account for the position of the nucleus in an atom

Thomson’s model failed to explain the scattering of alpha particles by thin metal foils

No experimental evidence in its support

Rutherfords’s Model

Rutherford model, also called Rutherford atomic model, nuclear atom, or planetary model of
the atom, description of the structure of atoms proposed (1911) by the New Zealand-born
physicist Ernest Rutherford. The model described the atom as a tiny, dense, positively charged
core called a nucleus, in which nearly all the mass is concentrated, around which the light,
negative constituents, called electrons, circulate at some distance, much like planets revolving
around the Sun.

The nucleus was postulated as small and dense to account for the scattering of alpha particles
from thin gold foil, as observed in a series of experiments performed by undergraduate Ernest
Marsden under the direction of Rutherford and German physicist Hans Geiger in 1909. A
radioactive source emitting alpha particles (i.e., positively charged particles, identical to the
helium atom nucleus and 7,000 times more massive than electrons) was enclosed within a
protective lead shield. The radiation was focused into a narrow beam after passing through a slit
in a lead screen. A thin section of gold foil was placed in front of the slit, and a screen coated
with zinc sulfide to render it fluorescent served as a counter to detect alpha particles. As each
alpha particle struck the fluorescent screen, it produced a burst of light called a scintillation,
which was visible through a viewing microscope attached to the back of the screen. The screen
4

itself was movable, allowing Rutherford and his associates to determine whether or not any
alpha particles were being deflected by the gold foil.

Most alpha particles passed straight through the gold foil, which implied that atoms are mostly
composed of open space. Some alpha particles were deflected slightly, suggesting interactions
with other positively charged particles within the atom. Still other alpha particles were
scattered at large angles, while a very few even bounced back toward the source. (Rutherford
famously said later, “It was almost as incredible as if you fired a 15-inch shell at a piece of tissue
paper and it came back and hit you.”) Only a positively charged and relatively heavy target
particle, such as the proposed nucleus, could account for such strong repulsion. The negative
electrons that balanced electrically the positive nuclear charge were regarded as traveling in
circular orbits about the nucleus. The electrostatic force of attraction between electrons and
nucleus was likened to the gravitational force of attraction between the revolving planets and
the Sun. Most of this planetary atom was open space and offered no resistance to the passage of
the alpha particles.

The Rutherford model supplanted the “plum-pudding” atomic model of English physicist Sir J.J.
Thomson, in which the electrons were embedded in a positively charged atom like plums in a
pudding. Based wholly on classical physics, the Rutherford model itself was superseded in a few
years by the Bohr atomic model, which incorporated some early quantum theory.

Atomic Spectra
The emission spectrum of a chemical element or chemical compound is the spectrum of
frequencies of electromagnetic radiation emitted due to an electron making a transition from a
high energy state to a lower energy state. The photon energy of the emitted photon is equal to
the energy difference between the two states. There are many possible electron transitions for
each atom, and each transition has a specific energy difference. This collection of different
transitions, leading to different radiated wavelengths, make up an emission spectrum. Each
element's emission spectrum is unique. Therefore, spectroscopy can be used to identify
elements in matter of unknown composition. Similarly, the emission spectra of molecules can
be used in chemical analysis of substances.

In physics, emission is the process by which a higher energy quantum mechanical state of a
particle becomes converted to a lower one through the emission of a photon, resulting in the
production of light. The frequency of light emitted is a function of the energy of the transition.
5

Since energy must be conserved, the energy difference between the two states equals the energy
carried off by the photon. The energy states of the transitions can lead to emissions over a very
large range of frequencies. For example, visible light is emitted by the coupling of electronic
states in atoms and molecules (then the phenomenon is called fluorescence or
phosphorescence). On the other hand, nuclear shell transitions can emit high energy gamma
rays, while nuclear spin transitions emit low energy radio waves.

The emittance of an object quantifies how much light is emitted by it. This may be related to
other properties of the object through the Stefan–Boltzmann law. For most substances, the
amount of emission varies with the temperature and the spectroscopic composition of the
object, leading to the appearance of color temperature and emission lines. Precise
measurements at many wavelengths allow the identification of a substance via emission
spectroscopy.

