206 A Physics
206 A Physics
206 A Physics
Atoms
12.1 INTRODUCTION
Ernst Rutherford (1871 1937) British physicist who did pioneering work on
radioactive radiation. He discovered alpha-rays and beta-rays. Along with
Federick Soddy, he created the modern theory of radioactivity. He studied the
emanation of thorium and discovered a new noble gas, an isotope of radon,
now known as thoron. By scattering alpha-rays from the metal foils, he
discovered the atomic nucleus and proposed the plenatery model of the atom.
He also estimated the approximate size of the nucleus.
(12.1)
where r is the distance between the -particle and the nucleus. The
force is directed along the line joining the -particle and the nucleus.
The magnitude and direction of the force on an -particle continuously
changes as it approaches the nucleus and recedes away from it.
Figure 12.4 Trajectory of -particles in the coulomb field of a target nucleus. The
impact parameter, b and scattering angle are also depicted.
Example 12.1 In the Rutherfords nuclear model of the atom, the nucleus
(radius about 1015 m) is analogous to the sun about which the electron move
in orbit (radius 1010 m) like the earth orbits around the sun. If the
dimensions of the solar system had the same proportions as those of the
atom, would the earth be closer to or farther away from the sun than actually it
is? The radius of earths orbit is about 1.5 1011 m. The radius of sun is taken
as 7 108 m.
Solution The ratio of the radius of electrons orbit to the radius of nucleus is
(1010 m)/(1015 m) = 105, that is, the radius of the electrons orbit is 105
times larger than the radius of nucleus. If the radius of the earths orbit around
the sun were 105 times larger than the radius of the sun, the radius of the
earths orbit would be 105 7 108 m =
7 1013 m. This is more than 100 times greater than the actual orbital radius
of earth. Thus, the earth would be much farther away from the sun.
It implies that an atom contains a much greater fraction of empty space than
our solar system does.
Example 12.2 In a Geiger-Marsden experiment, what is the distance of closest
approach to the nucleus of a 7.7 MeV -particle before it comes momentarily
to rest and reverses its direction?
Solution The key idea here is that throughout the scattering process, the total
mechanical energy of the system consisting of an -particle and a gold nucleus
is conserved. The systems initial mechanical energy is Ei, before the particle
and nucleus interact, and it is equal to its mechanical energy Ef when the -
particle momentarily stops. The initial energy Ei is just the kinetic energy K of
the incoming - particle. The final energy Ef is just the electric potential energy
U of the system. The potential energy U can be calculated from Eq. (12.1).
Let d be the centre-to-centre distance between the -particle and the gold
nucleus when the -particle is at its stopping point. Then we can write the
conservation of energy Ei = Ef as
= 3.84 1016 Z m
The atomic number of foil material gold is Z = 79, so that
14 m = 30 fm. (1 fm (i.e. fermi) = 1015 m.)
d (Au) = 3.0 10
The radius of gold nucleus is, therefore, less than 3.0 1014 m. This is not in
very good agreement with the observed result as the actual radius of gold
nucleus is 6 fm. The cause of discrepancy is that the distance of closest
approach is considerably larger than the sum of the radii of the gold nucleus
and the -particle. Thus, the -particle reverses its motion without ever actually
touching the gold nucleus.
(12.2)
Thus the relation between the orbit radius and the electron velocity is
(12.3)
The kinetic energy (K) and electrostatic potential energy (U) of the
electron in hydrogen atom are
(12.4)
The total energy of the electron is negative. This implies the fact that
the electron is bound to the nucleus. If E were positive, an electron will
not follow a closed orbit around the nucleus.
= 5.3 1011 m.
The velocity of the revolving electron can be computed from Eq. (12.3) with m
= 9.1 1031 kg,
(12.5)
where is the wavelength, R is a constant called the Rydberg
constant, and n may have integral values 3, 4, 5, etc. The value of R is
1.097 107 m1. This equation is also called Balmer formula.
Taking n = 3 in Eq. (12.5), one obtains the wavelength of the H line:
= 1.522 106 m1
i.e., = 656.3 nm
For n = 4, one obtains the wavelength of H line, etc. For n = , one
obtains the limit of the series, at = 364.6 nm. This is the shortest
wavelength in the Balmer series. Beyond this limit, no further distinct
lines appear, instead only a faint continuous spectrum is seen.
n = 2,3,4... (12.6)
Paschen series:
n = 4,5,6...
(12.7)
Brackett series:
n = 5,6,7... (12.8)
Pfund series:
n = 6,7,8... (12.9)
The Lyman series is in the ultraviolet, and the Paschen and Brackett
series are in the infrared region.
The Balmer formula Eq. (12.5) may be written in terms of frequency of
the light, recalling that
c =
or
Thus, Eq. (12.5) becomes
(12.10)
There are only a few elements (hydrogen, singly ionised helium, and
doubly ionised lithium) whose spectra can be represented by simple
formula like Eqs. (12.5) (12.9).
Equations (12.5) (12.9) are useful as they give the wavelengths that
hydrogen atoms radiate or absorb. However, these results are
empirical and do not give any reasoning why only certain frequencies
are observed in the hydrogen spectrum.
Niels Henrik David Bohr (1885 1962) Danish physicist who explained the
spectrum of hydrogen atom based on quantum ideas. He gave a theory of
nuclear fission based on the liquid-drop model of nucleus. Bohr contributed to
the clarification of conceptual problems in quantum mechanics, in particular by
proposing the comple- mentary principle.
Example 12.4 According to the classical electromagnetic theory, calculate the
initial frequency of the light emitted by the electron revolving around a proton in
hydrogen atom.
Solution From Example 12.3 we know that velocity of electron moving around
a proton in hydrogen atom in an orbit of radius 5.3 1011 m is 2.2 106 m/s.
Thus, the frequency of the electron moving around the proton is
It was Niels Bohr (1885 1962) who made certain modifications in this
model by adding the ideas of the newly developing quantum
hypothesis. Niels Bohr studied in Rutherfords laboratory for several
months in 1912 and he was convinced about the validity of Rutherford
nuclear model. Faced with the dilemma as discussed above, Bohr, in
1913, concluded that in spite of the success of electromagnetic theory
in explaining large-scale phenomena, it could not be applied to the
processes at the atomic scale. It became clear that a fairly radical
departure from the established principles of classical mechanics and
electromagnetism would be needed to understand the structure of
atoms and the relation of atomic structure to atomic spectra. Bohr
combined classical and early quantum concepts and gave his theory
in the form of three postulates. These are :
(i) Bohrs first postulate was that an electron in an atom could revolve
in certain stable orbits without the emission of radiant energy, contrary
to the predictions of electromagnetic theory. According to this
postulate, each atom has certain definite stable states in which it can
exist, and each possible state has definite total energy. These are
called the stationary states of the atom.
(ii) Bohrs second postulate defines these stable orbits. This postulate
states that the electron revolves around the nucleus only in those
orbits for which the angular momentum is some integral multiple of
h/2 where h is the Plancks constant (= 6.6 1034 J s). Thus the
angular momentum (L) of the orbiting electron is quantised. That is
L = nh/2 (12.11)
(iii) Bohrs third postulate incorporated into atomic theory the early
quantum concepts that had been developed by Planck and Einstein. It
states that an electron might make a transition from one of its
specified non-radiating orbits to another of lower energy. When it does
so, a photon is emitted having energy equal to the energy difference
between the initial and final states. The frequency of the emitted
photon is then given by
h = Ei Ef (12.12)
where Ei and Ef are the energies of the initial and final states and Ei >
Ef.
For a hydrogen atom, Eq. (12.4) gives the expression to determine the
energies of different energy states. But then this equation requires the
radius r of the electron orbit. To calculate r, Bohrs second postulate
about the angular momentum of the electronthe quantisation
condition is used. The angular momentum L is given by
L = mvr
Bohrs second postulate of quantisation [Eq. (12.11)] says that the
allowed values of angular momentum are integral multiples of h/2.
Ln = mvnrn = (12.13)
where n is an integer, rn is the radius of nth possible orbit and vn is the
speed of moving electron in the nth orbit. The allowed orbits are
numbered 1, 2, 3 ..., according to the values of n, which is called the
principal quantum number of the orbit.
From Eq. (12.3), the relation between vn and rn is
(12.14)
and
(12.15)
Eq. (12.14) depicts that the orbital speed in the nth orbit falls by a
factor of n. Using Eq. (12.15), the size of the innermost orbit (n = 1)
can be obtained as
This is called the Bohr radius, represented by the symbol a0. Thus,
(12.16)
Substitution of values of h, m, 0 and e gives a0 = 5.29 1011 m.
From Eq. (12.15), it can also be seen that the radii of the orbits
increase as n2.
or (12.17)
Substituting values, Eq. (12.17) yields
(12.18)
Atomic energies are often expressed in electron volts (eV) rather than
joules. Since 1 eV = 1.6 1019 J, Eq. (12.18) can be rewritten as
(12.19)
The negative sign of the total energy of an electron moving in an orbit
means that the electron is bound with the nucleus. Energy will thus be
required to remove the electron from the hydrogen atom to a distance
infinitely far away from its nucleus (or proton in hydrogen atom).
The derivation of Eqs. (12.17) (12.19) involves the assumption that
the electronic orbits are circular, though orbits under inverse square
force are, in general elliptical. (Planets move in elliptical orbits under
the inverse square gravitational force of the sun.) However, it was
shown by the German physicist Arnold Sommerfeld (1868 1951)
that, when the restriction of circular orbit is relaxed, these equations
continue to hold even for elliptic orbits.
The energy of an atom is the least (largest negative value) when its
electron is revolving in an orbit closest to the nucleus i.e., the one for
which n = 1. For n = 2, 3, ... the absolute value of the energy E is
smaller, hence the energy is progressively larger in the outer orbits.
The lowest state of the atom, called the ground state, is that of the
lowest energy, with the electron revolving in the orbit of smallest
radius, the Bohr radius, a0. The energy of this state (n = 1), E1 is
13.6 eV. Therefore, the minimum energy required to free the electron
from the ground state of the hydrogen atom is 13.6 eV. It is called the
ionisation energy of the hydrogen atom. This prediction of the Bohrs
model is in excellent agreement with the experimental value of
ionisation energy.
We are introduced to the Bohr Model of atom one time or the other in the
course of physics. This model has its place in the history of quantum
mechanics and particularly in explaining the structure of an atom. It has
become a milestone since Bohr introduced the revolutionary idea of definite
energy orbits for the electrons, contrary to the classical picture requiring an
accelerating particle to radiate. Bohr also introduced the idea of quantisation of
angular momentum of electrons moving in definite orbits. Thus it was a semi-
classical picture of the structure of atom.
Now with the development of quantum mechanics, we have a better
understanding of the structure of atom. Solutions of the Schrdinger wave
equation assign a wave-like description to the electrons bound in an atom due
to attractive forces of the protons.
An orbit of the electron in the Bohr model is the circular path of motion of an
electron around the nucleus. But according to quantum mechanics, we cannot
associate a definite path with the motion of the electrons in an atom. We can
only talk about the probability of finding an electron in a certain region of space
around the nucleus. This probability can be inferred from the one-electron
wave function called the orbital. This function depends only on the coordinates
of the electron.
It is therefore essential that we understand the subtle differences that exist in
the two models:
hif = (12.21)
or if = (12.22)
* An electron can have any total energy above E = 0 eV. In such
situations the electron is free. Thus there is a continuum of energy
states above E = 0 eV, as shown in Fig. 12.8.
Figure 12.8 The energy level diagram for the hydrogen atom. The electron in a
hydrogen atom at room temperature spendsmost of its time in the ground state. To
ionise a hydrogen atom an electron from the ground state, 13.6 eV of energy must
be supplied. (The horizontal lines specify the presence of allowed energy states.)
R= (12.23)
If we insert the values of various constants in Eq. (12.23), we get
R = 1.03 107 m1
This is a value very close to the value (1.097 107 m1) obtained from
the empirical Balmer formula. This agreement between the theoretical
and experimental values of the Rydberg constant provided a direct
and striking confirmation of the Bohrs model.
The existence of discrete energy levels in an atom was directly verified in 1914
by James Franck and Gustav Hertz. They studied the spectrum of mercury
vapour when electrons having different kinetic energies passed through the
vapour. The electron energy was varied by subjecting the electrons to electric
fields of varying strength. The electrons collide with the mercury atoms and
can transfer energy to the mercury atoms. This can only happen when the
energy of the electron is higher than the energy difference between an energy
level of Hg occupied by an electron and a higher unoccupied level (see
Figure). For instance, the difference between an occupied energy level of Hg
and a higher unoccupied level is 4.9 eV. If an electron of having an energy of
4.9 eV or more passes through mercury, an electron in mercury atom can
absorb energy from the bombarding electron and get excited to the higher
level [Fig (a)]. The colliding electrons kinetic energy would reduce by this
amount.
The excited electron would subsequently fall back to the ground state by
emission of radiation [Fig. (b)]. The wavelength of emitted radiation is:
= 253 nm
By direct measurement, Franck and Hertz found that the emission spectrum of
mercury has a line corresponding to this wavelength. For this experimental
verification of Bohrs basic ideas of discrete energy levels in atoms and the
process of photon emission, Frank and Hertz were awarded the Nobel prize in
1925.
The various lines in the atomic spectra are produced when electrons
jump from higher energy state to a lower energy state and photons are
emitted. These spectral lines are called emission lines. But when an
atom absorbs a photon that has precisely the same energy needed by
the electron in a lower energy state to make transitions to a higher
energy state, the process is called absorption. Thus if photons with a
continuous range of frequencies pass through a rarefied gas and then
are analysed with a spectrometer, a series of dark spectral absorption
lines appear in the continuous spectrum. The dark lines indicate the
frequencies that have been absorbed by the atoms of the gas.
The explanation of the hydrogen atom spectrum provided by Bohrs
model was a brilliant achievement, which greatly stimulated progress
towards the modern quantum theory. In 1922, Bohr was awarded
Nobel Prize in Physics.
Figure 12.9 Line spectra originate in transitions between energy levels.
hc/if =
The wavelengths of the first four lines in the Lyman series correspond
to transitions from ni = 2,3,4,5 to nf = 1. We know that
= =
= 913.4 ni2/(ni2 1)
Of all the postulates, Bohr made in his model of the atom, perhaps the
most puzzling is his second postulate. It states that the angular
momentum of the electron orbiting around the nucleus is quantised
(that is, Ln = nh/2; n = 1, 2, 3 ). Why should the angular
momentum have only those values that are integral multiples of h/2?
The French physicist Louis de Broglie explained this puzzle in 1923,
ten years after Bohr proposed his model.
We studied, in Chapter 11, about the de Broglies hypothesis that
material particles, such as electrons, also have a wave nature. C. J.
Davisson and L. H. Germer later experimentally verified the wave
nature of electrons in 1927. Louis de Broglie argued that the electron
in its circular orbit, as proposed by Bohr, must be seen as a particle
wave. In analogy to waves travelling on a string, particle waves too
can lead to standing waves under resonant conditions. From Chapter
15 of Class XI Physics textbook, we know that when a string is
plucked, a vast number of wavelengths are excited. However only
those wavelengths survive which have nodes at the ends and form the
standing wave in the string. It means that in a string, standing waves
are formed when the total distance travelled by a wave down the
string and back is one wavelength, two wavelengths, or any integral
number of wavelengths. Waves with other wavelengths interfere with
themselves upon reflection and their amplitudes quickly drop to zero.
For an electron moving in nth circular orbit of radius rn, the total
distance is the circumference of the orbit, 2rn. Thus
2 rn = n, n = 1, 2, 3... (12.24)
Figure 12.10 illustrates a standing particle wave on a circular orbit for
n = 4, i.e., 2rn = 4, where is the de Broglie wavelength of the
electron moving in nth orbit. From Chapter 11, we have = h/p, where
p is the magnitude of the electrons momentum. If the speed of the
electron is much less than the speed of light, the momentum is mvn.
Thus, = h/mvn. From Eq. (12.24), we have
2 rn = n h/mvn or m vn rn = nh/2
This is the quantum condition proposed by Bohr for the angular
momentum of the electron [Eq. (12.13)]. In Section 12.5, we saw that
this equation is the basis of explaining the discrete orbits and energy
levels in hydrogen atom. Thus de Broglie hypothesis provided an
explanation for Bohrs second postulate for the quantisation of angular
momentum of the orbiting electron. The quantised electron orbits and
energy states are due to the wave nature of the electron and only
resonant standing waves can persist.
Bohrs model, involving classical trajectory picture (planet-like electron
orbiting the nucleus), correctly predicts the gross features of the
hydrogenic atoms*, in particular, the frequencies of the radiation
emitted or selectively absorbed. This model however has many
limitations.
Some are:
Figure 12.10 A standing wave is shown on a circular orbit where four de Broglie
wavelengths fit into the circumference of the orbit.
(ii) While the Bohrs model correctly predicts the frequencies of the
light emitted by hydrogenic atoms, the model is unable to explain the
relative intensities of the frequencies in the spectrum. In emission
spectrum of hydrogen, some of the visible frequencies have weak
intensity, others strong. Why? Experimental observations depict that
some transitions are more favoured than others. Bohrs model is
unable to account for the intensity variations.
Bohrs model presents an elegant picture of an atom and cannot be
generalised to complex atoms. For complex atoms we have to use a
new and radical theory based on Quantum Mechanics, which provides
a more complete picture of the atomic structure.
LASER LIGHT
This is similar to the difference between light emitted by an ordinary source like
a candle or a bulb and that emitted by a laser. The acronym LASER stands for
Light Amplification by Stimulated Emission of Radiation. Since its development
in 1960, it has entered into all areas of science and technology. It has found
applications in physics, chemistry, biology, medicine, surgery, engineering, etc.
There are low power lasers, with a power of 0.5 mW, called pencil lasers,
which serve as pointers. There are also lasers of different power, suitable for
delicate surgery of eye or glands in the stomach. Finally, there are lasers
which can cut or weld steel.
Light is emitted from a source in the form of packets of waves. Light coming
out from an ordinary source contains a mixture of many wavelengths. There is
also no phase relation between the various waves. Therefore, such light, even
if it is passed through an aperture, spreads very fast and the beam size
increases rapidly with distance. In the case of laser light, the wavelength of
each packet is almost the same. Also the average length of the packet of
waves is much larger. This means that there is better phase correlation over a
longer duration of time. This results in reducing the divergence of a laser beam
substantially.
If there are N atoms in a source, each emitting light with intensity I, then the
total intensity produced by an ordinary source is proportional to NI, whereas in
a laser source, it is proportional to N2I. Considering that N is very large, we
see that the light from a laser can be much stronger than that from an ordinary
source.
When astronauts of the Apollo missions visited the moon, they placed a mirror
on its surface, facing the earth. Then scientists on the earth sent a strong laser
beam, which was reflected by the mirror on the moon and received back on
the earth. The size of the reflected laser beam and the time taken for the round
trip were measured. This allowed a very accurate determination of (a) the
extremely small divergence of a laser beam and (b) the distance of the moon
from the earth.
Summary
2
= 13.6 eV/n
The n = 1 state is called ground state. In hydrogen atom the ground state
energy is 13.6 eV. Higher values of n correspond to excited states (n > 1).
Atoms are excited to these higher states by collisions with other atoms or
electrons or by absorption of a photon of right frequency.
EXERCISES
12.1 Choose the correct alternative from the clues given at the end
of the each statement:
(a) The size of the atom in Thomsons model is .......... the atomic
size in Rutherfords model. (much greater than/no different
from/much less than.)
(b) In the ground state of .......... electrons are in stable equilibrium,
while in .......... electrons always experience a net force.
(Thomsons model/ Rutherfords model.)
NUCLEI
13.1 Introduction
In the previous chapter, we have learnt that in every atom, the positive
charge and mass are densely concentrated at the centre of the atom
forming its nucleus. The overall dimensions of a nucleus are much
smaller than those of an atom. Experiments on scattering of -
particles demonstrated that the radius of a nucleus was smaller than
the radius of an atom by a factor of about 104. This means the volume
of a nucleus is about 1012 times the volume of the atom. In other
words, an atom is almost empty. If an atom is enlarged to the size of a
classroom, the nucleus would be of the size of pinhead. Nevertheless,
the nucleus contains most (more than 99.9%) of the mass of an atom.
Does the nucleus have a structure, just as the atom does? If so, what
are the constituents of the nucleus? How are these held together? In
this chapter, we shall look for answers to such questions. We shall
discuss various properties of nuclei such as their size, mass and
stability, and also associated nuclear phenomena such as
radioactivity, fission and fusion.
(13.1)
The atomic masses of various elements expressed in atomic mass
unit (u) are close to being integral multiples of the mass of a hydrogen
atom. There are, however, many striking exceptions to this rule. For
example, the atomic mass of chlorine atom is 35.46 u.
=
= 35.47 u
which agrees with the atomic mass of chlorine.
Even the lightest element, hydrogen has three isotopes having
masses 1.0078 u, 2.0141 u, and 3.0160 u. The nucleus of the lightest
atom of hydrogen, which has a relative abundance of 99.985%, is
called the proton. The mass of a proton is
(13.2)
This is equal to the mass of the hydrogen atom (= 1.00783u), minus
the mass of a single electron (me = 0.00055 u). The other two
isotopes of hydrogen are called deuterium and tritium. Tritium nuclei,
being unstable, do not occur naturally and are produced artificially in
laboratories.
The positive charge in the nucleus is that of the protons. A proton
carries one unit of fundamental charge and is stable. It was earlier
thought that the nucleus may contain electrons, but this was ruled out
later using arguments based on quantum theory. All the electrons of
an atom are outside the nucleus. We know that the number of these
electrons outside the nucleus of the atom is Z, the atomic number.
The total charge of the atomic electrons is thus (Ze), and since the
atom is neutral, the charge of the nucleus is (+Ze). The number of
protons in the nucleus of the atom is, therefore, exactly Z, the atomic
number.
Discovery of Neutron
mn = 1.00866 u = 1.67491027 kg
(13.3)
Chadwick was awarded the 1935 Nobel Prize in Physics for his
discovery of the neutron.
A free neutron, unlike a free proton, is unstable. It decays into a
proton, an electron and a antineutrino (another elementary particle),
and has a mean life of about 1000s. It is, however, stable inside the
nucleus.
The composition of a nucleus can now be described using the
following terms and symbols:
one proton and one neutron. Its other isotope tritium, , contains
one proton and two neutrons. The element gold has 32 isotopes,
ranging from A =173 to A = 204. We have already mentioned that
chemical properties of elements depend on their electronic structure.
As the atoms of isotopes have identical electronic structure they have
identical chemical behaviour and are placed in the same location in
the periodic table.
All nuclides with same mass number A are called isobars. For
Example 13.1 Given the mass of iron nucleus as 55.85u and A=56, find the
nuclear density?
Solution
Nuclear density = =
17 3
= 2.29 10 kg m
The density of matter in neutron stars (an astrophysical object) is comparable
to this density. This shows that matter in these objects has been compressed
to such an extent that they resemble a big nucleus.
= 8 2.01593 u = 16.12744 u.
Thus, we find that the mass of the nucleus is less than the total
mass of its constituents by 0.13691u. The difference in mass of a
nucleus and its constituents, M, is called the mass defect, and is
given by
(13.7)
What is the meaning of the mass defect? It is here that Einsteins
equivalence of mass and energy plays a role. Since the mass of the
oxygen nucleus is less that the sum of the masses of its constituents
(8 protons and 8 neutrons, in the unbound state), the equivalent
energy of the oxygen nucleus is less than that of the sum of the
equivalent energies of its constituents. If one wants to break the
oxygen nucleus into 8 protons and 8 neutrons, this extra energy M
c2, has to supplied. This energy required Eb is related to the mass
defect by
Eb = M c2 (13.8)
Example 13.3 Find the energy equivalent of one atomic mass unit, first in
Joules and then in MeV. Using this, express the mass defect of in
MeV/c2.
Solution
1u = 1.6605 1027 kg n
To convert it into energy units, we multiply it by c2 and find that energy
equivalent = 1.6605 1027 (2.9979 108)2 kg m2/s2
= 1.4924 1010 J
=
= 0.9315 109 eV
= 931.5 MeV
or , 1u = 931.5 MeV/c2
= 127.5 MeV/c2
We can think of binding energy per nucleon as the average energy per
nucleon needed to separate a nucleus into its individual nucleons.
Figure 13.1 is a plot of the binding energy per nucleon Ebn versus the
mass number A for a large number of nuclei. We notice the following
main features of
the plot:
(i) the binding energy per nucleon, Ebn, is practically constant, i.e.
practically independent of the atomic number for nuclei of middle
mass number ( 30 < A < 170). The curve has a maximum of about
8.75 MeV for A = 56 and has a value of 7.6 MeV for A = 238.
(ii) Ebn is lower for both light nuclei (A<30) and heavy nuclei (A>170).
We can draw some conclusions from these two observations:
(i) The force is attractive and sufficiently strong to produce a binding
energy of a few MeV per nucleon.
(ii) The constancy of the binding energy in the range 30 < A < 170 is a
consequence of the fact that the nuclear force is short-ranged.
Consider a particular nucleon inside a sufficiently large nucleus. It will
be under the influence of only some of its neighbours, which come
within the range of the nuclear force. If any other nucleon is at a
distance more than the range of the nuclear force from the particular
nucleon it will have no influence on the binding energy of the nucleon
under consideration. If a nucleon can have a maximum of p
neighbours within the range of nuclear force, its binding energy would
be proportional to p. Let the binding energy of the nucleus be pk,
where k is a constant having the dimensions of energy. If we increase
A by adding nucleons they will not change the binding energy of a
nucleon inside. Since most of the nucleons in a large nucleus reside
inside it and not on the surface, the change in binding energy per
nucleon would be small. The binding energy per nucleon is a constant
and is approximately equal to pk. The property that a given nucleon
influences only nucleons close to it is also referred to as saturation
property of the nuclear force.
(iii) A very heavy nucleus, say A = 240, has lower binding energy per
nucleon compared to that of a nucleus with A = 120. Thus if a nucleus
A = 240 breaks into two A = 120 nuclei, nucleons get more tightly
bound. This implies energy would be released in the process. It has
very important implications for energy production through fission, to be
discussed later in Section 13.7.1.
(iv) Consider two very light nuclei (A 10) joining to form a heavier
nucleus. The binding energy per nucleon of the fused heavier nuclei is
more than the binding energy per nucleon of the lighter nuclei. This
means that the final system is more tightly bound than the initial
system. Again energy would be released in such a process of fusion.
This is the energy source of sun, to be discussed later in Section
13.7.3.
For a separation greater than r0, the force is attractive and for separations
less than r0, the force is strongly repulsive.
The force that determines the motion of atomic electrons is the familiar
Coulomb force. In Section 13.4, we have seen that for average mass
nuclei the binding energy per nucleon is approximately 8 MeV, which
is much larger than the binding energy in atoms. Therefore, to bind a
nucleus together there must be a strong attractive force of a totally
different kind. It must be strong enough to overcome the repulsion
between the (positively charged) protons and to bind both protons and
neutrons into the tiny nuclear volume. We have already seen that the
constancy of binding energy per nucleon can be understood in terms
of its short-range. Many features of the nuclear binding force are
summarised below. These are obtained from a variety of experiments
carried out during 1930 to 1950.
(i) The nuclear force is much stronger than the Coulomb force acting
between charges or the gravitational forces between masses. The
nuclear binding force has to dominate over the Coulomb repulsive
force between protons inside the nucleus. This happens only because
the nuclear force is much stronger than the coulomb force. The
gravitational force is much weaker than even Coulomb force.
(ii) The nuclear force between two nucleons falls rapidly to zero as
their distance is more than a few femtometres. This leads to saturation
of forces in a medium or a large-sized nucleus, which is the reason for
the constancy of the binding energy per nucleon.
A rough plot of the potential energy between two nucleons as a
function of distance is shown in the Fig. 13.2. The potential energy is a
minimum at a distance r0 of about 0.8 fm. This means that the force is
attractive for distances larger than 0.8 fm and repulsive if they are
separated by distances less than 0.8 fm.
(iii) The nuclear force between neutron-neutron, proton-neutron and
proton-proton is approximately the same. The nuclear force does not
depend on the electric charge.
Unlike Coulombs law or the Newtons law of gravitation there is no
simple mathematical form of the nuclear force.
13.6 RADIOACTIVITY
(13.11)
or, ln N ln N0 = (t t0) (13.12)
ln (13.13)
which gives
N(t) = N0 e t (13.14)
Note, for example, the light bulbs follow no such exponential decay
law. If we test 1000 bulbs for their life (time span before they burn out
or fuse), we expect that they will decay (that is, burn out) at more or
less the same time. The decay of radionuclides follows quite a
different law, the law of radioactive decay represented by Eq. (13.14).
The total decay rate R of a sample is the number of nuclei
disintegrating per unit time. Suppose in a time interval dt, the decay
count measured is N. Then dN = N.
The positive quantity R is then defined as
R=
R = N0 e t
or, R = R0 e t (13.15)
T1/2 = = (13.17)
Clearly if N0 reduces to half its value in time T1/2, R0 will also reduce
to half its value in the same time according to Eq. (13.16).
Another related measure is the average or mean life . This again can
be obtained from Eq. (13.14). The number of nuclei which decay in the
time interval t to t + t is R(t)t (= N0ett). Each of them has lived
for time t. Thus the total life of all these nuclei would be t N0et t. It
is clear that some nuclei may live for a short time while others may live
longer. Therefore to obtain the mean life, we have to sum (or
integrate) this expression over all times from 0 to , and divide by the
total number N0 of nuclei at t = 0. Thus,
T1/2 = = ln 2 (13.18)
Radioactive elements (e.g., tritium, plutonium) which are short-lived
i.e., have half-lives much less than the age of the universe ( 15
billion years) have obviously decayed long ago and are not found in
nature. They can, however, be produced artificially in nuclear
reactions.
= 1.42 1017s
One k mol of any isotope contains Avogadros number of atoms, and so 1g of
contains
= =
= 1.23 104 s1
= 1.23 104 Bq
Example 13.5 Tritium has a half-life of 12.5 y undergoing beta decay. What
fraction of a sample of pure tritium will remain undecayed
after 25 y.
Solution
By definition of half-life, half of the initial sample will remain undecayed after
12.5 y. In the next 12.5 y, one-half of these nuclei would have decayed.
Hence, one fourth of the sample of the initial pure tritium will remain
undecayed.
+ (-decay) (13.19)
In -decay, the mass number of the product nucleus (daughter nucleus) is four
less than that of the decaying nucleus (parent nucleus), while the atomic
+ (13.20)
From Einsteins mass-energy equivalance relation [Eq. (13.6)] and energy
conservation, it is clear that this spontaneous decay is possible only when the
total mass of the decay products is less than the mass of the initial nucleus.
This difference in mass appears as kinetic energy of the products. By referring
to a table of nuclear masses, one can check that the total mass of
= 238.05079 u = 4.00260 u
= 234.04363 u = 1.00783 u
= 237.05121 u
Here the symbol Pa is for the element protactinium (Z = 91).
(a) The alpha decay of is given by Eq. (13.20). The energy released
in this process is given by
= (0.00456 u) c2
= (0.00456 u) (931.5 MeV/u)
= 4.25 MeV.
+
The Q for this process to happen is
= ( 0.00825 u) c2
= (0.00825 u)(931.5 MeV/u)
= 7.68 MeV
Thus, the Q of the process is negative and therefore it cannot proceed
(13.22)
(13.23)
The decays are governed by the Eqs. (13.14) and (13.15), so that one can
never predict which nucleus will undergo decay, but one can characterize the
decay by a half-life T1/2 . For example, T1/2 for the decays above is
respectively 14.3 d and 2.6y. The emission of electron in decay is
accompanied by the emission of an antineutrino ( ); in + decay, instead, a
neutrino () is generated. Neutrinos are neutral particles with very small
(possiblly, even zero) mass compared to electrons. They have only weak
interaction with other particles. They are, therefore, very difficult to detect,
since they can penetrate large quantity of matter (even earth) without any
interaction.
In both and + decay, the mass number A remains unchanged. In decay,
the atomic number Z of the nucleus goes up by 1, while in + decay Z goes
down by 1. The basic nuclear process underlying decay is the conversion of
neutron to proton
n p + e + (13.24)
Like an atom, a nucleus also has discrete energy levels - the ground state and
excited states. The scale of energy is, however, very different. Atomic energy
level spacings are of the order of eV, while the difference in nuclear energy
levels is of the order of MeV. When a nucleus in an excited state
spontaneously decays to its ground state (or to a lower energy state), a photon
is emitted with energy equal to the difference in the two energy levels of the
nucleus. This is the so-called gamma decay. The energy (MeV) corresponds to
radiation of extremely short wavelength, shorter than the hard X-ray region.
Figure 13.4 -decay of nucleus followed by emission of two rays from
deexcitation of the daughter
nucleus .
MeV and 1.33 MeV from the deexcitation of nuclei formed from
decay of .
13.7.1 Fission
New possibilities emerge when we go beyond natural radioactive decays and
study nuclear reactions by bombarding nuclei with other nuclear particles such
as proton, neutron, -particle, etc.
A most important neutron-induced nuclear reaction is fission. An example of
(13.26)
The same reaction can produce other pairs of intermediate mass fragments
(13.27)
Or, as another example,
(13.28)
Notice one fact of great importance in the fission reactions given in Eqs.
(13.26) to (13.28). There is a release of extra neutron (s) in the fission process.
Averagely, 2 neutrons are released per fission of uranium nucleus. It is a
fraction since in some fission events 2 neutrons are produced, in some 3, etc.
The extra neutrons in turn can initiate fission processes, producing still more
neutrons, and so on. This leads to the possibility of a chain reaction, as was
first suggested by Enrico Fermi. If the chain reaction is controlled suitably, we
can get a steady energy output. This is what happens in a nuclear reactor. If
the chain reaction is uncontrolled, it leads to explosive energy output, as in a
nuclear bomb.
The atomic energy programme in India was launched around the time of
independence under the leadership of Homi J. Bhabha (1909-1966). An
early historic achievement was the design and construction of the first
nuclear reactor in India (named Apsara) which went critical on August 4,
1956. It used enriched uranium as fuel and water as moderator. Following
this was another notable landmark: the construction of CIRUS (Canada
India Research U.S.) reactor in 1960. This 40 MW reactor used natural
uranium as fuel and heavy water as moderator. Apsara and CIRUS
spurred research in a wide range of areas of basic and applied nuclear
science. An important milestone in the first two decades of the programme
was the indigenous design and construction of the plutonium plant at
Trombay, which ushered in the technology of fuel reprocessing (separating
useful fissile and fertile nuclear materials from the spent fuel of a reactor)
in India. Research reactors that have been subsequently commissioned
include ZERLINA, PURNIMA (I, II and III), DHRUVA and KAMINI. KAMINI
is the countrys first large research reactor that uses U-233 as fuel. As the
name suggests, the primary objective of a research reactor is not
generation of power but to provide a facility for research on different
aspects of nuclear science and technology. Research reactors are also an
excellent source for production of a variety of radioactive isotopes that find
application in diverse fields: industry, medicine and agriculture.
The main objectives of the Indian Atomic Energy programme are to
provide safe and reliable electric power for the countrys social and
economic progress and to be self-reliant in all aspects of nuclear
technology. Exploration of atomic minerals in India undertaken since the
early fifties has indicated that India has limited reserves of uranium, but
fairly abundant reserves of thorium. Accordingly, our country has adopted
a three-stage strategy of nuclear power generation. The first stage involves
the use of natural uranium as a fuel, with heavy water as moderator. The
Plutonium-239 obtained from reprocessing of the discharged fuel from the
reactors then serves as a fuel for the second stage the fast breeder
reactors. They are so called because they use fast neutrons for sustaining
the chain reaction (hence no moderator is needed) and, besides
generating power, also breed more fissile species (plutonium) than they
consume. The third stage, most significant in the long term, involves using
fast breeder reactors to produce fissile Uranium-233 from Thorium-232
and to build power reactors based on them.
India is currently well into the second stage of the programme and
considerable work has also been done on the third the thorium
utilisation stage. The country has mastered the complex technologies of
mineral exploration and mining, fuel fabrication, heavy water production,
reactor design, construction and operation, fuel reprocessing, etc.
Pressurised Heavy Water Reactors (PHWRs) built at different sites in the
country mark the accomplishment of the first stage of the programme.
India is now more than self-sufficient in heavy water production. Elaborate
safety measures both in the design and operation of reactors, as also
adhering to stringent standards of radiological protection are the hallmark
of the Indian Atomic Energy Programme.
(13.29)
Plutonium undergoes fission with slow neutrons.
Figure 13.5 shows the schematic diagram of a nuclear reactor based on
thermal neutron fission. The core of the reactor is the site of nuclear fission. It
contains the fuel elements in suitably fabricated form. The fuel may be say
Figure 13.5 Schematic diagram of a nuclear reactor based on thermal neutron fission.