Emission of radiation is typically described using semi-classical quantum mechanics: the


particle's energy levels and spacings are determined from quantum mechanics, and light is
treated as an oscillating electric field that can drive a transition if it is in resonance with the
system's natural frequency. The quantum mechanics problem is treated using time-dependent
perturbation theory and leads to the general result known as Fermi's golden rule. The
description has been superseded by quantum electrodynamics, although the semi-classical
version continues to be more useful in most practical computations.
6

Excitation of Atoms
In quantum mechanics, an excited state of a system (such as an atom, molecule or nucleus) is
any quantum state of the system that has a higher energy than the ground state (that is, more
energy than the absolute minimum). Excitation refers to an increase in energy level above a
chosen starting point, usually the ground state, but sometimes an already excited state. The
temperature of a group of particles is indicative of the level of excitation (with the notable
exception of systems that exhibit negative temperature).

The lifetime of a system in an excited state is usually short: spontaneous or induced emission of
a quantum of energy (such as a photon or a phonon) usually occurs shortly after the system is
promoted to the excited state, returning the system to a state with lower energy (a less excited
state or the ground state). This return to a lower energy level is often loosely described as decay
and is the inverse of excitation.

Long-lived excited states are often called metastable. Long-lived nuclear isomers and singlet
oxygen are two examples of this.

Frank- Hertz Experiment

The Franck–Hertz experiment was the first electrical measurement to clearly show the quantum
nature of atoms, and thus "transformed our understanding of the world". It was presented on
7

April 24, 1914, to the German Physical Society in a paper by James Franck and Gustav Hertz.
Franck and Hertz had designed a vacuum tube for studying energetic electrons that flew
through a thin vapor of mercury atoms. They discovered that, when an electron collided with a
mercury atom, it could lose only a specific quantity (4.9 electron volts) of its kinetic energy
before flying away. This energy loss corresponds to decelerating the electron from a speed of
about 1.3 million meters per second to zero. A faster electron does not decelerate completely
after a collision, but loses precisely the same amount of its kinetic energy. Slower electrons
merely bounce off mercury atoms without losing any significant speed or kinetic energy.

Franck and Hertz's original experiment used a heated vacuum tube containing a drop of
mercury; they reported a tube temperature of 115 °C, at which the vapor pressure of mercury is
about 100 pascals (and far below atmospheric pressure).[2][12] A contemporary Franck–Hertz
tube is shown in the photograph. It is fitted with three electrodes: an electron-emitting, hot
cathode; a metal mesh grid; and an anode. The grid's voltage is positive relative to the cathode,
so that electrons emitted from the hot cathode are drawn to it. The electric current measured in
the experiment is due to electrons that pass through the grid and reach the anode. The anode's
electric potential is slightly negative relative to the grid, so that electrons that reach the anode
have at least a corresponding amount of kinetic energy after passing the grid.

Wavelengths of light emitted by a mercury vapor discharge and by a Franck–Hertz tube in


operation at 10 V. The Franck–Hertz tube primarily emits light with a wavelength near 254
nanometers; the discharge emits light at many wavelengths. Based on the original 1914 figure.
The graphs published by Franck and Hertz (see figure) show the dependence of the electric
current flowing out of the anode upon the electric potential between the grid and the cathode.

At low potential differences—up to 4.9 volts—the current through the tube increased steadily
with increasing potential difference. This behavior is typical of true vacuum tubes that don't
contain mercury vapor; larger voltages lead to larger "space-charge limited current".

At 4.9 volts the current drops sharply, almost back to zero.

The current then increases steadily once again as the voltage is increased further, until 9.8 volts
is reached (exactly 4.9+4.9 volts).

At 9.8 volts a similar sharp drop is observed.

While it isn't evident in the original measurements of the figure, this series of dips in current at
approximately 4.9 volt increments continues to potentials of at least 70 volts.
8

Franck and Hertz noted in their first paper that the 4.9 eV characteristic energy of their
experiment corresponded well to one of the wavelengths of light emitted by mercury atoms in
gas discharges. They were using a quantum relationship between the energy of excitation and
the corresponding wavelength of light, which they broadly attributed to Johannes Stark and to
Arnold Sommerfeld; it predicts that 4.9 eV corresponds to light with a 254 nm wavelength.[2]
The same relationship was also incorporated in Einstein's 1905 photon theory of the
photoelectric effect.[15] In a second paper, Franck and Hertz reported the optical emission from
their tubes, which emitted light with a single prominent wavelength 254 nm.[7] The figure at
the right shows the spectrum of a Franck–Hertz tube; nearly all of the light emitted has a single
wavelength. For reference, the figure also shows the spectrum for a mercury gas discharge light,
which emits light at several wavelengths besides 254 nm. The figure is based on the original
spectra published by Franck and Hertz in 1914. The fact that the Franck–Hertz tube emitted just
the single wavelength, corresponding nearly exactly to the voltage period they had measured,
was very important.