When two light nuclei fuse to form a larger nucleus, energy is released, since
the larger nucleus is more tightly bound, as seen from the binding energy
curve in Fig.13.1. Some examples of such energy liberating nuclear fusion
reactions are :
or
(13.31)
SUMMARY
1. An atom has a nucleus. The nucleus is positively charged. The radius of the
nucleus is smaller than the radius of an atom by a factor of 104. More than
99.9% mass of the atom is concentrated in the nucleus.
2. On the atomic scale, mass is measured in atomic mass units (u). By
definition, 1 atomic mass unit (1u) is 1/12th mass of one atom of 12C;
1u = 1.660563 1027 kg.
3. A nucleus contains a neutral particle called neutron. Its mass is almost the
same as that of proton
4. The atomic number Z is the number of protons in the atomic nucleus of an
element. The mass number A is the total number of protons and neutrons in
the atomic nucleus; A = Z+N; Here N denotes the number of neutrons in the
nucleus.
R = R0 A1/3,
where R0 = a constant = 1.2 fm. This implies that the nuclear density is
independent of A. It is of the order of 1017 kg/m3.
6. Neutrons and protons are bound in a nucleus by the short-range strong
nuclear force. The nuclear force does not distinguish between neutron and
proton.
7. The nuclear mass M is always less than the total mass, m, of its
constituents. The difference in mass of a nucleus and its constituents is called
the mass defect,
=(
M Z mp + (A Z)mn) M
Using Einsteins mass energy relation, we express this mass difference in
terms of energy as
b=
E M c2
The energy Eb represents the binding energy of the nucleus. In the mass
number range A = 30 to 170, the binding energy per nucleon is nearly
constant, about 8 MeV/nucleon.
8. Energies associated with nuclear processes are about a million times larger
than chemical process.
9. The Q-value of a nuclear process is
12. Energy is released when less tightly bound nuclei are transmuted into
more tightly bound nuclei. In fission, a heavy nucleus like breaks into
Exercises
You may find the following data useful in solving the exercises:
e = 1.61019C N = 6.0231023 per mole
mH = 1.007825 u mn = 1.008665 u
m( ) = 4.002603 u me = 0.000548 u
, given m =14.00307 u
m( ) = 55.934939 u m ( ) = 208.980388 u
13.5 A given coin has a mass of 3.0 g. Calculate the nuclear
energy that would be required to separate all the neutrons and
protons from each other. For simplicity assume that the coin is
m( ) = 220.01137 u, m ( ) = 216.00189 u.
13.13 The radionuclide 11C decays according to
m( ) = 22.994466 u
m( ) = 22.089770 u.
13.15 The Q value of a nuclear reaction A + b C + d is defined
by
Q = [ mA + mb mC md]c2
where the masses refer to the respective nuclei. Determine from
the given data the Q-value of the following reactions and state
whether the reactions are exothermic or endothermic.
(i)
(ii)
Atomic masses are given to be
m( ) = 2.014102 u
m( ) = 3.016049 u
m( ) = 12.000000 u
m( ) = 19.992439 u
m( ) = 27.98191 u.
undergo fission?
13.18 A 1000 MW fission reactor consumes half of its fuel in 5.00
Additional Exercises
13.23 In a periodic table the average atomic mass of magnesium
is given as 24.312 u. The average value is based on their relative
natural abundance on earth. The three isotopes and their masses
m( ) = 39.962591 u
m( ) = 40.962278 u
m( ) = 25.986895 u
m( ) = 26.981541 u
come from . How long one must wait until 90% do so?
13.26 Under certain circumstances, a nucleus can decay by
emitting a particle more massive than an -particle. Consider the
following decay processes:
Calculate the Q-values for these decays and determine that both
are energetically allowed.
m( ) =238.05079 u
m( ) =139.90543 u
m( ) = 98.90594 u
(a) Calculate the energy released in MeV in this reaction from the
data:
m( )=2.014102 u
m( ) =3.016049 u
(b) Consider the radius of both deuterium and tritium to be
approximately 2.0 fm. What is the kinetic energy needed to
overcome the coulomb repulsion between the two nuclei? To what
temperature must the gas be heated to initiate the reaction?
(Hint: Kinetic energy required for one fusion event =average
thermal kinetic energy available with the interacting particles =
2(3kT/2); k = Boltzmans constant, T = absolute temperature.)
Figure13.6
13.30 Calculate and compare the energy released by a) fusion of
1.0 kg of hydrogen deep within Sun and b) the fission of 1.0 kg of
235U in a fission reactor.
13.31 Suppose India had a target of producing by 2020 AD,
200,000 MW of electric power, ten percent of which was to be
obtained from nuclear power plants. Suppose we are given that,
on an average, the efficiency of utilization (i.e. conversion to
electric energy) of thermal energy produced in a reactor was 25%.
How much amount of fissionable uranium would our country need
per year by 2020? Take the heat energy per fission of 235U to be
about 200MeV.
Chapter Eleven
11.1 INTRODUCTION
(ii) Field emission: By applying a very strong electric field (of the order
of 108 V m1) to a metal, electrons can be pulled out of the metal, as in
a spark plug.
(iii) Photo-electric emission: When light of suitable frequency
illuminates a metal surface, electrons are emitted from the metal
surface. These photo(light)-generated electrons are called
photoelectrons.
Figure 11.4 Variation of photoelectric current with collector plate potential for
different frequencies of incident radiation.
Albert Einstein (1879 1955) Einstein, one of the greatest physicists of all
time, was born in Ulm, Germany. In 1905, he published three path-breaking
papers. In the first paper, he introduced the notion of light quanta (now called
photons) and used it to explain the features of photoelectric effect. In the
second paper, he developed a theory of Brownian motion, confirmed
experimentally a few years later and provided a convincing evidence of the
atomic picture of matter. The third paper gave birth to the special theory of
relativity. In 1916, he published the general theory of relativity. Some of
Einsteins most significant later contributions are: the notion of stimulated
emission introduced in an alternative derivation of Plancks blackbody radiation
law, static model of the universe which started modern cosmology, quantum
statistics of a gas of massive bosons, and a critical analysis of the foundations
of quantum mechanics. In 1921, he was awarded the Nobel Prize in physics
for his contribution to theoretical physics and the photoelectric effect.
According to Eq. (11.2), Kmax depends linearly on , and is
independent of intensity of radiation, in agreement with
observation. This has happened because in Einsteins picture,
photoelectric effect arises from the absorption of a single
quantum of radiation by a single electron. The intensity of
radiation (that is proportional to the number of energy quanta per
unit area per unit time) is irrelevant to this basic process.
Since Kmax must be non-negative, Eq. (11.2 ) implies that
photoelectric emission is possible only if
h > 0
or > 0 , where
0 = (11.3)
Equation (11.3) shows that the greater the work function 0, the
higher the minimum or threshold frequency 0 needed to emit
photoelectrons. Thus, there exists a threshold frequency0 (=
0/h) for the metal surface, below which no photoelectric
emission is possible, no matter how intense the incident radiation
may be or how long it falls on the surface.
In this picture, intensity of radiation as noted above, is
proportional to the number of energy quanta per unit area per
unit time. The greater the number of energy quanta available, the
greater is the number of electrons absorbing the energy quanta
and greater, therefore, is the number of electrons coming out of
the metal (for > 0). This explains why, for > 0 , photoelectric
current is proportional to intensity.
In Einsteins picture, the basic elementary process involved in
photoelectric effect is the absorption of a light quantum by an
electron. This process is instantaneous. Thus, whatever may be
the intensity i.e., the number of quanta of radiation per unit area
per unit time, photoelectric emission is instantaneous. Low
intensity does not mean delay in emission, since the basic
elementary process is the same. Intensity only determines how
many electrons are able to participate in the elementary process
(absorption of a light quantum by a single electron) and,
therefore, the photoelectric current.
Using Eq. (11.1), the photoelectric equation, Eq. (11.2), can be
written as
e V0 = h 0; for
or V0 = (11.4)
This is an important result. It predicts that the V0 versus curve is a
straight line with slope = (h/e), independent of the nature of the
material. During 1906-1916, Millikan performed a series of
experiments on photoelectric effect, aimed at disproving Einsteins
photoelectric equation. He measured the slope of the straight line
obtained for sodium, similar to that shown in Fig. 11.5. Using the
known value of e, he determined the value of Plancks constant h.
34
This value was close to the value of Plancks contant (= 6.626 10 J
s) determined in an entirely different context. In this way, in 1916,
Millikan proved the validity of Einsteins photoelectric equation, instead
of disproving it.
The successful explanation of photoelectric effect using the
hypothesis of light quanta and the experimental determination of
values of h and 0, in agreement with values obtained from other
experiments, led to the acceptance of Einsteins picture of
photoelectric effect. Millikan verified photoelectric equation with great
precision, for a number of alkali metals over a wide range of radiation
frequencies.
= 3.98 1019 J
(b) If N is the number of photons emitted by the source per second, the power
P transmitted in the beam equals N times the energy per photon E, so that P =
N E. Then
N=
Example 11.2 The work function of caesium is 2.14 eV. Find (a) the threshold
frequency for caesium, and (b) the wavelength of the incident light if the
photocurrent is brought to zero by a stopping potential of 0.60 V.
Solution
(a) For the cut-off or threshold frequency, the energy h 0 of the incident
radiation must be equal to work function 0, so that
0 =
Thus, for frequencies less than this threshold frequency, no photoelectrons are
ejected.
(b) Photocurrent reduces to zero, when maximum kinetic energy of the emitted
photoelectrons equals the potential energy e V0 by the retarding potential V0.
Einsteins Photoelectric equation is
eV0 = h 0 = 0
or, = hc/(eV0 + 0)
Example 11.3 The wavelength of light in the visible region is about 390 nm for
violet colour, about 550 nm (average wavelength) for yellow-green colour and
about 760 nm for red colour.
(a) What are the energies of photons in (eV) at the (i) violet end, (ii) average
wavelength, yellow-green colour, and (iii) red end of the visible spectrum?
(Take h = 6.631034 J s and 1 eV = 1.610 19J.)
(b) From which of the photosensitive materials with work functions listed in
Table 11.1 and using the results of (i), (ii) and (iii) of (a), can you build a
photoelectric device that operates with visible light?
Solution
(a) Energy of the incident photon, E = h = hc/
E = (6.631034J s) (3108 m/s)/
= 5.10 1019J
= 3.19 eV
(ii) For yellow-green light, 2 = 550 nm (average wavelength)
= 3.621019 J = 2.26 eV
(iii) For red light, 3 = 760 nm (higher wavelength end)
= 2.621019 J = 1.64 eV
(b) For a photoelectric device to operate, we require incident light energy E to
be equal to or greater than the work function 0 of the material. Thus, the
photoelectric device will operate with violet light (with E = 3.19 eV)
photosensitive material Na (with 0 = 2.75 eV), K (with 0 = 2.30 eV) and Cs
(with 0 = 2.14 eV). It will also operate with yellow-green light (with E = 2.26
eV) for Cs (with 0 = 2.14 eV) only. However, it will not operate with red light
(with E = 1.64 eV) for any of these photosensitive materials.
= (11.5)
where m is the mass of the particle and v its speed. Equation (11.5) is
known as the de Broglie relation and the wavelength of the matter
wave is called de Broglie wavelength. The dual aspect of matter is
evident in the de Broglie relation. On the left hand side of Eq. (11.5),
is the attribute of a wave while on the right hand side the momentum p
is a typical attribute of a particle. Plancks constant h relates the two
attributes.
Equation (11.5) for a material particle is basically a hypothesis whose
validity can be tested only by experiment. However, it is interesting to
see that it is satisfied also by a photon. For a photon, as we have
seen,
p = h /c (11.6)
Therefore,
(11.7)
= = = 2.76 1034 m
Photocell
A photo cell
When light of suitable wavelength falls on the emitter C, photoelectrons are
emitted. These photoelectrons are drawn to the collector A. Photocurrent of
the order of a few microampere can be normally obtained from a photo cell.
A photocell converts a change in intensity of illumination into a change in
photocurrent. This current can be used to operate control systems and in light
measuring devices. A photocell of lead sulphide sensitive to infrared radiation
is used in electronic ignition circuits.
In scientific work, photo cells are used whenever it is necessary to measure
the intensity of light. Light meters in photographic cameras make use of photo
cells to measure the intensity of incident light. The photocells, inserted in the
door light electric circuit, are used as automatic door opener. A person
approaching a doorway may interrupt a light beam which is incident on a
photocell. The abrupt change in photocurrent may be used to start a motor
which opens the door or rings an alarm. They are used in the control of a
counting device which records every interruption of the light beam caused by a
person or object passing across the beam. So photocells help count the
persons entering an auditorium, provided they enter the hall one by one. They
are used for detection of traffic law defaulters: an alarm may be sounded
whenever a beam of (invisible) radiation is intercepted.
In burglar alarm, (invisible) ultraviolet light is continuously made to fall on a
photocell installed at the doorway. A person entering the door interrupts the
beam falling on the photocell. The abrupt change in photocurrent is used to
start an electric bell ringing. In fire alarm, a number of photocells are installed
at suitable places in a building. In the event of breaking out of fire, light
radiations fall upon the photocell. This completes the electric circuit through an
electric bell or a siren which starts operating as a warning signal.
Photocells are used in the reproduction of sound in motion pictures and in the
television camera for scanning and telecasting scenes. They are used in
industries for detecting minor flaws or holes in metal sheets.
of the electron equals the work done (eV ) on it by the electric field:
K=eV (11.8)
Now , K = m v2 = , so that
p= (11.9)
The de Broglie wavelength of the electron is then
= (11.10)
Substituting the numerical values of h, m, e,
we get
(11.11)
Louis Victor de Broglie (1892 1987) French physicist who put forth
revolutionary idea of wave nature of matter. This idea was developed by Erwin
Schrdinger into a full-fledged theory of quantum mechanics commonly known
as wave mechanics. In 1929, he was awarded the Nobel Prize in Physics for
his discovery of the wave nature of electrons.
Figure 11.6 (a) The wave packet description of an electron. The wave packet
corresponds to a spread of wavelength around some central wavelength (and
hence by de Broglie relation, a spread in momentum). Consequently, it is
associated with an uncertainty in position (x) and an uncertainty in momentum
(p).(b) The matter wave corresponding to a definite momentum of an
electron extends all over space. In this case, p = 0 and x .
Solution
(a) For the electron:
Mass m = 9.111031 kg, speed v = 5.4106 m/s. Then, momentum p = m v =
9.111031 (kg) 5.4 106 (m/s)
p = 4.92 1024 kg m/s
= 0.135 nm
(b) For the ball:
Mass m = 0.150 kg, speed v = 30.0 m/s.
Then momentum p = m v = 0.150 (kg) 30.0 (m/s)
p= 4.50 kg m/s
= 1.47 1034 m
The de Broglie wavelength of electron is comparable with X-ray wavelengths.
19
However, for the ball it is about 10 times the size of the proton, quite
beyond experimental measurement.
Example 11.5 An electron, an -particle, and a proton have the same kinetic
energy. Which of these particles has the shortest de Broglie wavelength?
Solution
For a particle, de Broglie wavelength, = h/p
2
Kinetic energy, K = p /2m
Then,
For the same kinetic energy K, the de Broglie wavelength associated with the
particle is inversely proportional to the square root of their masses. A proton
It is worth pausing here to reflect on just what a matter wave associated with a
particle, say, an electron, means. Actually, a truly satisfactory physical
understanding of the dual nature of matter and radiation has not emerged so
far. The great founders of quantum mechanics (Niels Bohr, Albert Einstein,
and many others) struggled with this and related concepts for long. Still the
deep physical interpretation of quantum mechanics continues to be an area of
active research. Despite this, the concept of matter wave has been
mathematically introduced in modern quantum mechanics with great success.
An important milestone in this connection was when Max Born (1882-1970)
suggested a probability interpretation to the matter wave amplitude. According
to this, the intensity (square of the amplitude) of the matter wave at a point
determines the probability density of the particle at that point. Probability
density means probability per unit volume. Thus, if A is the amplitude of the
wave at a point, |A|2 V is the probability of the particle being found in a small
volume V around that point. Thus, if the intensity of matter wave is large in a
certain region, there is a greater probability of the particle being found there
than where the intensity is small.
= h /p nm
nm = 0.123 nm
= h /p nm
nm = 0.167 nm
Thus, there is an excellent agreement between the theoretical value
and the experimentally obtained value of de Broglie wavelength.
Davisson-Germer experiment thus strikingly confirms the wave nature
of electrons and the de Broglie relation. More recently, in 1989, the
wave nature of a beam of electrons was experimentally demonstrated
in a double-slit experiment, similar to that used for the wave nature of
light. Also, in an experiment in 1994, interference fringes were
obtained with the beams of iodine molecules, which are about a
million times more massive than electrons.
The de Broglie hypothesis has been basic to the development of
modern quantum mechanics. It has also led to the field of electron
optics. The wave properties of electrons have been utilised in the
design of electron microscope which is a great improvement, with
higher resolution, over the optical microscope.
Summary
m v2max = V0 e = h 0 = h ( 0)
This photoelectric equation explains all the features of the photoelectric effect.
Millikans first precise measurements confirmed the Einsteins photoelectric
equation and obtained an accurate value of Plancks constant h. This led to the
acceptance of particle or photon description (nature) of electromagnetic
radiation, introduced by Einstein.
8. Radiation has dual nature: wave and particle. The nature of experiment
determines whether a wave or particle description is best suited for
understanding the experimental result. Reasoning that radiation and matter
should be symmetrical in nature, Louis Victor de Broglie attributed a wave-like
character to matter (material particles). The waves associated with the moving
material particles are called matter waves or de Broglie waves.
9. The de Broglie wavelength () associated with a moving particle is related to
its momentum p as: = h/p. The dualism of matter is inherent in the de Broglie
relation which contains a wave concept () and a particle concept (p). The de
Broglie wavelength is independent of the charge and nature of the material
particle. It is significantly measurable (of the order of the atomic-planes
spacing in crystals) only in case of sub-atomic particles like electrons, protons,
etc. (due to smallness of their masses and hence, momenta). However, it is
indeed very small, quite beyond measurement, in case of macroscopic objects,
commonly encountered in everyday life.
10. Electron diffraction experiments by Davisson and Germer, and by G. P.
Thomson, as well as many later experiments, have verified and confirmed the
wave-nature of electrons. The de Broglie hypothesis of matter waves supports
the Bohrs concept of stationary orbits.
Points to Ponder
1. Free electrons in a metal are free in the sense that they move inside the
metal in a constant potential (This is only an approximation). They are not free
to move out of the metal. They need additional energy to get out of the metal.
2. Free electrons in a metal do not all have the same energy. Like molecules in
a gas jar, the electrons have a certain energy distribution at a given
temperature. This distribution is different from the usual Maxwells distribution
that you have learnt in the study of kinetic theory of gases. You will learn about
it in later courses, but the difference has to do with the fact that electrons obey
Paulis exclusion principle.
3. Because of the energy distribution of free electrons in a metal, the energy
required by an electron to come out of the metal is different for different
electrons. Electrons with higher energy require less additional energy to come
out of the metal than those with lower energies. Work function is the least
energy required by an electron to come out of the metal.
4. Observations on photoelectric effect imply that in the event of matter-light
interaction, absorption of energy takes place in discrete units of h. This is not
quite the same as saying that light consists of particles, each of energy h.
5. Observations on the stopping potential (its independence of intensity and
dependence on frequency) are the crucial discriminator between the wave-
picture and photon-picture of photoelectric effect.
Exercises
(a) Find the energy and momentum of each photon in the light
beam,
(b) How many photons per second, on the average, arrive at a
target irradiated by this beam? (Assume the beam to have uniform
cross-section which is less than the target area), and
(c) How fast does a hydrogen atom have to travel in order to have
the same momentum as that of the photon?
11.5 The energy flux of sunlight reaching the surface of the earth
is
1.388 103 W/m2. How many photons (nearly) per square metre
are incident on the Earth per second? Assume that the photons in
the sunlight have an average wavelength of 550 nm.
11.6 In an experiment on photoelectric effect, the slope of the cut-
off voltage versus frequency of incident light is found to be 4.12
1015 V s. Calculate the value of Plancks constant.
11.7 A 100W sodium lamp radiates energy uniformly in all
directions. The lamp is located at the centre of a large sphere that
absorbs all the sodium light which is incident on it. The wavelength
of the sodium light is 589 nm. (a) What is the energy per photon
associated with the sodium light? (b) At what rate are the photons
delivered to the sphere?
11.8 The threshold frequency for a certain metal is 3.3 1014 Hz.
If light of frequency 8.2 1014 Hz is incident on the metal, predict
the cut-off voltage for the photoelectric emission.
11.9 The work function for a certain metal is 4.2 eV. Will this metal
give photoelectric emission for incident radiation of wavelength
330 nm?
11.10 Light of frequency 7.21 1014 Hz is incident on a metal
surface. Electrons with a maximum speed of 6.0 105 m/s are
ejected from the surface. What is the threshold frequency for
photoemission of electrons?
11.11 Light of wavelength 488 nm is produced by an argon laser
which is used in the photoelectric effect. When light from this
spectral line is incident on the emitter, the stopping (cut-off)
potential of photoelectrons is 0.38 V. Find the work function of the
material from which the emitter is made.
11.12 Calculate the
(a) momentum,
(b) speed, and
(c) de Broglie wavelength of an electron with kinetic energy of
120 eV.
11.14 The wavelength of light from the spectral emission line of
sodium is 589 nm. Find the kinetic energy at which
(a) an electron, and
11.20 (a) Estimate the speed with which electrons emitted from a
heated emitter of an evacuated tube impinge on the collector
maintained at a potential difference of 500 V with respect to the
emitter. Ignore the small initial speeds of the electrons.
The specific charge of the electron, i.e., its e/m is given to be 1.76
1011 C kg1.
(b) Use the same formula you employ in (a) to obtain electron
speed for an collector potential of 10 MV. Do you see what is
wrong ? In what way is the formula to be modified?
11.21 (a) A monoenergetic electron beam with electron speed
[Note: You will notice that to get h from the data, you will need to
know e (which you can take to be 1.6 1019 C). Experiments of
this kind on Na, Li, K, etc. were performed by Millikan, who, using
his own value of e (from the oil-drop experiment) confirmed
Einsteins photoelectric equation and at the same time gave an
independent estimate of the value of h.]
11.29 The work function for the following metals is given:
Na: 2.75 eV; K: 2.30 eV; Mo: 4.17 eV; Ni: 5.15 eV. Which of these
metals will not give photoelectric emission for a radiation of
wavelength 3300 from a He-Cd laser placed 1 m away from the
photocell? What happens if the laser is brought nearer and placed
50 cm away?
11.30 Light of intensity 105 W m2 falls on a sodium photo-cell of
surface area 2 cm2. Assuming that the top 5 layers of sodium
absorb the incident energy, estimate time required for
photoelectric emission in the wave-picture of radiation. The work
function for the metal is given to be about 2 eV. What is the
implication of your answer?
11.31 Crystal diffraction experiments can be performed using X-
rays, or electrons accelerated through appropriate voltage. Which
probe has greater energy? (For quantitative comparison, take the
wavelength of the probe equal to 1 , which is of the order of inter-
atomic spacing in the lattice) (me=9.11 1031 kg).
11.32 (a) Obtain the de Broglie wavelength of a neutron of kinetic
energy 150 eV. As you have seen in Exercise 11.31, an electron
beam of this energy is suitable for crystal diffraction experiments.
Would a neutron beam of the same energy be equally suitable?
Explain. (mn = 1.675 1027 kg)
(b) Obtain the de Broglie wavelength associated with thermal
neutrons at room temperature (27 C). Hence explain why a fast
neutron beam needs to be thermalised with the environment
before it can be used for neutron diffraction experiments.
11.33 An electron microscope uses electrons accelerated by a
voltage of 50 kV. Determine the de Broglie wavelength associated
with the electrons. If other factors (such as numerical aperture,
etc.) are taken to be roughly the same, how does the resolving
power of an electron microscope compare with that of an optical
microscope which uses yellow light?
11.34 The wavelength of a probe is roughly a measure of the size
of a structure that it can probe in some detail. The quark
structure of protons and neutrons appears at the minute length-
scale of
1015 m or less. This structure was first probed in early 1970s
using high energy electron beams produced by a linear
accelerator at Stanford, USA. Guess what might have been the
order of energy of these electron beams. (Rest mass energy of
electron = 0.511 MeV.)
11.35 Find the typical de Broglie wavelength associated with a He
atom in helium gas at room temperature (27 C) and 1 atm
pressure; and compare it with the mean separation between two
atoms under these conditions.
11.36 Compute the typical de Broglie wavelength of an electron in
a metal at 27 C and compare it with the mean separation
between two electrons in a metal which is given to be about 2
1010 m.
[Note: Exercises 11.35 and 11.36 reveal that while the wave-
packets associated with gaseous molecules under ordinary
conditions are non-overlapping, the electron wave-packets in a
metal strongly overlap with one another. This suggests that
whereas molecules in an ordinary gas can be distinguished apart,
electrons in a metal cannot be distintguished apart from one
another. This indistinguishibility has many fundamental
implications which you will explore in more advanced Physics
courses.]
11.37 Answer the following questions:
E = h , p =
Appendix
Objectives
We are mostly surrounded by solids and we use them more often than liquids
and gases. For different applications we need solids with widely different
properties. These properties depend upon the nature of constituent particles
and the binding forces operating between them. Therefore, study of the
structure of solids is important. The correlation between structure and
properties helps in discovering new solid materials with desired properties
like high temperature superconductors, magnetic materials, biodegradable
polymers for packaging, biocompliant solids for surgical implants, etc.
From our earlier studies, we know that liquids and gases are called fluids
because of their ability to flow. The fluidity in both of these states is due to
the fact that the molecules are free to move about. On the contrary, the
constituent particles in solids have fixed positions and can only oscillate
about their mean positions. This explains the rigidity in solids. In crystalline
solids, the constituent particles are arranged in regular patterns.
In Class XI you have learnt that matter can exist in three states namely, solid,
liquid and gas. Under a given set of conditions of temperature and pressure,
which of these would be the most stable state of a given substance depends
upon the net effect of two opposing factors. Intermolecular forces tend to
keep the molecules (or atoms or ions) closer, whereas thermal energy tends to
keep them apart by making them move faster. At sufficiently low
temperature, the thermal energy is low and intermolecular forces bring them
so close that they cling to one another and occupy fixed positions. These can
still oscillate about their mean positions and the substance exists in solid
state. The following are the characteristic properties of the solid state:
Fig. 1.1: Two dimensional structure of (a) quartz and (b) quartz glass
Crystalline solids have a sharp melting point. On the other hand, amorphous
solids soften over a range of temperature and can be moulded and blown into
various shapes. On heating they become crystalline at some temperature.
Some glass objects from ancient civilisations are found to become milky in
appearance because of some crystallisation. Like liquids, amorphous solids
have a tendency to flow, though very slowly. Therefore, sometimes these are
called pseudo solids or super cooled liquids. Glass panes fixed to windows or
doors of old buildings are invariably found to be slightly thicker at the bottom
than at the top. This is because the glass flows down very slowly and makes
the bottom portion slightly thicker.
Crystalline solids are anisotropic in nature, that is, some of their physical
properties like electrical resistance or refractive index show different values
when measured along different directions in the same crystals. This arises
from different arrangement of particles in different directions. This is
illustrated in Fig. 1.2. Since the arrangement of particles is different along
different directions, the value of same physical property is found to be
different along each direction.
Fig. 1.2: Anisotropy in crystals is due to different arrangement of particles along different
directions.
Definite characteristic
Shape Irregular shape
geometrical shape
Order in
Pseudo solids or
arrangement
super cooled liquids
of Long range order
Only short range
constituent
order.
particles
Amorphous solids are useful materials. Glass, rubber and plastics find many
applications in our daily lives. Amorphous silicon is one of the best
photovoltaic material available for conversion of sunlight into electricity.
Intext Questions
1.5 Refractive index of a solid is observed to have the same value along
all directions. Comment on the nature of this solid. Would it show
cleavage property?
1.3 Classification of Crystalline Solids
In Section 1.2, we have learnt about amorphous substances and that they have
only short range order. However, most of the solid substances are crystalline
in nature. For example, all the metallic elements like iron, copper and silver;
non metallic elements like sulphur, phosphorus and iodine and compounds
like sodium chloride, zinc sulphide and naphthalene form crystalline solids.
Molecules are the constituent particles of molecular solids. These are further
sub divided into the following categories:
(i) Non polar Molecular Solids: They comprise of either atoms, for example,
argon and helium or the molecules formed by non polar covalent bonds for
example H2, Cl2 and I2. In these solids, the atoms or molecules are held by
weak dispersion forces or London forces about which you have learnt in
Class XI. These solids are soft and non-conductors of electricity. They have
low melting points and are usually in liquid or gaseous state at room
temperature and pressure.
(ii) Polar Molecular Solids: The molecules of substances like HCl, SO2, etc.
are formed by polar covalent bonds. The molecules in such solids are held
together by relatively stronger dipole-dipole interactions. These solids are
soft and non-conductors of electricity. Their melting points are higher than
those of non polar molecular solids yet most of these are gases or liquids
under room temperature and pressure. Solid SO2 and solid NH3 are some
examples of such solids.
Ions are the constituent particles of ionic solids. Such solids are formed by
the three dimensional arrangements of cations and anions bound by strong
coulombic (electrostatic) forces. These solids are hard and brittle in nature.
They have high melting and boiling points. Since the ions are not free to
move about, they are electrical insulators in the solid state. However, in the
molten state or when dissolved in water, the ions become free to move about
and they conduct electricity.
The different properties of the four types of solids are listed in Table 1.2.
(1)
Molecular
solids
Dispersion or London Ar, CCl4,
forces Soft
(i) Non H2, I2, CO
polar Molecules Dipole-dipole Soft
interactions HCl, SO2
(ii) Polar Hydrogen bonding H2O (ice) Hard
(iii)
Hydrogen
Bonded
NaCl,
(2) Ionic Coulombic or MgO, Hard but
Ions
solids electrostatic Brittle
ZnS, CaF2
Positive
(3) Ions in a
Hard but
Metallic sea of Metalic Binding Fe, Cu, Ag
malleable
solids delocalised
electrons
SiO2
(4)
Covalent (quartz), Hard
or Atoms Covalent bonding SiC, C
network (diamond), Soft
solids AlN, C
graphite)
Intext Questions
1.8 Ionic solids conduct electricity in molten state but not in solid state.
Explain.
1.9 What type of solids are electrical conductors, malleable and ductile?
Fig. 1.5: A portion of a three dimensional cubic lattice and its unit cell.
There are only 14 possible three dimensional lattices. These are called
Bravais Lattices (after the French mathematician who first described them).
The following are the characteristics of a crystal lattice:
(b) Each point in a crystal lattice represents one constituent particle which
may be an atom, a molecule (group of atoms) or an ion.
(c) Lattice points are joined by straight lines to bring out the geometry of the
lattice. Unit cell is the smallest portion of a crystal lattice which, when
repeated in different directions, generates the entire lattice.
(i) its dimensions along the three edges, a, b and c. These edges may or may
not be mutually perpendicular.
(ii) angles between the edges, (between b and c) (between a and c) and
(between a and b). Thus, a unit cell is characterised by six parameters, a, b, c,
, and . These parameters of a typical unit cell are shown in Fig. 1.6.
When constituent particles are present only on the corner positions of a unit
cell, it is called as primitive unit cell.
(i) Body-Centred Unit Cells: Such a unit cell contains one constituent particle
(atom, molecule or ion) at its body-centre besides the ones that are at its
corners.
(ii) Face-Centred Unit Cells: Such a unit cell contains one constituent
particle present at the centre of each face, besides the ones that are at its
corners.
(iii) End-Centred Unit Cells: In such a unit cell, one constituent particle is
present at the centre of any two opposite faces besides the ones present at its
corners.
In all, there are seven types of primitive unit cells (Fig. 1.7).
Fig. 1.7: Seven primitive unit cells in crystals
Their characteristics along with the centred unit cells they can form have
been listed in Table 1.3.
Table 1.3: Seven Primitive Unit Cells and their Possible Variations as
Centred Unit Cells
Axial
Crystal Possible distances Axial
Examples
system Variations or edge angles
lengths
Primitive,
Body- =
NaCl, Zinc
Cubic centred, a=b=c = =
blende, Cu
Face- 90
centred
=
=90 Graphite,
Hexagonal Primitive a=bc
= ZnO,CdS,
120
= Calcite
Rhombohedral (CaCO3), HgS
Primitive a=b=c =
or Trigonal
90 (cinnabar)
= Monoclinic
Primitive,
=90 sulphur,
Monoclinic End- a b c
Na2SO4.10H2O
centred
90
K2Cr2O7,
Triclinic Primitive abc CuSO4. 5H2O,
90 H3BO3
We know that any crystal lattice is made up of a very large number of unit
cells and every lattice point is occupied by one constituent particle (atom,
molecule or ion). Let us now work out what portion of each particle belongs
to a particular unit cell.
We shall consider three types of cubic unit cells and for simplicity assume
that the constituent particle is an atom.
Primitive cubic unit cell has atoms only at its corner. Each atom at a corner is
shared between eight adjacent unit cells as shown in
Fig. 1.8, four unit cells in the same layer and four unit cells of the upper (or
particular unit cell. In Fig. 1.9, a primitive cubic unit cell has been depicted in
three different ways. Each small sphere in Fig. 1.9 (a) represents only the
centre of the particle occupying that position and not its actual size. Such
structures are called open structures. The arrangement of particles is easier to
follow in open structures.
Fig. 1.9 (b) depicts space-filling representation of the unit cell with actual
particle size and Fig. 1.9 (c) shows the actual portions of different atoms
present in a cubic unit cell.
In all, since each cubic unit cell has
8 atoms on its corners, the total number of atoms in one unit cell is
atom.
Fig. 1.8: In a simple cubic unit cell, each corner atom is shared between 8 unit cells.
Fig. 1.9: A primitive cubic unit cell (a) open structure (b) space-filling structure (c) actual
portions of atoms belonging to one unit cell.
1.5.2 Body-Centred Cubic Unit Cell
A body-centred cubic (bcc) unit cell has an atom at each of its corners and
also one atom at its body centre. Fig. 1.10 depicts (a) open structure (b) space
filling model and (c) the unit cell with portions of atoms actually belonging to
it. It can be seen that the atom at the body centre wholly belongs to the unit
cell in which it is present. Thus in a body-centered cubic (bcc) unit cell:
Fig. 1.10: A body-centred cubic unit cell (a) open structure (b) spacefilling structure (c)
actual portions of atoms belonging to one unit cell.
A face-centred cubic (fcc) unit cell contains atoms at all the corners and at the
centre of all the faces of the cube. It can be seen in Fig. 1.11 that each atom
located at the face-centre is shared between two adjacent unit cells and only
of each atom belongs to a unit cell. Fig. 1.12 depicts (a) open structure (b)
space-filling model and (c) the unit cell with portions of atoms actually
belonging to it. Thus, in a face-centred cubic (fcc) unit cell:
Fig. 1.11: An atom at face centre of unit cell is shared between 2 unit cells
Fig 1.12: A face-centred cubic unit cell (a) open structure (b) space filling structure (c)
actual portions of atoms belonging to one unit cell.
Intext Questions
1.13 Explain how much portion of an atom located at (i) corner and (ii)
bodycentre of a cubic unit cell is part of its neighbouring unit cell.
In this arrangement, each sphere is in contact with two of its neighbours. The
number of nearest neighbours of a particle is called its coordination number.
Thus, in one dimensional close packed arrangement, the coordination number
is 2.
(i) The second row may be placed in contact with the first one such that the
spheres of the second row are exactly above those of the first row. The
spheres of the two rows are aligned horizontally as well as vertically. If we
call the first row as A type row, the second row being exactly the same as
the first one, is also of A type. Similarly, we may place more rows to obtain
AAA type of arrangement as shown in Fig. 1.14 (a).
Fig. 1.14: (a) Square close packing (b) hexagonal close packing of spheres in two
dimensions
(ii) The second row may be placed above the first one in a staggered manner
such that its spheres fit in the depressions of the first row. If the arrangement
of spheres in the first row is called A type, the one in the second row is
different and may be called B type. When the third row is placed adjacent to
the second in staggered manner, its spheres are aligned with those of the first
layer. Hence this layer is also of A type. The spheres of similarly placed
fourth row will be aligned with those of the second row (B type). Hence this
arrangement is of ABAB type. In this arrangement there is less free space and
this packing is more efficient than the square close packing. Each sphere is in
contact with six of its neighbours and the two dimensional coordination
number is 6. The centres of these six spheres are at the corners of a regular
hexagon (Fig. 1.14b) hence this packing is called two dimensional hexagonal
close-packing. It can be seen in Figure 1.14 (b) that in this layer there are
some voids (empty spaces). These are triangular in shape. The triangular
voids are of two different types. In one row, the apex of the triangles are
pointing upwards and in the next layer downwards.
All real structures are three dimensional structures. They can be obtained by
stacking two dimensional layers one above the other. In the last Section, we
discussed close packing in two dimensions which can be of two types; square
close-packed and hexagonal close-packed. Let us see what types of three
dimensional close packing can be obtained from these.