Radioactivity
Radioactive decay (also known as nuclear decay, radioactivity, radioactive disintegration, or
nuclear disintegration) is the process by which an unstable atomic nucleus loses energy by
radiation. A material containing unstable nuclei is considered radioactive. Three of the most
common types of decay are alpha decay (α-decay), beta decay (β-decay), and gamma decay
(γ-decay), all of which involve emitting one or more particles. The weak force is the mechanism
that is responsible for beta decay, while the other two are governed by the electromagnetism
and nuclear force.[1] A fourth type of common decay is electron capture, in which an unstable
nucleus captures an inner electron from one of the electron shells. The loss of that electron
9

from the shell results in a cascade of electrons dropping down to that lower shell resulting in
emission of discrete X-rays from the transitions. A common example is iodine-125 commonly
used in medical settings.

Alpha Decay

Alpha decay or α-decay is a type of radioactive decay in which an atomic nucleus emits an alpha
particle (helium nucleus) and thereby transforms or 'decays' into a different atomic nucleus,
with a mass number that is reduced by four and an atomic number that is reduced by two. An
alpha particle is identical to the nucleus of a helium-4 atom, which consists of two protons and
two neutrons. It has a charge of +2 e and a mass of 4 u. For example, uranium-238 decays to
form thorium-234.

While alpha particles have a charge +2 e, this is not usually shown because a nuclear equation
describes a nuclear reaction without considering the electrons – a convention that does not
imply that the nuclei necessarily occur in neutral atoms.

Alpha decay typically occurs in the heaviest nuclides. Theoretically, it can occur only in nuclei
somewhat heavier than nickel (element 28), where the overall binding energy per nucleon is no
longer a maximum and the nuclides are therefore unstable toward spontaneous fission-type
processes. In practice, this mode of decay has only been observed in nuclides considerably
heavier than nickel, with the lightest known alpha emitters being the lightest isotopes (mass
numbers 104–109) of tellurium (element 52). Exceptionally, however, beryllium-8 decays to two
alpha particles.

Alpha decay is by far the most common form of cluster decay, where the parent atom ejects a
defined daughter collection of nucleons, leaving another defined product behind. It is the most
common form because of the combined extremely high nuclear binding energy and relatively
small mass of the alpha particle. Like other cluster decays, alpha decay is fundamentally a
quantum tunneling process. Unlike beta decay, it is governed by the interplay between both the
strong nuclear force and the electromagnetic force.

Alpha particles have a typical kinetic energy of 5 MeV (or ≈ 0.13% of their total energy, 110
TJ/kg) and have a speed of about 15,000,000 m/s, or 5% of the speed of light. There is
surprisingly small variation around this energy, due to the strong dependence of the half-life of
this process on the energy produced. Because of their relatively large mass, the electric charge
of +2 e and relatively low velocity, alpha particles are very likely to interact with other atoms
and lose their energy, and their forward motion can be stopped by a few centimeters of air.
10

Approximately 99% of the helium produced on Earth is the result of the alpha decay of
underground deposits of minerals containing uranium or thorium. The helium is brought to the
surface as a by-product of natural gas production.

Beta Decay

In nuclear physics, beta decay (β-decay) is a type of radioactive decay in which a beta particle
(fast energetic electron or positron) is emitted from an atomic nucleus, transforming the
original nuclide to an isobar of that nuclide. For example, beta decay of a neutron transforms it
into a proton by the emission of an electron accompanied by an antineutrino; or, conversely a
proton is converted into a neutron by the emission of a positron with a neutrino in so-called
positron emission. Neither the beta particle nor its associated (anti-)neutrino exist within the
nucleus prior to beta decay, but are created in the decay process. By this process, unstable
11

atoms obtain a more stable ratio of protons to neutrons. The probability of a nuclide decaying
due to beta and other forms of decay is determined by its nuclear binding energy. The binding
energies of all existing nuclides form what is called the nuclear band or valley of stability.[1] For
either electron or positron emission to be energetically possible, the energy release (see below)
or Q value must be positive.

Beta decay is a consequence of the weak force, which is characterized by relatively lengthy
decay times. Nucleons are composed of up quarks and down quarks,[2] and the weak force
allows a quark to change its flavour by emission of a W boson leading to creation of an
electron/antineutrino or positron/neutrino pair. For example, a neutron, composed of two down
quarks and an up quark, decays to a proton composed of a down quark and two up quarks.

Electron capture is sometimes included as a type of beta decay,[3] because the basic nuclear
process, mediated by the weak force, is the same. In electron capture, an inner atomic electron
is captured by a proton in the nucleus, transforming it into a neutron, and an electron neutrino
is released.