(i) Three dimensional close packing from two dimensional square close-
packed layers: While placing the second square close-packed layer above the
first we follow the same rule that was followed when one row was placed
adjacent to the other. The second layer is placed over the first layer such that
the spheres of the upper layer are exactly above those of the first layer. In this
arrangement spheres of both the layers are perfectly aligned horizontally as
well as vertically as shown in Fig. 1.15.
Similarly, we may place more layers one above the other. If the arrangement
of spheres in the first layer is called A type, all the layers have the same
arrangement. Thus this lattice has AAA.... type pattern. The lattice thus
generated is the simple cubic lattice, and its unit cell is the primitive cubic
unit cell (See Fig. 1.9).
Fig. 1.15: Simple cubic lattice formed by A A A .... arrangement
(ii) Three dimensional close packing from two dimensional hexagonal close
packed layers: Three dimensional close packed structure can be generated by
placing layers one over the other.
Let us take a two dimensional hexagonal close packed layer A and place a
similar layer above it such that the spheres of the second layer are placed in
the depressions of the first layer. Since the spheres of the two layers are
aligned differently, let us call the second layer as B. It can be observed from
Fig. 1.16 that not all the triangular voids of the first layer are covered by the
spheres of the second layer. This gives rise to different arrangements.
Wherever a sphere of the second layer is above the void of the first layer (or
vice versa) a tetrahedral void is formed. These voids are called tetrahedral
voids because a tetrahedron is formed when the centres of these four spheres
are joined. They have been marked as T in Fig. 1.16. One such void has
been shown separately in Fig. 1.17.
Fig. 1.16: A stack of two layers of close packed spheres and voids generated in them. T =
Tetrahedral void; O = Octahedral void
Fig 1.17 Tetrahedral and octahedral voids (a) top view (b) exploded side view and (c)
geometrical shape of the void.
At other places, the triangular voids in the second layer are above the
triangular voids in the first layer, and the triangular shapes of these do not
overlap. One of them has the apex of the triangle pointing upwards and the
other downwards. These voids have been marked as O in Fig. 1.16. Such
voids are surrounded by six spheres and are called octahedral voids. One such
void has been shown separately in Fig. 1.17. The number of these two types
of voids depend upon the number of close packed spheres.
(i) Covering Tetrahedral Voids: Tetrahedral voids of the second layer may be
covered by the spheres of the third layer. In this case, the spheres of the third
layer are exactly aligned with those of the first layer. Thus, the pattern of
spheres is repeated in alternate layers. This pattern is often written as ABAB
....... pattern. This structure is called hexagonal close packed (hcp) structure
(Fig. 1.18). This sort of arrangement of atoms is found in many metals like
magnesium and zinc.
Fig. 1.18 (a) Hexagonal cubic close-packing exploded view showing stacking of layers of
spheres (b) four layers stacked in each case and (c) geometry of packing.
Fig. 1.19 (a) ABCABC... arrangement of layers when octahedral void is covered (b)
fragment of structure formed by this arrangement resulting in cubic closed packed (ccp) or
face centred cubic (fcc) structure.
(ii) Covering Octahedral Voids: The third layer may be placed above the
second layer in a manner such that its spheres cover the octahedral voids.
When placed in this manner, the spheres of the third layer are not aligned
with those of either the first or the second layer. This arrangement is called
C type. Only when fourth layer is placed, its spheres are aligned with those
of the first layer as shown in Figs. 1.18 and 1.19. This pattern of layers is
often written as ABCABC ........... This structure is called cubic close packed
(ccp) or face-centred cubic (fcc) structure. Metals such as copper and silver
crystallise in this structure.
Both these types of close packing are highly efficient and 74% space in the
crystal is filled. In either of them, each sphere is in contact with twelve
spheres. Thus, the coordination number is 12 in either of these two structures.
1.6.1 Formula of a Compound and Number of Voids
Filled
Earlier in the section, we have learnt that when particles are close-packed
resulting in either ccp or hcp structure, two types of voids are generated.
While the number of octahedral voids present in a lattice is equal to the
number of close packed particles, the number of tetrahedral voids generated
is twice this number. In ionic solids, the bigger ions (usually anions) form the
close packed structure and the smaller ions (usually cations) occupy the
voids. If the latter ion is small enough then tetrahedral voids are occupied, if
bigger, then octahedral voids. Not all octahedral or tetrahedral voids are
occupied. In a given compound, the fraction of octahedral or tetrahedral voids
that are occupied, depends upon the chemical formula of the compound, as
can be seen from the following examples.
Example 1.1
Solution
Example 1.2
Atoms of element B form hcp lattice and those of the element A occupy
2/3rd of tetrahedral voids. What is the formula of the compound formed
by the elements A and B?
Solution
Let us consider a unit cell of ccp or fcc lattice [Fig. 1(a)]. The unit cell is
divided into eight small cubes.
Each small cube has atoms at alternate corners [Fig. 1(a)]. In all, each
small cube has 4 atoms. When joined to each other, they make a regular
tetrahedron. Thus, there is one tetrahedral void in each small cube and
eight tetrahedral voids in total. Each of the eight small cubes have one
void in one unit cell of ccp structure. We know that ccp structure has 4
atoms per unit cell. Thus, the number of tetrahedral voids is twice the
number of atoms.
Fig. 1: (a) Eight tetrahedral voids per unit cell of ccp structure (b) one tetrahedral
void showing the geometry.
Let us again consider a unit cell of ccp or fcc lattice [Fig. 2(a)]. The body
centre of the cube, C is not occupied but it is surrounded by six atoms on
face centres. If these face centres are joined, an octahedron is generated.
Thus, this unit cell has one octahedral void at the body centre of the
cube.
Besides the body centre, there is one octahedral void at the centre of each
of the 12 edges. [Fig. 2(b)]. It is surrounded by six atoms, four belonging
to the same unit cell (2 on the corners and 2 on face centre) and two
belonging to two adjacent unit cells. Since each edge of the cube is
shared between four adjacent unit cells, so is the octahedral void located
Fig. 2: Location of octahedral voids per unit cell of ccp or fcc lattice (a) at the body
centre of the cube and (b) at the centre of each edge (only one such void is shown).
12 octahedral voids located at each edge and shared between four unit
cells
We know that in ccp structure, each unit cell has 4 atoms. Thus, the
number of octahedral voids is equal to this number.
1.7 Packing Efficiency
Both types of close packing (hcp and ccp) are equally efficient. Let us
calculate the efficiency of packing in ccp structure. In Fig. 1.20 let the unit
cell edge length be a and face diagonal AC = b.
In
= a2+a2 = 2a2 or
b=
b = 4r =
or a =
(we can also write,
We know, that each unit cell in ccp structure, has effectively 4 spheres. Total
is a3 or .
Therefore,
Fig. 1.20: Cubic close packing other sides are not provided with spheres for sake of clarity.
1.7.2 Efficiency of Packing in Body- Centred Cubic
Structures
From Fig. 1.21, it is clear that the atom at the centre will be in touch with the
other two atoms diagonally arranged.
In EFD,
b2 = a2 + a2 = 2a2
b=
Now in AFD
c2 = a2 + b2 = a2 + 2a2 = 3a2
c=
The length of the body diagonal c is equal to 4r, where r is the radius of the
sphere (atom), as all the three spheres along the diagonal touch each other.
Therefore, = 4r
a=
or .
Therefore,
Fig. 1.21: Body-centred cubic unit cell (sphere along the body diagonal are shown
with solid boundaries).
In a simple cubic lattice the atoms are located only on the corners of the cube.
The particles touch each other along the edge (Fig. 1.22).
Thus, the edge length or side of the cube a, and the radius of each particle, r
are related as
a = 2r
Packing efficiency
=
= 52.36% = 52.4 %
Thus, we may conclude that ccp and hcp structures have maximum packing
efficiency.
Fig. 1.22 Simple cubic unit cell. The spheres are in contact with each other along the edge
of the cube.
From the unit cell dimensions, it is possible to calculate the volume of the
unit cell. Knowing the density of the metal, we can calculate the mass of the
atoms in the unit cell. The determination of the mass of a single atom gives
an accurate method of determination of Avogadro constant. Suppose, edge
length of a unit cell of a cubic crystal determined by X-ray diffraction is a, d
the density of the solid substance and M the molar mass. In case of cubic
crystal:
(Here z is the number of atoms present in one unit cell and m is the mass of a
single atom)
m (M is molar mass)
Remember, the density of the unit cell is the same as the density of the
substance. The density of the solid can always be determined by other
methods. Out of the five parameters (d, z M, a and NA), if any four are
known, we can determine the fifth.
Example 1.3
Since each bcc cubic unit cell contains 2 atoms, therefore, the total
number
of atoms in 208 g = 2 (atoms/unit cell) 12.08 1023 unit cells
= 24.161023 atoms
Example 1.4
X-ray diffraction studies show that copper crystallises in an fcc unit cell
with cell edge of 3.60810-8 cm. In a separate experiment, copper is
determined to have a density of 8.92 g/cm3, calculate the atomic mass of
copper
Solution
Therefore,
= 63.1 g/mol
Atomic mass of copper = 63.1u
Example 1.5
Silver forms ccp lattice and X-ray studies of its crystals show that the
edge length of its unit cell is 408.6 pm. Calculate the density of silver
(Atomic mass = 107.9 u).
Solution
Since the lattice is ccp, the number of silver atoms per unit cell = z = 4
Density,
Intext Questions
1.14 What is the two dimensional coordination number of a molecule in
square close-packed layer?
1.17 Which of the following lattices has the highest packing efficiency
(i) simple cubic (ii) body-centred cubic and (iii) hexagonal close-packed
lattice?
1.18 An element with molar mass 2.710-2 kg mol-1 forms a cubic unit
cell with edge length 405 pm. If its density is 2.7103 kg m-3, what is the
nature of the cubic unit cell?
Although crystalline solids have short range as well as long range order in the
arrangement of their constituent particles, yet crystals are not perfect. Usually
a solid consists of an aggregate of large number of small crystals. These
small crystals have defects in them. This happens when crystallisation
process occurs at fast or moderate rate. Single crystals are formed when the
process of crystallisation occurs at extremely slow rate. Even these crystals
are not free of defects. The defects are basically irregularities in the
arrangement of constituent particles. Broadly speaking, the defects are of two
types, namely, point defects and line defects. Point defects are the
irregularities or deviations from ideal arrangement around a point or an atom
in a crystalline substance, whereas the line defects are the irregularities or
deviations from ideal arrangement in entire rows of lattice points. These
irregularities are called crystal defects. We shall confine our discussion to
point defects only.
Point defects can be classified into three types : (i) stoichiometric defects (ii)
impurity defects and (iii) non-stoichiometric defects.
These are the point defects that do not disturb the stoichiometry of the solid.
They are also called intrinsic or thermodynamic defects. Basically these are
of two types, vacancy defects and interstitial defects.
(i) Vacancy Defect: When some of the lattice sites are vacant, the crystal is
said to have vacancy defect (Fig. 1.23). This results in decrease in density of
the substance. This defect can also develop when a substance is heated.
(iii) Frenkel Defect: This defect is shown by ionic solids. The smaller ion
(usually cation) is dislocated from its normal site to an interstitial site (Fig.
1.25). It creates a vacancy defect at its original site and an interstitial defect
at its new location.
Frenkel defect is also called dislocation defect. It does not change the
density of the solid. Frenkel defect is shown by ionic substance in which
there is a large difference in the size of ions, for example, ZnS, AgCl, AgBr
and AgI due to small size of Zn2+ and Ag+ ions.
Like simple vacancy defect, Schottky defect also decreases the density of the
substance. Number of such defects in ionic solids is quite significant. For
example, in NaCl there are approximately 106 Schottky pairs per cm3 at room
temperature. In 1 cm3 there are about 1022 ions. Thus, there is one Schottky
defect per 1016 ions. Schottky defect is shown by ionic substances in which
the cation and anion are of almost similar sizes. For example, NaCl, KCl,
CsCl and AgBr. It may be noted that AgBr shows both, Frenkel as well as
Schottky defects.
Now there is excess of zinc in the crystal and its formula becomes Zn1+xO.
The excess Zn2+ ions move to interstitial sites and the electrons to
neighbouring interstitial sites.
There are many solids which are difficult to prepare in the stoichiometric
composition and contain less amount of the metal as compared to the
stoichiometric proportion. A typical example of this type is FeO which is
mostly found with a composition of Fe0.95O. It may actually range from
Fe0.93O to Fe0.96O. In crystals of FeO some Fe2+ cations are missing and the
loss of positive charge is made up by the presence of required number of Fe3+
ions.
Fig. 1.28: An F-centre in a crystal
(i) Conductors: The solids with conductivities ranging between 104 to 107
ohm1m1 are called conductors. Metals have conductivities in the order of
107 ohm1m1 are good conductors.
(ii) Insulators : These are the solids with very low conductivities ranging
between 1020 to 1010 ohm1m1.
If the gap between filled valence band and the next higher unoccupied band
(conduction band) is large, electrons cannot jump to it and such a substance
has very small conductivity and it behaves as an insulator (Fig. 1.29 b).
In case of semiconductors, the gap between the valence band and conduction
band is small (Fig. 1.29c). Therefore, some electrons may jump to conduction
band and show some conductivity. Electrical conductivity of semiconductors
increases with rise in temperature, since more electrons can jump to the
conduction band. Substances like silicon and germanium show this type of
behaviour and are called intrinsic semiconductors.
Fig. 1.29 Distinction among (a) metals (b) insulators and (c) semiconductors. In each case,
an unshaded area represents a conduction band.
Silicon and germanium belong to group 14 of the periodic table and have four
valence electrons each. In their crystals each atom forms four covalent bonds
with its neighbours (Fig. 1.30 a). When doped with a group 15 element like P
or As, which contains five valence electrons, they occupy some of the lattice
sites in silicon or germanium crystal (Fig. 1.30 b). Four out of five electrons
are used in the formation of four covalent bonds with the four neighbouring
silicon atoms. The fifth electron is extra and becomes delocalised. These
delocalised electrons increase the conductivity of doped silicon (or
germanium). Here the increase in conductivity is due to the negatively
charged electron, hence silicon doped with electron-rich impurity is called n-
type semiconductor.
Fig. 1.30: Creation of n-type and p-type semiconductors by doping groups 13 and 15
elements.
Applications of n-type and p-type semiconductors
Fig.1.31: Demonstration of the magnetic moment associated with (a) an orbiting electron
and (b) a spinning electron.
Intext Questions
1.19 What type of defect can arise when a solid is heated? Which
physical property is affected by it and in what way?
1.20 What type of stoichiometric defect is shown by: (i) ZnS (ii) AgBr
Summary
Solids have definite mass, volume and shape. This is due to the fixed
position of their constituent particles, short distances and strong
interactions between them. In amorphous solids, the arrangement of
constituent particles has only short range order and consequently they
behave like super cooled liquids, do not have sharp melting points and
are isotropic in nature. In crystalline solids there is long range order in
the arrangement of their constituent particles. They have sharp melting
points, are anisotropic in nature and their particles have characteristic
shapes. Properties of crystalline solids depend upon the nature of
interactions between their constituent particles. On this basis, they can be
divided into four categories, namely: molecular, ionic, metallic and
covalent solids. They differ widely in their properties.
Exercises
1.2 What makes a glass different from a solid such as quartz? Under
what conditions could quartz be converted into glass?
(iv) I2 (ix) Rb
1.5 How can you determine the atomic mass of an unknown metal if you
know its density and the dimension of its unit cell? Explain.
1.8 How many lattice points are there in one unit cell of each of the
following lattice?
(i) The basis of similarities and differences between metallic and ionic
crystals.
(ii) Ionic solids are hard and brittle.
1.14 If the radius of the octahedral void is r and radius of the atoms in
closepacking is R, derive relation between r and R.
1.15 Copper crystallises into a fcc lattice with edge length 3.61 108
cm. Show that the calculated density is in agreement with its measured
value of 8.92 g cm3.
1.16 Analysis shows that nickel oxide has the formula Ni0.98O1.00. What
fractions of nickel exist as NI2+ and Ni3+ ions?
1.17 What is a semiconductor? Describe the two main types of
semiconductors and contrast their conduction mechanism.
1.14 4
1.15 Total number of voids = 9.033 1023 Number of tetrahedral voids =
6.022 1023
1.16 M2N3
1.18 ccp
Chapter Two
2.1 INTRODUCTION
In Chapters 6 and 8 (Class XI), the notion of potential energy was
introduced. When an external force does work in taking a body from a
point to another against a force like spring force or gravitational force,
that work gets stored as potential energy of the body. When the
external force is removed, the body moves, gaining kinetic energy and
losing an equal amount of potential energy. The sum of kinetic
and potential energies is thus conserved. Forces of this kind are called
conservative forces. Spring force and gravitational force are examples
of conservative forces.
Figure 2.1 A test charge q (> 0) is moved from the point R to the point P against
the repulsive force on it by the charge Q(> 0) placed at the origin.
Two remarks may be made here. First, we assume that the test
charge q is so small that it does not disturb the original configuration,
namely the charge Q at the origin (or else, we keep Q fixed at the
origin by some unspecified force). Second, in bringing the charge q
from R to P, we apply an external force Fext just enough to counter the
repulsive electric force FE (i.e, Fext= FE). This means there is no net
force on or acceleration of the charge q when it is brought from R to P,
i.e., it is brought with infinitesimally slow constant speed. In this
situation, work done by the external force is the negative of the work
done by the electric force, and gets fully stored in the form of potential
energy of the charge q. If the external force is removed on reaching P,
the electric force will take the charge away from Q the stored energy
(potential energy) at P is used to provide kinetic energy to the charge
q in such a way that the sum of the kinetic and potential energies is
conserved.
Thus, work done by external forces in moving a charge q from R to P
is
(2.1)
This work done is against electrostatic repulsive force and gets stored
as potential energy.
At every point in electric field, a particle with charge q possesses a
certain electrostatic potential energy, this work done increases its
potential energy by an amount equal to potential energy difference
between points R and P.
Thus, potential energy difference
(2.2)
(2.3)
Since the point P is arbitrary, Eq. (2.3) provides us with a definition of
potential energy of a charge q at any point. Potential energy of charge
q at a point (in the presence of field due to any charge configuration)
is the work done by the external force (equal and opposite to the
electric force) in bringing the charge q from infinity to that point.
= VP VR (2.4)
where VP and VR are the electrostatic potentials at P and R,
respectively. Note, as before, that it is not the actual value of potential
but the potential difference that is physically significant. If, as before,
we choose the potential to be zero at infinity, Eq. (2.4) implies:
Work done by an external force in bringing a unit positive charge from
infinity to a point = electrostatic potential (V) at that point.
Figure 2.2 Work done on a test charge q by the electrostatic field due to any given
charge configuration is independent of the path, and depends only on its initial and
final positions.
where is the unit vector along OP. Work done against this force
from r to r + r is
(2.6)
The negative sign appears because for r < 0, W is positive . Total
work done (W) by the external force is obtained by integrating Eq.
(2.6) from r = to r = r,
(2.7)
(2.8)
Equation (2.8) is true for any sign of the charge Q, though we
considered Q > 0 in its derivation. For Q < 0, V < 0, i.e., work done (by
the external force) per unit positive test charge in bringing it from
infinity to the point is negative. This is equivalent to saying that work
done by the electrostatic force in bringing the unit positive charge form
infinity to the point P is positive. [This is as it should be, since for Q <
0, the force on a unit positive test charge is attractive, so that the
electrostatic force and the displacement (from infinity to P) are in the
same direction.] Finally, we note that Eq. (2.8) is consistent with the
choice that potential at infinity be zero.
Figure 2.4 Variation of potential V with r [in units of (Q/40) m-1] (blue curve) and
field with r [in units of (Q/40) m-2] (black curve) for a point charge Q.
Figure (2.4) shows how the electrostatic potential ( 1/r) and the
electrostatic field ( 1/r2 ) varies with r.
Example 2.1
(a) Calculate the potential at a point P due to a charge of 4 10
7C located 9 cm away.
(a)
= 4 104 V
(b)
= 8 105 J
No, work done will be path independent. Any arbitrary infinitesimal
path can be resolved into two perpendicular displacements: One
along r and another perpendicular to r. The work done
corresponding to the later will be zero.
2.4 POTENTIAL DUE TO AN ELECTRIC DIPOLE
As we learnt in the last chapter, an electric dipole consists of two
charges q and q separated by a (small) distance 2a. Its total charge
is zero. It is characterised by a dipole moment vector p whose
magnitude is q 2a and which points in the direction from q to q (Fig.
2.5). We also saw that the electric field of a dipole at a point with
position vector r depends not just on the magnitude r, but also on the
angle between r and p. Further, the field falls off, at large distance, not
as 1/r2 (typical of field due to a single charge) but as 1/r3. We, now,
determine the electric potential due to a dipole and contrast it with the
potential due to a single charge.
As before, we take the origin at the centre of the dipole. Now we know
that the electric field obeys the superposition principle. Since potential
is related to the work done by the field, electrostatic potential also
follows the superposition principle. Thus, the potential due to the
dipole is the sum of potentials due to the charges q and q
(2.9)
where r1 and r2 are the distances of the point P from q and q,
respectively.
Now, by geometry,
cos
cos (2.10)
Figure 2.5 Quantities involved in the calculation of potential due to a dipole.
(2.11)
Similarly,
(2.12)
Using the Binomial theorem and retaining terms upto the first order in
a/r ; we obtain,
[2.13(a)]
[2.13(b)]
Using Eqs. (2.9) and (2.13) and p = 2qa, we get
(2.14)
Now, p cos =
; (r >> a) (2.15)
Equation (2.15) is, as indicated, approximately true only for distances
large compared to the size of the dipole, so that higher order terms
in a/r are negligible. For a point dipole p at the origin, Eq. (2.15) is,
however, exact.
From Eq. (2.15), potential on the dipole axis ( = 0, ) is given by
(2.16)
(Positive sign for = 0, negative sign for = .) The potential in the
equatorial plane ( = /2) is zero.
The important contrasting features of electric potential of a dipole from
that due to a single charge are clear from Eqs. (2.8) and (2.15):
(i) The potential due to a dipole depends not just on r but also on the
angle between the position vector r and the dipole moment vector p.
(It is, however, axially symmetric about p. That is, if you rotate the
position vector r about p, keeping fixed, the points corresponding to
P on the cone so generated will have the same potential as at P.)
(ii) The electric dipole potential falls off, at large distance, as 1/r2, not
as 1/r, characteristic of the potential due to a single charge. (You can
refer to the Fig. 2.5 for graphs of 1/r2 versus r and 1/r versus r, drawn
there in another context.)
,
where r2P and r3P are the distances of P from charges q2 and q3,
respectively; and so on for the potential due to other charges. By the
superposition principle, the potential V at P due to the total charge
configuration is the algebraic sum of the potentials due to the
individual charges
V = V1 + V2 + ... + Vn (2.17)
(2.18)
Figure 2.6 Potential at a point due to a system of charges is the sum of potentials
due to individual charges.
[2.19(a)]
where q is the total charge on the shell and R its radius. The electric
field inside the shell is zero. This implies (Section 2.6) that potential is
constant inside the shell (as no work is done in moving a charge
inside the shell), and, therefore, equals its value at the surface, which
is
[2.19(b)]
which gives
x = 45 cm
Thus, electric potential is zero at 9 cm and 45 cm away from the
positive charge on the side of the negative charge. Note that the
formula for potential used in the calculation required choosing
potential to be zero at infinity.
Example 2.3 Figures 2.8 (a) and (b) show the field lines of a
positive and negative point charge respectively.
Figure 2.8
(d) Give the sign of the work done by the external agency in
moving a small negative charge from B to A.
(e) Does the kinetic energy of a small negative charge increase or
decrease in going from B to A?
Solution
electric field lines are radial, starting from the charge if q > 0.
Now the electric field lines for a single charge q are radial lines
starting from or ending at the charge, depending on whether q is
positive or negative. Clearly, the electric field at every point is normal
to the equipotential surface passing through that point. This is true in
general: for any charge configuration, equipotential surface through a
point is normal to the electric field at that point. The proof of this
statement is simple.
If the field were not normal to the equipotential surface, it would have
non-zero component along the surface. To move a unit test charge
against the direction of the component of the field, work would have to
be done. But this is in contradiction to the definition of an equipotential
surface: there is no potential difference between any two points on the
surface and no work is required to move a test charge on the surface.
The electric field must, therefore, be normal to the equipotential
surface at every point. Equipotential surfaces offer an alternative
visual picture in addition to the picture of electric field lines around a
charge configuration.
For a uniform electric field E, say, along the -axis, the equipotential
surfaces are planes normal to the -axis, i.e., planes parallel to the y-
z plane (Fig. 2.10). Equipotential surfaces for (a) a dipole and (b) two
identical positive charges are shown in Fig. 2.11.
Figure 2.11 Some equipotential surfaces for (a) a dipole, (b) two identical positive
charges.
Thus,
|E| l = V (V +V)= V
where r1P is the distance of a point P in space from the location of q1.
From the definition of potential, work done in bringing charge q2 from
infinity to the point r2 is q2 times the potential at r2 due to q1:
work done on q2 =
where r12 is the distance between points 1 and 2.
(2.22)
Obviously, if q2 was brought first to its present location and q1 brought
later, the potential energy U would be the same. More generally, the
potential energy expression, Eq. (2.22), is unaltered whatever way the
charges are brought to the specified locations, because of path-
independence of work for electrostatic force.
Equation (2.22) is true for any sign of q1and q2. If q1q2 > 0, potential
energy is positive. This is as expected, since for like charges (q1q2 >
0), electrostatic force is repulsive and a positive amount of work is
needed to be done against this force to bring the charges from infinity
to a finite distance apart. For unlike charges (q1 q2 < 0), the
electrostatic force is attractive. In that case, a positive amount of work
is needed against this force to take the charges from the given
location to infinity. In other words, a negative amount of work is
needed for the reverse path (from infinity to the present locations), so
the potential energy is negative.
2.14 Potential energy of a system of three charges is given by Eq. (2.26), with the
notation given in the figure.
(2.23)
The charges q1 and q2 produce a potential, which at any point P is
given by
(2.24)
(2.25)
The total work done in assembling the charges at the given locations
is obtained by adding the work done in different steps [Eq. (2.23) and
Eq. (2.25)],
(2.26)
Again, because of the conservative nature of the electrostatic force (or
equivalently, the path independence of work done), the final
expression for U, Eq. (2.26), is independent of the manner in which
the configuration is assembled. The potential energy is characteristic
of the present state of configuration, and not the way the state is
achieved.
Figure 2.15
Solution
(a) Since the work done depends on the final arrangement of the
charges, and not on how they are put together, we calculate work
needed for one way of putting the charges at A, B, C and D.
Suppose, first the charge +q is brought to A, and then the charges
q, +q, and q are brought to B, C and D, respectively. The total
work needed can be calculated in steps:
Add the work done in steps (i), (ii), (iii) and (iv). The total
work required is
= qV(r) (2.27)
where V(r) is the external potential at the point r.
Thus, if an electron with charge q = e = 1.61019 C is accelerated by
a potential difference of V = 1 volt, it would gain energy of qV = 1.6
1019J. This unit of energy is defined as 1 electron volt or 1eV, i.e.,
1 eV=1.6 1019J. The units based on eV are most commonly used in
atomic, nuclear and particle physics, (1 keV = 103eV = 1.6 1016J, 1
MeV
= 106eV = 1.6 1013J, 1 GeV = 109eV = 1.6 1010J and 1 TeV =
1012eV
= 1.6 107J). [This has already been defined on Page 117, XI
Physics Part I, Table 6.1.]
where r12 is the distance between q1 and q2. We have made use of
Eqs. (2.27) and (2.22). By the superposition principle for fields, we add
up the work done on q2 against the two fields (E and that due to q1):
Work done in bringing q2 to r2
(2.28)
Thus,
Potential energy of the system
= the total work done in assembling the configuration
(2.29)
Example 2.5
(a) Determine the electrostatic potential energy of a system
consisting of two charges 7 C and 2 C (and with no external
field) placed at (9 cm, 0, 0) and (9 cm, 0, 0) respectively.
(b) How much work is required to separate the two charges
infinitely away from each other?
(c) Suppose that the same system of charges is now placed in an
external electric field E = A (1/r2); A = 9 105 C m2. What would
the electrostatic energy of the configuration be?
Solution
(a) = 0.7 J.
(b) W = U2 U1 = 0 U = 0 (0.7) = 0.7 J.
(2.31)
This work is stored as the potential energy of the system. We can then
associate potential energy U() with an inclination of the dipole.
Similar to other potential energies, there is a freedom in choosing the
angle where the potential energy U is taken to be zero. A natural
choice is to take
0 = / 2. (n explanation for it is provided towards the end of
discussion.) We can then write,
(2.32)
(2.33)
Here, r1 and r2 denote the position vectors of +q and q. Now, the
potential difference between positions r1 and r2 equals the work
done in bringing a unit positive charge against field from r2 to r1. The
displacement parallel to the force is 2a cos. Thus, [V(r1)V (r2)] = E
2a cos . We thus obtain,
(2.34)
We note that U () differs from U() by a quantity which is just a
constant for a given dipole. Since a constant is insignificant for
potential energy, we can drop the second term in Eq. (2.34) and it
then reduces to Eq. (2.32).
We can now understand why we took 0= /2. In this case, the work
done against the external field E in bringing +q and q are equal and
opposite and cancel out, i.e., q [V (r1) V (r2)]=0.
(2.35)
Just inside the surface, the electrostatic field is zero; just outside, the
field is normal to the surface with magnitude E. Thus, the contribution
to the total flux through the pill box comes only from the outside
(circular) cross-section of the pill box. This equals ES (positive for
> 0, negative for < 0), since over the small area S, E may be
considered constant and E and S are parallel or antiparallel. The
charge enclosed by the pill box is S.
By Gausss law
ES =
E= (2.36)
Including the fact that electric field is normal to the surface, we get the
vector relation, Eq. (2.35), which is true for both signs of . For > 0,
electric field is normal to the surface outward; for < 0, electric field is
normal to the surface inward.
6. Electrostatic shielding
Consider a conductor with a cavity, with no charges inside the cavity.
A remarkable result is that the electric field inside the cavity is zero,
whatever be the size and shape of the cavity and whatever be the
charge on the conductor and the external fields in which it might be
placed. We have proved a simple case of this result already: the
electric field inside a charged spherical shell is zero. The proof of the
result for the shell makes use of the spherical symmetry of the shell
(see Chapter 1). But the vanishing of electric field in the (charge-free)
cavity of a conductor is, as mentioned above, a very general result. A
related result is that even if the conductor is charged or charges are
induced on a neutral conductor by an external field, all charges reside
only on the outer surface of a conductor with cavity.
Figure 2.18 The electric field inside a cavity of any conductor is zero. All charges
reside only on the outer surface of a conductor with cavity. (There are no charges
placed in the cavity.)
The proofs of the results noted in Fig. 2.18 are omitted here, but we
note their important implication. Whatever be the charge and field
configuration outside, any cavity in a conductor remains shielded from
outside electric influence: the field inside the cavity is always zero.
This is known as electrostatic shielding. The effect can be made use
of in protecting sensitive instruments from outside electrical influence.
Figure 2.19 gives a summary of the important electrostatic properties
of a conductor.
Figure 2.22 A dielectric develops a net dipole moment in an external electric field.
(a) Non-polar molecules, (b) Polar molecules.
A dielectric with polar molecules also develops a net dipole moment in
an external field, but for a different reason. In the absence of any
external field, the different permanent dipoles are oriented randomly
due to thermal agitation; so the total dipole moment is zero. When an
external field is applied, the individual dipole moments tend to align
with the field. When summed over all the molecules, there is then a
net dipole moment in the direction of the external field, i.e., the
dielectric is polarised. The extent of polarisation depends on the
relative strength of two mutually opposite factors: the dipole potential
energy in the external field tending to align the dipoles with the field
and thermal energy tending to disrupt the alignment. There may be, in
addition, the induced dipole moment effect as for non-polar
molecules, but generally the alignment effect is more important for
polar molecules.
Thus in either case, whether polar or non-polar, a dielectric develops a
net dipole moment in the presence of an external field. The dipole
moment per unit volume is called polarisation and is denoted by P. For
linear isotropic dielectrics,
(2.37)
where e is a constant characteristic of the dielectric and is known as
the electric susceptibility of the dielectric medium.
It is possible to relate e to the molecular properties of the substance,
but we shall not pursue that here.
The question is: how does the polarised dielectric modify the original
external field inside it? Let us consider, for simplicity, a rectangular
dielectric slab placed in a uniform external field E0 parallel to two of its
faces. The field causes a uniform polarisation P of the dielectric. Thus
every volume element v of the slab has a dipole moment P v in the
direction of the field. The volume element v is macroscopically small
but contains a very large number of molecular dipoles. Anywhere
inside the dielectric, the volume element v has no net charge (though
it has net dipole moment). This is, because, the positive charge of one
dipole sits close to the negative charge of the adjacent dipole.
However, at the surfaces of the dielectric normal to the electric field,
there is evidently a net charge density. As seen in Fig 2.23, the
positive ends of the dipoles remain unneutralised at the right surface
and the negative ends at the left surface. The unbalanced charges are
the induced charges due to the external field.
Figure 2.23 A uniformly polarised dielectric amounts to induced surface
charge density, but no volume charge density.
(2.38)
The constant C is called the capacitance of the capacitor. C is
independent of Q or V, as stated above. The capacitance C depends
only on the geometrical configuration (shape, size, separation) of the
system of two conductors. [As we shall see later, it also depends on
the nature of the insulator (dielectric) separating the two conductors.]
The SI unit of capacitance is 1 farad (=1 coulomb volt-1) or 1 F = 1 C
V1. A capacitor with fixed capacitance is symbolically shown as ---||---
, while the one with variable capacitance is shown as .
Equation (2.38) shows that for large C, V is small for a given Q. This
means a capacitor with large capacitance can hold large amount of
charge Q at a relatively small V. This is of practical importance. High
potential difference implies strong electric field around the conductors.
A strong electric field can ionise the surrounding air and accelerate the
charges so produced to the oppositely charged plates, thereby
neutralising the charge on the capacitor plates, at least partly. In other
words, the charge of the capacitor leaks away due to the reduction in
insulating power of the intervening medium.
The maximum electric field that a dielectric medium can withstand
without break-down (of its insulating property) is called its dielectric
strength; for air it is about 3 106 Vm1. For a separation between
conductors of the order of 1 cm or so, this field corresponds to a
potential difference of 3 104 V between the conductors. Thus, for a
capacitor to store a large amount of charge without leaking, its
capacitance should be high enough so that the potential difference
and hence the electric field do not exceed the break-down limits. Put
differently, there is a limit to the amount of charge that can be stored
on a given capacitor without significant leaking. In practice, a farad is
a very big unit; the most common units are its sub-multiples 1 F =
106 F, 1 nF = 109 F, 1 pF = 1012 F, etc. Besides its use in storing
charge, a capacitor is a key element of most ac circuits with important
functions, as described in Chapter 7.
2.12 THE PARALLEL PLATE CAPACITOR
A parallel plate capacitor consists of two large plane
parallel conducting plates separated by a small distance (Fig. 2.25).
(2.40)
In the inner region between the plates 1 and 2, the electric fields due
to the two charged plates add up, giving
(2.41)
The direction of electric field is from the positive to the negative plate.
Thus, the electric field is localised between the two plates and is
uniform throughout. For plates with finite area, this will not be true
near the outer boundaries of the plates. The field lines bend outward
at the edges an effect called fringing of the field. By the same
token, will not be strictly uniform on the entire plate. [E and are
related by Eq. (2.35).] However, for d2 << A, these effects can be
ignored in the regions sufficiently far from the edges, and the field
there is given by Eq. (2.41). Now for uniform electric field, potential
difference is simply the electric field times the distance between the
plates, that is,
(2.42)
The capacitance C of the parallel plate capacitor is then
= (2.43)
(2.44)
(2.45)
which is a plate about 30 km in length and breadth!
http://micro.magnet.fsu.edu/electromag/java/capacitance/
2.13 EFFECT OF DIELECTRIC ON CAPACITANCE
With the understanding of the behavior of dielectrics in an external
field developed in Section 2.10, let us see how the capacitance of a
parallel plate capacitor is modified when a dielectric is present. As
before, we have two large plates, each of area A, separated by a
distance d. The charge on the plates is Q, corresponding to the
charge density (with = Q/A). When there is vacuum between the
plates,
(2.46)
Consider next a dielectric inserted between the plates fully occupying
the intervening region. The dielectric is polarised by the field and, as
explained in Section 2.10, the effect is equivalent to two charged
sheets (at the surfaces of the dielectric normal to the field) with
surface charge densities p and p. The electric field in the dielectric
then corresponds to the case when the net surface charge density on
the plates is ( p).
That is,
(2.47)
(2.48)
For linear dielectrics, we expect p to be proportional to E0, i.e., to .
Thus, ( p) is proportional to and we can write
(2.49)
where K is a constant characteristic of the dielectric. Clearly, K > 1.