Gamma Decay

A gamma ray, also known as gamma radiation.

Gamma decay, is a penetrating form of electromagnetic radiation arising from the radioactive
decay of atomic nuclei. It consists of the shortest wavelength electromagnetic waves, typically
shorter than those of X-rays. With frequencies above 30 exahertz (3×1019 Hz), it imparts the
highest photon energy. Paul Villard, a French chemist and physicist, discovered gamma
radiation in 1900 while studying radiation emitted by radium. In 1903, Ernest Rutherford named
this radiation gamma rays based on their relatively strong penetration of matter; in 1900 he had
already named two less penetrating types of decay radiation (discovered by Henri Becquerel)
alpha rays and beta rays in ascending order of penetrating power.
12

Gamma rays from radioactive decay are in the energy range from a few kiloelectronvolts (keV)
to approximately 8 megaelectronvolts (MeV), corresponding to the typical energy levels in
nuclei with reasonably long lifetimes. The energy spectrum of gamma rays can be used to
identify the decaying radionuclides using gamma spectroscopy. Very-high-energy gamma rays
in the 100–1000 teraelectronvolt (TeV) range have been observed from sources such as the
Cygnus X-3 microquasar.

Natural sources of gamma rays originating on Earth are mostly a result of radioactive decay and
secondary radiation from atmospheric interactions with cosmic ray particles. However, there are
other rare natural sources, such as terrestrial gamma-ray flashes, which produce gamma rays
from electron action upon the nucleus. Notable artificial sources of gamma rays include fission,
such as that which occurs in nuclear reactors, and high energy physics experiments, such as
neutral pion decay and nuclear fusion.

Gamma rays and X-rays are both electromagnetic radiation, and since they overlap in the
electromagnetic spectrum, the terminology varies between scientific disciplines. In some fields
of physics, they are distinguished by their origin: Gamma rays are created by nuclear decay
while X-rays originate outside the nucleus. In astrophysics, gamma rays are conventionally
defined as having photon energies above 100 keV and are the subject of gamma ray astronomy,
while radiation below 100 keV is classified as X-rays and is the subject of X-ray astronomy.

Gamma rays are ionizing radiation and are thus hazardous to life. Due to their high penetration
power, they can damage bone marrow and internal organs. Unlike alpha and beta rays, they
easily pass through the body and thus pose a formidable radiation protection challenge,
requiring shielding made from dense materials such as lead or concrete. On Earth, the
magnetosphere protects life from most types of lethal cosmic radiation other than gamma rays.
13

Half-life

half-life, in radioactivity, the interval of time required for one-half of the atomic nuclei of a
radioactive sample to decay (change spontaneously into other nuclear species by emitting
particles and energy), or, equivalently, the time interval required for the number of
disintegrations per second of a radioactive material to decrease by one-half.

The radioactive isotope cobalt-60, which is used for radiotherapy, has, for example, a half-life of
5.26 years. Thus after that interval, a sample originally containing 8 g of cobalt-60 would
contain only 4 g of cobalt-60 and would emit only half as much radiation. After another interval
14

of 5.26 years, the sample would contain only 2 g of cobalt-60. Neither the volume nor the mass
of the original sample visibly decreases, however, because the unstable cobalt-60 nuclei decay
into stable nickel-60 nuclei, which remain with the still-undecayed cobalt.

Half-lives are characteristic properties of the various unstable atomic nuclei and the particular
way in which they decay. Alpha and beta decay are generally slower processes than gamma
decay. Half-lives for beta decay range upward from one-hundredth of a second and, for alpha
decay, upward from about one one-millionth of a second. Half-lives for gamma decay may be
too short to measure (around 10-14 second), though a wide range of half-lives for gamma
emission has been reported.

The Discovery of Protons and Neutrons


Discovery of Protons

The discovery of protons dates back to the year 1815 when the English chemist William Prout
suggested that all atoms are made up of hydrogen atoms (which he referred to as protyles).
When canal rays (positively charged ions formed by gases) were discovered by the German
physicist Eugen Goldstein in the year 1886, it was observed that the charge-to-mass ratio of the
hydrogen ion was the highest among all gases. It was also observed that the hydrogen ion had
the smallest size among all ionized gases.

The nucleus of the atom was discovered by Ernest Rutherford in the year 1911 in his famous
gold foil experiment. He concluded that all the positively charged particles in an atom were
concentrated in a singular core and that most of the atom’s volume was empty. He also stated
that the total number of positively charged particles in the nucleus is equal to the total number
of negatively charged electrons present around it.

How was the Proton Discovered?