We then have
(2.50)
The capacitance C, with dielectric between the plates, is then
(2.51)
The product 0K is called the permittivity of the medium and is
denoted by
= 0 K (2.52)
For vacuum K = 1 and = 0; 0 is called the permittivity of the
vacuum. The dimensionless ratio
(2.53)
(2.54)
ELECTRIC DISPLACEMENT
We have introduced the notion of dielectric constant and arrived at
Eq. (2.54), without giving the explicit relation between the induced
charge density p and the polarisation P. We take without proof
or (0 E + P) =
The quantity 0 E + P is called the electric displacement and is
denoted by D. It is a vector quantity. Thus,
D = 0 E + P, D = ,
The significance of D is this : in vacuum, E is related to the free
charge density . When a dielectric medium is present, the
corresponding role is taken up by D. For a dielectric medium, it
is D not E that is directly related to free charge density , as seen
in above equation. Since P is in the same direction as E, all the
three vectors P, E and D are parallel.
The ratio of the magnitudes of D and E is
Thus,
D = 0 K E
and P = D 0E = 0 (K 1)E
This gives for the electric susceptibility e defined in Eq. (2.37)
e =0 (K1)
Example 2.8 A slab of material of dielectric constant K has the
same area as the plates of a parallel-plate capacitor but has a
thickness (3/4)d, where d is the separation of the plates. How is
the capacitance changed when the slab is inserted between the
plates?
V = V1 + V2 = (2.55)
i.e., , (2.56)
Figure 2.26 Combination of two capacitors in series.
(2.57)
We compare Eq. (2.57) with Eq. (2.56), and obtain
(2.58)
The proof clearly goes through for any number of capacitors arranged
in a similar way. Equation (2.55), for n capacitors arranged in series,
generalises to
(2.59)
Following the same steps as for the case of two capacitors, we get the
general formula for effective capacitance of a series combination of n
capacitors:
(2.60)
Figure 2.28 (a) shows two capacitors arranged in parallel. In this case,
the same potential difference is applied across both the capacitors.
But the plate charges (Q1) on capacitor 1 and the plate charges
(Q2) on the capacitor 2 are not necessarily the same:
Q1 = C1V, Q2 = C2V (2.61)
The equivalent capacitor is one with charge
Q = Q1 + Q2 (2.62)
Solution
(a) In the given network, C1, C2 and C3 are connected in series.
The effective capacitance C of these three capacitors is given by
C = C + C4 = F =13.3F
(b) Clearly, from the figure, the charge on each of the capacitors,
C1, C2 and C3 is the same, say Q. Let the charge on C4 be Q.
Now, since the potential difference across AB is Q/C1, across BC
is Q/C2, across CD is Q/C3 , we have
.
and
(2.68)
Since Q can be made as small as we like, Eq. (2.68) can be written
as
(2.69)
Equations (2.68) and (2.69) are identical because the term of second
order in Q, i.e., Q2/2C, is negligible, since Q is arbitrarily small.
The total work done (W) is the sum of the small work ( W) over the
very large number of steps involved in building the charge Q from
zero to Q.
= (2.70)
(2.71)
(2.72)
The same result can be obtained directly from Eq. (2.68) by
integration
= (2.74)
The surface charge density is related to the electric field E between
the plates,
(2.75)
From Eqs. (2.74) and (2.75) , we get
Energy stored in the capacitor
U= (2.76)
Note that Ad is the volume of the region between the plates (where
electric field alone exists). If we define energy density as energy
stored per unit volume of space, Eq (2.76) shows that
Energy density of electric field,
u =(1/2)0E2 (2.77)
Though we derived Eq. (2.77) for the case of a parallel plate capacitor,
the result on energy density of an electric field is, in fact, very general
and holds true for electric field due to any configuration of charges.
Solution
(a) The charge on the capacitor is
Q = CV = 900 1012 F 100 V = 9 108 C
= constant
(2.78)
Now, as shown in Fig. 2.32, let us suppose that in some way we
introduce a small sphere of radius r, carrying some charge q, into the
large one, and place it at the centre. The potential due to this new
charge clearly has the following values at the radii indicated:
Potential due to small sphere of radius r carrying charge q
SUMMARY
C=
where A is the area of each plate and d the separation between
them.
C = C1 + C2 + C3 + ...
where C1, C2, C3... are individual capacitances.
12. The energy U stored in a capacitor of capacitance C, with
charge Q and voltage V is
Points to Ponder
1. Electrostatics deals with forces between charges at rest. But if
there is a force on a charge, how can it be at rest? Thus, when we
are talking of r force between charges, it should be understood
that each charge is being kept at rest by some unspecified force
that opposes the net Coulomb force on the charge.
2. A capacitor is so configured that it confines the electric field
lines within a small region of space. Thus, even though field may
have considerable strength, the potential difference between the
two conductors of a capacitor is small.
3. Electric field is discontinuous across the surface of a spherical
EXERCISES
2.1 Two charges 5 108 C and 3 108 C are located 16 cm
apart. At what point(s) on the line joining the two charges is the
electric potential zero? Take the potential at infinity to be zero.
2.2 A regular hexagon of side 10 cm has a charge 5 C at each of
its vertices. Calculate the potential at the centre of the hexagon.
2.3 Two charges 2 C and 2 C are placed at points A and B 6
cm apart.
(a) Identify an equipotential surface of the system.
(b) What is the direction of the electric field at every point on this
surface?
2.4 A spherical conductor of radius 12 cm has a charge of 1.6
107C distributed uniformly on its surface. What is the electric field
(a) inside the sphere
(b) just outside the sphere
(c) at a point 18 cm from the centre of the sphere?
2.5 A parallel plate capacitor with air between the plates has a
capacitance of 8 pF (1pF = 1012 F). What will be the capacitance
if the distance between the plates is reduced by half, and the
space between them is filled with a substance of dielectric
constant 6?
2.8 In a parallel plate capacitor with air between the plates, each
plate has an area of 6 103 m2 and the distance between the
plates is 3 mm. Calculate the capacitance of the capacitor. If this
capacitor is connected to a 100 V supply, what is the charge on
each plate of the capacitor?
Additional Exercises
2.12 A charge of 8 mC is located at the origin. Calculate the work
done in taking a small charge of 2 109 C from a point P (0, 0, 3
cm) to a point Q (0, 4 cm, 0), via a point R (0, 6 cm, 9 cm).
2.13 A cube of side b has a charge q at each of its vertices.
Determine the potential and electric field due to this charge array
at the centre of the cube.
2.14 Two tiny spheres carrying charges 1.5 C and 2.5 C are
located 30 cm apart. Find the potential and electric field:
(a) at the mid-point of the line joining the two charges, and
(b) at a point 10 cm from this midpoint in a plane normal to the line
and passing through the mid-point.
2.15 A spherical conducting shell of inner radius r1 and outer
radius r2 has a charge Q.
(a) A charge q is placed at the centre of the shell. What is the
surface charge density on the inner and outer surfaces of the
shell?
(b) Is the electric field inside a cavity (with no charge) zero, even if
the shell is not spherical, but has any irregular shape? Explain.
2.16 (a) Show that the normal component of electrostatic field has
a discontinuity from one side of a charged surface to another given
by
(c) How much work is done in moving a small test charge from the
point (5,0,0) to (7,0,0) along the x-axis? Does the answer change
if the path of the test charge between the same points is not along
the x-axis?
2.22 Figure 2.34 shows a charge array known as an electric
quadrupole. For a point on the axis of the quadrupole, obtain the
dependence
of potential on r for r/a >> 1, and contrast your results with that due
to an electric dipole, and an electric monopole (i.e., a single
charge).
Figure 2.34
Figure 2.35
SEMICONDUCTOR ELECTRONICS:
MATERIALS, DEVICES AND SIMPLE
CIRCUITS
14.1 INTRODUCTION
(i) Metals: They possess very low resistivity (or high conductivity).
~ 102 108 m
~ 102 108 S m1
(ii) Semiconductors: They have resistivity or conductivity intermediate
to metals and insulators.
~ 105 106 m
~ 105 106 S m1
(iii) Insulators: They have high resistivity (or low conductivity).
~ 1011 1019 m
~ 1011 1019 S m1
The values of and given above are indicative of magnitude and
could well go outside the ranges as well. Relative values of the
resistivity are not the only criteria for distinguishing metals, insulators
and semiconductors from each other. There are some other
differences, which will become clear as we go along in this chapter.
Our interest in this chapter is in the study of semiconductors which
could be:
(i) Elemental semiconductors: Si and Ge
(ii) Compound semiconductors: Examples are:
Suppose these atoms start coming nearer to each other to form a solid. The
energies of these electrons in the outermost orbit may change (both increase
and decrease) due to the interaction between the electrons of different atoms.
The 6N states for l = 1, which originally had identical energies in the isolated
atoms, spread out and form an energy band [region B in Figure]. Similarly, the
2N states for l = 0, having identical energies in the isolated atoms, split into a
second band (carefully see the region B of Figure) separated from the first one
by an energy gap.
At still smaller spacing, however, there comes a region in which the bands
merge with each other. The lowest energy state that is a split from the upper
atomic level appears to drop below the upper state that has come from the
lower atomic level. In this region (region C in Figure), no energy gap exists
where the upper and lower energy states get mixed.
Finally, if the distance between the atoms further decreases, the energy bands
again split apart and are separated by an energy gap Eg (region D in Figure).
The total number of available energy states 8N has been re-apportioned
between the two bands (4N states each in the lower and upper energy bands).
Here the significant point is that there are exactly as many states in the lower
band (4N) as there are available valence electrons from the atoms (4N).
Therefore, this band (called the valence band) is completely filled while the
upper band is completely empty. The upper band is called the conduction
band.
Figure 14.1 The energy band positions in a semiconductor at 0 K. The upper band,
called the conduction band, consists of infinitely large number of closely spaced
energy states. The lower band, called the valence band, consists of closely spaced
completely filled energy states.
The gap between the top of the valence band and bottom of the
conduction band is called the energy band gap (Energy gap Eg). It
may be large, small, or zero, depending upon the material. These
different situations, are depicted in Fig. 14.2 and discussed below:
Case I: This refers to a situation, as shown in Fig. 14.2(a). One can
have a metal either when the conduction band is partially filled and the
balanced band is partially empty or when the conduction and valance
bands overlap. When there is overlap electrons from valence band
can easily move into the conduction band. This situation makes a
large number of electrons available for electrical conduction. When the
valence band is partially empty, electrons from its lower level can
move to higher level making conduction possible. Therefore, the
resistance of such materials is low or the conductivity is high.
Figure 14.2 Difference between energy bands of (a) metals, (b) insulators and (c)
semiconductors.
Case II: In this case, as shown in Fig. 14.2(b), a large band gap Eg
exists (Eg > 3 eV). There are no electrons in the conduction band, and
therefore no electrical conduction is possible. Note that the energy
gap is so large that electrons cannot be excited from the valence band
to the conduction band by thermal excitation. This is the case of
insulators.
Case III: This situation is shown in Fig. 14.2(c). Here a finite but small
band gap (Eg < 3 eV) exists. Because of the small band gap, at room
temperature some electrons from valence band can acquire enough
energy to cross the energy gap and enter the conduction band. These
electrons (though small in numbers) can move in the conduction band.
Hence, the resistance of semiconductors is not as high as that of the
insulators.
In this section we have made a broad classification of metals,
conductors and semiconductors. In the section which follows you will
learn the conduction process in semiconductors.
Thus, after such a jump, the hole is at site 2 and the site 1 has now an
electron. Therefore, apparently, the hole has moved from site 1 to site
2. Note that the electron originally set free [Fig. 14.5(a)] is not involved
in this process of hole motion. The free electron moves completely
independently as conduction electron and gives rise to an electron
current, Ie under an applied electric field. Remember that the motion
of hole is only a convenient way of describing the actual motion of
bound electrons, whenever there is an empty bond anywhere in the
crystal. Under the action of an electric field, these holes move towards
negative potential giving the hole current, Ih. The total current, I is thus
the sum of the electron current Ie and the hole current Ih:
I = Ie + Ih (14.2)
It may be noted that apart from the process of generation of
conduction electrons and holes, a simultaneous process of
recombination occurs in which the electrons recombine with the holes.
At equilibrium, the rate of generation is equal to the rate of
recombination of charge carriers. The recombination occurs due to an
electron colliding with a hole.
(a)
(b)
Figure 14.5 (a) Schematic model of generation of hole at site 1 and conduction
electron due to thermal energy at moderate temperatures. (b) Simplified
representation of possible thermal motion of a hole. The electron from the lower
left hand covalent bond (site 2) goes to the earlier hole site1, leaving a hole at its
site indicating an
apparent movement of the hole from site 1 to site 2.
When a small amount, say, a few parts per million (ppm), of a suitable
impurity is added to the pure semiconductor, the conductivity of the
semiconductor is increased manifold. Such materials are known as
extrinsic semiconductors or impurity semiconductors. The deliberate
addition of a desirable impurity is called doping and the impurity atoms
are called dopants. Such a material is also called a doped
semiconductor. The dopant has to be such that it does not distort the
original pure semiconductor lattice. It occupies only a very few of the
original semiconductor atom sites in the crystal. A necessary condition
to attain this is that the sizes of the dopant and the semiconductor
atoms should be nearly the same.
Figure 14.9 Energy bands of (a) n-type semiconductor at T > 0K, (b) p-
type semiconductor at T > 0K.
Example 14.2 Suppose a pure Si crystal has 5 1028 atoms m3. It is doped
by 1 ppm concentration of pentavalent As. Calculate the number of electrons
and holes. Given that ni =1.5 1016 m3.
Solution Note that thermally generated electrons (ni ~1016 m3) are negligibly
small as compared to those produced by doping.
Therefore, ne ND.
Since nenh = ni2, The number of holes
nh = (2.25 1032)/(5 1022)
~ 4.5 109 m3
The loss of electrons from the n-region and the gain of electron by the
p-region causes a difference of potential across the junction of the two
regions. The polarity of this potential is such as to oppose further flow
of carriers so that a condition of equilibrium exists. Figure 14.11 shows
the p-n junction at equilibrium and the potential across the junction.
The
n-material has lost electrons, and p material has acquired electrons.
The n material is thus positive relative to the p material. Since this
potential tends to prevent the movement of electron from the n region
into the p region, it is often called a barrier potential.
Example 14.3 Can we take one slab of p-type semiconductor and physically
join it to another n-type semiconductor to get p-n junction?
Solution No! Any slab, howsoever flat, will have roughness much larger than
the inter-atomic crystal spacing (~2 to 3 ) and hence continuous contact at
the atomic level will not be possible. The junction will behave as a discontinuity
for the flowing charge carriers.
Figure 14.13 (a) p-n junction diode under forward bias, (b) Barrier potential (1)
without battery, (2) Low battery voltage, and (3) High voltage battery.
Due to the applied voltage, electrons from n-side cross the depletion
region and reach p-side (where they are minority carries). Similarly,
holes from p-side cross the junction and reach the n-side (where they
are minority carries). This process under forward bias is known as
minority carrier injection. At the junction boundary, on each side, the
minority carrier concentration increases significantly compared to the
locations far from the junction.
When an external voltage (V) is applied across the diode such that n-
side is positive and p-side is negative, it is said to be reverse biased
[Fig.14.15(a)]. The applied voltage mostly drops across the depletion
region. The direction of applied voltage is same as the direction of
barrier potential. As a result, the barrier height increases and the
depletion region widens due to the change in the electric field. The
effective barrier height under reverse bias is (V0 + V), [Fig. 14.15(b)].
This suppresses the flow of electrons from n p and holes from p
n. Thus, diffusion current, decreases enormously compared to the
diode under forward bias.
Figure 14.15 (a) Diode under reverse bias, (b) Barrier potential under reverse bias.
his voltage is called the threshold voltage or cut-in voltage (~0.2V for
germanium diode and ~0.7 V for silicon diode).
For the diode in reverse bias, the current is very small (~A) and
almost remains constant with change in bias. It is called reverse
saturation current. However, for special cases, at very high reverse
bias (break down voltage), the current suddenly increases. This
special action of the diode is discussed later in Section 14.8. The
general purpose diode are not used beyond the reverse saturation
current region.
The above discussion shows that the p-n junction diode primerly
allows the flow of current only in one direction (forward bias). The
forward bias resistance is low as compared to the reverse bias
resistance. This property is used for rectification of ac voltages as
discussed in the next section. For diodes, we define a quantity called
dynamic resistance as the ratio of small change in voltage V to a
small change in current I:
(14.6)
Example 14.4 The V-I characteristic of a silicon diode is shown in the Fig.
14.17. Calculate the resistance of the diode at (a) ID = 15 mA and (b) VD =
10 V.
Figure 14.17
Solution Considering the diode characteristics as a straight line between I =
10 mA to I = 20 mA passing through the origin, we can calculate the resistance
using Ohms law.
(a) From the curve, at I = 20 mA, V = 0.8 V, I = 10 mA, V = 0.7 V
rfb = V/I = 0.1V/10 mA = 10
(b) From the curve at V = 10 V, I = 1 A,
Therefore,
rrb = 10 V/1A= 1.0 107
Figure 14.18 (a) Half-wave rectifier circuit, (b) Input ac voltage and output voltage
waveforms from the rectifier circuit.
Figure 14.20 (a) A full-wave rectifier with capacitor filter, (b) Input and
output voltage of rectifier in (a).
Zener diode is fabricated by heavily doping both p-, and n- sides of the
junction. Due to this, depletion region formed is very thin (<106 m)
and the electric field of the junction is extremely high (~5106 V/m)
even for a small reverse bias voltage of about 5V. The I-V
characteristics of a Zener diode is shown in Fig. 14.21(b). It is seen
that when the applied reverse bias voltage(V) reaches the breakdown
voltage (Vz) of the Zener diode, there is a large change in the current.
Note that after the breakdown voltage Vz, a large change in the
current can be produced by almost insignificant change in the reverse
bias voltage. In other words, Zener voltage remains constant, even
though current through the Zener diode varies over a wide range. This
property of the Zener diode is used for regulating supply voltages so
that they are constant.
Let us understand how reverse current suddenly increases at the
breakdown voltage. We know that reverse current is due to the flow of
electrons (minority carriers) from p n and holes from n p. As the
reverse bias voltage is increased, the electric field at the junction
becomes significant. When the reverse bias voltage V = Vz, then the
electric field strength is high enough to pull valence electrons from the
host atoms on the p-side which are accelerated to n-side. These
electrons account for high current observed at the breakdown. The
emission of electrons from the host atoms due to the high electric field
is known as internal field emission or field ionisation. The electric field
required for field ionisation is of the order of 106 V/m.