Ernest Rutherford observed that his scintillation detectors detected hydrogen nuclei when a
beam of alpha particles was shot into the air.

After investigating further, Rutherford found that these hydrogen nuclei were produced from
the nitrogen atoms present in the atmosphere.

He then proceeded to fire beams of alpha particles into pure nitrogen gas and observed that a
greater number of hydrogen nuclei were produced.
15

He concluded that the hydrogen nuclei originated from the nitrogen atom, proving that the
hydrogen nucleus was a part of all other atoms.

This experiment was the first to report a nuclear reaction, given by the equation: 14N + α → 17O
+ p [Where α is an alpha particle which contains two protons and two neutrons, and ‘p’ is a
proton]

The hydrogen nucleus was later named ‘proton’ and recognized as one of the building blocks of
the atomic nucleus.

Discovery of Neutrons

The discovery of neutrons can be traced back to the year 1930 when the German nuclear
physicists Herbert Becker and Walther Bothe observed that a penetrating form of radiation was
produced when the alpha particles emitted by polonium was incident on relatively light
elements such as lithium, beryllium, and boron. This penetrating radiation was unaffected by
electric fields and was, therefore, assumed to be gamma radiation.

In the year 1932, the French scientists Frederic Joliot-Curie and Irene Joliot-Curie observed that
this unusually penetrating radiation, when incident on paraffin wax (or other compounds rich
in hydrogen), caused the ejection of high energy protons (~5 MeV). The Italian physicist Ettore
Majorana suggested the existence of a neutral particle in the nucleus of the atom which was
responsible for the manner in which the radiation interacted with protons.

The presence of neutral particles in the nuclei of atoms was also suggested by Ernest Rutherford
in the year 1920. He suggested that a neutrally charged particle, consisting of a proton and an
electron bound to each other, also resided in the nuclei of atoms. He coined the term ‘neutron’
to refer to these neutrally charged particles.

How were Neutrons Discovered?

James Chadwick fired alpha radiation at beryllium sheet from a polonium source. This led to the
production of an uncharged, penetrating radiation.

This radiation was made incident on paraffin wax, a hydrocarbon having a relatively high
hydrogen content.
16

The protons ejected from the paraffin wax (when struck by the uncharged radiation) were
observed with the help of an ionization chamber.

The range of the liberated protons was measured and the interaction between the uncharged
radiation and the atoms of several gases was studied by Chadwick.

He concluded that the unusually penetrating radiation consisted of uncharged particles having
(approximately) the same mass as a proton. These particles were later termed ‘neutrons’.

Nuclear Reactions
In nuclear physics and nuclear chemistry, a nuclear reaction is a process in which two nuclei, or
a nucleus and an external subatomic particle, collide to produce one or more new nuclides.
Thus, a nuclear reaction must cause a transformation of at least one nuclide to another. If a
nucleus interacts with another nucleus or particle and they then separate without changing the
nature of any nuclide, the process is simply referred to as a type of nuclear scattering, rather
than a nuclear reaction.

In principle, a reaction can involve more than two particles colliding, but because the
probability of three or more nuclei to meet at the same time at the same place is much less than
for two nuclei, such an event is exceptionally rare (see triple alpha process for an example very
close to a three-body nuclear reaction). The term "nuclear reaction" may refer either to a change
in a nuclide induced by collision with another particle or to a spontaneous change of a nuclide
without collision.

Natural nuclear reactions occur in the interaction between cosmic rays and matter, and nuclear
reactions can be employed artificially to obtain nuclear energy, at an adjustable rate,
on-demand. Nuclear chain reactions in fissionable materials produce induced nuclear fission.
Various nuclear fusion reactions of light elements power the energy production of the Sun and
stars.

Nuclear reactions may be shown in a form similar to chemical equations, for which invariant
mass must balance for each side of the equation, and in which transformations of particles must
follow certain conservation laws, such as conservation of charge and baryon number (total
atomic mass number). An example of this notation follows:
17

To balance the equation above for mass, charge and mass number, the second nucleus to the
right must have atomic number 2 and mass number 4; it is therefore also helium-4. The
complete equation therefore reads:

or more simply:

Instead of using the full equations in the style above, in many situations a compact notation is
used to describe nuclear reactions. This style of the form A(b,c)D is equivalent to A + b
producing c + D. Common light particles are often abbreviated in this shorthand, typically p for
proton, n for neutron, d for deuteron, α representing an alpha particle or helium-4, β for beta
particle or electron, γ for gamma photon, etc. The reaction above would be written as 6Li(d,α)α
18

You might also like