Figure 14.21 Zener diode, (a) symbol, (b) I-V characteristics.
Example 14.6 The current in the forward bias is known to be more (~mA) than
the current in the reverse bias (~A). What is the reason then to operate the
photodiodes in reverse bias?
Solution Consider the case of an n-type semiconductor. Obviously, the
majority carrier density (n) is considerably larger than the minority hole density
p (i.e., n >> p). On illumination, let the excess electrons and holes generated
be n and p, respectively:
n = n + n
p = p + p
Here n and p are the electron and hole concentrations* at any particular
illumination and n and p are carriers concentration when there is no
illumination. Remember n = p and n >> p. Hence, the fractional change in
the majority carriers (i.e., n/n) would be much less than that in the minority
carriers (i.e., p/p). In general, we can state that the fractional change due to
the photo-effects on the minority carrier dominated reverse bias current is
more easily measurable than the fractional change in the forward bias current.
Hence, photodiodes are preferably used in the reverse bias condition for
measuring light intensity.
* Note that, to create an e-h pair, we spend some energy (photoexcitation, thermal
excitation, etc.). Therefore when an electron and hole recombine the energy is
released in the form of light (radiative recombination) or heat (non-radiative
recombination). It depends on semiconductor and the method of fabrication of the
p-n junction. For the fabrication of LEDs, semiconductors like GaAs, GaAs-GaP
are used in which radiative recombination dominates.
When the diode is forward biased, electrons are sent from n p
(where they are minority carriers) and holes are sent from p n
(where they are minority carriers). At the junction boundary the
concentration of minority carriers increases compared to the
equilibrium concentration (i.e., when there is no bias). Thus at the
junction boundary on either side of the junction, excess minority
carriers are there which recombine with majority carriers near the
junction. On recombination, the energy is released in the form of
photons. Photons with energy equal to or slightly less than the band
gap are emitted. When the forward current of the diode is small, the
intensity of light emitted is small. As the forward current increases,
intensity of light increases and reaches a maximum. Further increase
in the forward current results in decrease of light intensity. LEDs are
biased such that the light emitting efficiency is maximum.
The V-I characteristics of a LED is similar to that of a Si junction diode.
But the threshold voltages are much higher and slightly different for
each colour. The reverse breakdown voltages of LEDs are very low,
typically around 5V. So care should be taken that high reverse
voltages do not appear across them.
LEDs that can emit red, yellow, orange, green and blue light are
commercially available. The semiconductor used for fabrication of
visible LEDs must at least have a band gap of 1.8 eV (spectral range
of visible light is from about 0.4 m to 0.7 m, i.e., from about 3 eV to
1.8 eV). The compound semiconductor Gallium Arsenide Phosphide
(GaAs1xPx) is used for making LEDs of different colours. GaAs0.6
P0.4 (Eg ~ 1.9 eV) is used for red LED. GaAs (Eg ~ 1.4 eV) is used for
making infrared LED. These LEDs find extensive use in remote
controls, burglar alarm systems, optical communication, etc. Extensive
research is being done for developing white LEDs which can replace
incandescent lamps.
Figure 14.24 (a) Typical p-n junction solar cell; (b) Cross-sectional view.
Example 14.7 Why are Si and GaAs are preferred materials for solar cells?
Solution The solar radiation spectrum received by us is shown in Fig. 14.26.
Figure 14.26
The maxima is near 1.5 eV. For photo-excitation, h > Eg. Hence,
semiconductor with band gap ~1.5 eV or lower is likely to give better solar
conversion efficiency. Silicon has Eg ~ 1.1 eV while for GaAs it is ~1.53 eV. In
fact, GaAs is better (in spite of its higher band gap) than Si because of its
relatively higher absorption coefficient. If we choose materials like CdS or
CdSe (Eg ~ 2.4 eV), we can use only the high energy component of the solar
energy for photo-conversion and a significant part of energy will be of no use.
The question arises: why we do not use material like PbS (Eg ~ 0.4 eV) which
satisfy the condition h > Eg for maxima corresponding to the solar radiation
spectra? If we do so, most of the solar radiation will be absorbed on the top-
layer of solar cell and will not reach in or near the depletion region. For
effective electron-hole separation, due to the junction field, we want the photo-
generation to occur in the junction region only.
14.9 JUNCTION TRANSISTOR
First we shall see what gives the transistor its amplifying capabilities.
The transistor works as an amplifier, with its emitter-base junction
forward biased and the base-collector junction reverse biased. This
situation is shown in Fig. 14.28, where VCC and VEE are used for
creating the respective biasing. When the transistor is biased in this
way it is said to be in active state.We represent the voltage between
emitter and base as VEB and that between the collector and the base
as VCB. In Fig. 14.28, base is a common terminal for the two power
supplies whose other terminals are connected to emitter and collector,
respectively. So the two power supplies are represented as VEE, and
VCC, respectively. In circuits, where emitter is the common terminal,
the power supply between the base and the emitter is represented as
VBB and that between collector and emitter as VCC.
Let us see now the paths of current carriers in the transistor with
emitter-base junction forward biased and base-collector junction
reverse biased. The heavily doped emitter has a high concentration of
majority carriers, which will be holes in a p-n-p transistor and electrons
in an n-p-n transistor. These majority carriers enter the base region in
large numbers. The base is thin and lightly doped. So the majority
carriers there would be few. In a p-n-p transistor the majority carriers
in the base are electrons since base is of n-type semiconductor. The
large number of holes entering the base from the emitter swamps the
small number of electrons there. As the base collector-junction is
reverse-biased, these holes, which appear as minority carriers at the
junction, can easily cross the junction and enter the collector. The
holes in the base could move either towards the base terminal to
combine with the electrons entering from outside or cross the junction
to enter into the collector and reach the collector terminal. The base is
made thin so that most of the holes find themselves near the reverse-
biased base-collector junction and so cross the junction instead of
moving to the base terminal.
Figure 14.28 Bias Voltage applied on: (a) p-n-p transistor and (b) n-p-n transistor.
Figure 14.29 Circuit arrangement for studying the input and output characteristics
of n-p-n transistor in CE configuration.
The linear segments of both the input and output characteristics can
be used to calculate some important ac parameters of transistors as
shown below.
(i) Input resistance (ri): This is defined as the ratio of change in
base-emitter voltage (VBE) to the resulting change in base current
(IB) at constant collector-emitter voltage (VCE). This is dynamic (ac
resistance) and as can be seen from the input characteristic, its value
varies with the operating current in the transistor:
(14.8)
The value of ri can be anything from a few hundreds to a few
thousand ohms.
(ii) Output resistance (ro): This is defined as the ratio of change in
collector-emitter voltage (VCE) to the change in collector current (IC)
at a constant base current IB.
(14.9)
The output characteristics show that initially for very small values of
VCE, IC increases almost linearly. This happens because the base-
collector junction is not reverse biased and the transistor is not in
active state. In fact, the transistor is in the saturation state and the
current is controlled by the supply voltage VCC (=VCE) in this part of
the characteristic. When VCE is more than that required to reverse bias
the base-collector junction, IC increases very little with VCE. The
reciprocal of the slope of the linear part of the output characteristic
gives the values of ro. The output resistance of the transistor is mainly
controlled by the bias of the base-collector junction. The high
magnitude of the output resistance (of the order of 100 k) is due to
the reverse-biased state of this diode. This also explains why the
resistance at the initial part of the characteristic, when the transistor is
in saturation state, is very low.
(iii) Current amplification factor (): This is defined as the ratio of
the change in collector current to the change in base current at a
constant collector-emitter voltage (VCE) when the transistor is in active
state.
(14.10)
This is also known as small signal current gain and its value is very
large.
(14.11)
Since IC increases with IB almost linearly and IC = 0 when IB = 0, the
values of both dc and ac are nearly equal. So, for most calculations
dc can be used. Both ac and dc vary with VCE and IB (or IC) slightly.
Example 14.8 From the output characteristics shown in Fig. 14.30(b),
calculate the values of ac and dc of the transistor when VCE is 10 V and IC
= 4.0 mA.
Solution
,
For determining ac and dc at the stated values of VCE and IC one can
proceed as follows. Consider any two characteristics for two values of IB which
lie above and below the given value of IC . Here IC = 4.0 mA. (Choose
characteristics for IB= 30 and 20 A.) At VCE = 10 V we read the two values of
IC from the graph. Then
IB = (30 20) A = 10 A, IC = (4.5 3.0) mA = 1.5 mA
Therefore, ac = 1.5 mA/ 10 A = 150
For determining dc, either estimate the value of IB corresponding to
IC = 4.0 mA at VCE = 10 V or calculate the two values of dc for the two
characteristics chosen and find their mean.
Therefore, for IC = 4.5 mA and IB = 30 A,
dc = 4.5 mA/ 30 A = 150
and for IC = 3.0 mA and IB = 20 A
dc =3.0 mA / 20 A = 150
Hence, dc =(150 + 150) /2 = 150
We shall treat VBB as the dc input voltage Vi and VCE as the dc output
voltage VO. So, we have
Vi = IBRB + VBE and
Vo = VCC ICRC.
Let us see how Vo changes as Vi increases from zero onwards. In the
case of Si transistor, as long as input Vi is less than 0.6 V, the
transistor will be in cut off state and current IC will be zero.
Hence Vo = VCC
When Vi becomes greater than 0.6 V the transistor is in active state
with some current IC in the output path and the output Vo decrease as
the term ICRC increases. With increase of Vi , IC increases almost
linearly and so Vo decreases linearly till its value becomes less than
about 1.0 V.
Beyond this, the change becomes non linear and transistor goes into
saturation state. With further increase in Vi the output voltage is found
to decrease further towards zero though it may never become zero. If
we plot the Vo vs Vi curve, [also called the transfer characteristics of
the base-biased transistor (Fig. 14.31(b)], we see that between cut off
state and active state and also between active state and saturation
state there are regions of non-linearity showing that the transition from
cutoff state to active state and from active state to saturation state are
not sharply defined.
Figure 14.31 (a) Base-biased transistor in CE configuration, (b) Transfer
characteristic.
If the VBB voltage has a fixed value corresponding to the mid point of
the active region, the circuit will behave as a CE amplifier with voltage
gain Vo/ Vi. We can express the voltage gain AV in terms of the
resistors in the circuit and the current gain of the transistor as follows.
We have, Vo = VCC ICRC
Therefore, Vo = 0 RC IC
Similarly, from Vi = IBRB + VBE
Vi = RB IB + VBE
But VBE is negligibly small in comparison to IBRB in this circuit.
= r IB
The change in IB causes a change in Ic. We define a parameter ac,
which is similar to the dc defined in Eq. (14.11), as
(14.17)
which is also known as the ac current gain Ai. Usually ac is close to
dc in the linear region of the output characteristics.
The change in Ic due to a change in IB causes a change in VCE and
the voltage drop across the resistor RL because VCC is fixed.
These changes can be given by Eq. (14.15) as
VCC = VCE + RL IC = 0
or VCE = RL IC
The change in VCE is the output voltage v0. From Eq. (14.10), we get
v0 = VCE = ac RL IB
The voltage gain of the amplifier is
(14.18)
The negative sign represents that output voltage is opposite with
phase with the input voltage.
Example 14.9 In Fig. 14.31(a), the VBB supply can be varied from 0V to 5.0 V.
The Si transistor has dc = 250 and RB = 100 k, RC = 1 K, VCC = 5.0V.
Assume that when the transistor is saturated, VCE = 0V and VBE = 0.8V.
Calculate (a) the minimum base current, for which the transistor will reach
saturation. Hence, (b) determine V1 when the transistor is switched on. (c)
find the ranges of V1 for which the transistor is switched off and switched on.
Solution
Given at saturation VCE = 0V, VBE = 0.8V
VCE = VCC ICRC
IC = VCC/RC = 5.0V/1.0k = 5.0 mA
Therefore IB = IC/ = 5.0 mA/250 = 20A
The input voltage at which the transistor will go into saturation is given by
VIH = VBB = IBRB +VBE
= 20A 100 k + 0.8V = 2.8V
The value of input voltage below which the transistor remains cutoff is given by
VIL = 0.6V, VIH = 2.8V
Between 0.0V and 0.6V, the transistor will be in the switched off state.
Between 2.8V and 5.0V, it will be in switched on state.
Note that the transistor is in active state when IB varies from 0.0mA to 20mA.
In this range, IC = IB is valid. In the saturation range,
IC IB.
Example 14.10 For a CE transistor amplifier, the audio signal voltage across
the collector resistance of 2.0 k is 2.0 V. Suppose the current amplification
factor of the transistor is 100, What should be the value of RB in series with
VBB supply of 2.0 V if the dc base current has to be 10 times the signal
current. Also calculate the dc drop across the collector resistance. (Refer to
Fig. 14.33).
Solution The output ac voltage is 2.0 V. So, the ac collector current iC =
2.0/2000 = 1.0 mA. The signal current through the base is, therefore given by
iB = iC / = 1.0 mA/100 = 0.010 mA. The dc base current has to be 10 0.010
= 0.10 mA.
From Eq.14.16, RB = (VBB - VBE ) /IB. Assuming VBE = 0.6 V,
RB = (2.0 0.6 )/0.10 = 14 k.
The dc collector current IC = 1000.10 = 10 mA.
(14.20)
In the circuit of Fig. 14.33(b), the tank or tuned circuit is connected in
the collector side. Hence, it is known as tuned collector oscillator. If
the tuned circuit is on the base side, it will be known as tuned base
oscillator. There are many other types of tank circuits (say RC) or
feedback circuits giving different types of oscillators like Colpitts
oscillator, Hartley oscillator, RC-oscillator.
(b)
Figure 14.35
(a) Logic symbol,
(b) Truth table of
NOT gate.
(b)
Apart from carrying out the above mathematical logic operation, this
gate can be used for modifying the pulse waveform as explained in
the following example.
Example 14.11 Justify the output waveform (Y) of the OR gate for the
following inputs A and B given in Fig. 14.37.
Solution Note the following:
At t < t1; A = 0, B = 0; Hence Y = 0
For t1 to t2; A = 1, B = 0; Hence Y = 1
For t2 to t3; A = 1, B = 1; Hence Y = 1
For t3 to t4; A = 0, B = 1; Hence Y = 1
For t4 to t5; A = 0, B = 0; Hence Y = 0
For t5 to t6; A = 1, B = 0; Hence Y = 1
For t > t6; A = 0, B = 1; Hence Y = 1
Therefore the waveform Y will be as shown in the Fig. 14.37.
Figure 14.37
(b)
Figure 14.38 (a) Logic symbol, (b) Truth table of AND gate.
(b)
Figure 14.40 (a) Logic symbol, (b) Truth table of NAND gate.
Example 14.13 Sketch the output Y from a NAND gate having inputs A and B
given below:
Solution
For t < t1; A = 1, B = 1; Hence Y = 0
For t1 to t2; A = 0, B = 0; Hence Y = 1
For t2 to t3; A = 0, B = 1; Hence Y = 1
For t3 to t4; A = 1, B = 0; Hence Y = 1
For t4 to t5; A = 1, B = 1; Hence Y = 0
For t5 to t6; A = 0, B = 0; Hence Y = 1
For t > t6; A = 0, B = 1; Hence Y = 1
Figure 14.41
(b)
Figure 14 .42 (a) Logic symbol, (b) Truth table of NOR gate.
NOR gates are considered as universal gates because you can obtain
all the gates like AND, OR, NOT by using only NOR gates (Exercises
14.18 and 14.19).
The Integrated Chip (IC) is at the heart of all computer systems. In fact ICs are
found in almost all electrical devices like cars, televisions, CD players, cell
phones etc. The miniaturisation that made the modern personal computer
possible could never have happened without the IC. ICs are electronic devices
that contain many transistors, resistors, capacitors, connecting wires all in
one package. You must have heard of the microprocessor. The
microprocessor is an IC that processes all information in a computer, like
keeping track of what keys are pressed, running programmes, games etc. The
IC was first invented by Jack Kilky at Texas Instruments in 1958 and he was
awarded Nobel Prize for this in 2000. ICs are produced on a piece of
semiconductor crystal (or chip) by a process called photolithography. Thus, the
entire Information Technology (IT) industry hinges on semiconductors. Over
the years, the complexity of ICs has increased while the size of its features
continued to shrink. In the past five decades, a dramatic miniaturisation in
computer technology has made modern day computers faster and smaller. In
the 1970s, Gordon Moore, co-founder of INTEL, pointed out that the memory
capacity of a chip (IC) approximately doubled every one and a half years. This
is popularly known as Moores law. The number of transistors per chip has
risen exponentially and each year computers are becoming more powerful, yet
cheaper than the year before. It is intimated from current trends that the
computers available in 2020 will operate at 40 GHz (40,000 MHz) and would
be much smaller, more efficient and less expensive than present day
computers. The explosive growth in the semiconductor industry and computer
technology is best expressed by a famous quote from Gordon Moore: If the
auto industry advanced as rapidly as the semiconductor industry, a Rolls
Royce would get half a million miles per gallon, and it would be cheaper to
throw it away than to park it.
SUMMARY
1. Semiconductors are the basic materials used in the present solid state
electronic devices like diode, transistor, ICs, etc.
2. Lattice structure and the atomic structure of constituent elements decide
whether a particular material will be insulator, metal or semiconductor.
3. Metals have low resistivity (102 to 108 m), insulators have very high
resistivity (>108 m1), while semiconductors have intermediate values of
resistivity.
4. Semiconductors are elemental (Si, Ge) as well as compound (GaAs, CdS,
etc.).
5. Pure semiconductors are called intrinsic semiconductors. The presence of
charge carriers (electrons and holes) is an intrinsic property of the material
and these are obtained as a result of thermal excitation. The number of
electrons (ne) is equal to the number of holes (nh ) in intrinsic conductors.
Holes are essentially electron vacancies with an effective positive charge.
6. The number of charge carriers can be changed by doping of a suitable
impurity in pure semiconductors. Such semiconductors are known as extrinsic
semiconductors. These are of two types (n-type and p-type).
7. In n-type semiconductors, ne >> nh while in p-type semiconductors nh >>
ne.
8. n-type semiconducting Si or Ge is obtained by doping with pentavalent
atoms (donors) like As, Sb, P, etc., while p-type Si or Ge can be obtained by
doping with trivalent atom (acceptors) like B, Al, In etc.
9. nenh = ni2 in all cases. Further, the material possesses an overall charge
neutrality.
10. There are two distinct band of energies (called valence band and
conduction band) in which the electrons in a material lie. Valence band
energies are low as compared to conduction band energies. All energy levels
in the valence band are filled while energy levels in the conduction band may
be fully empty or partially filled. The electrons in the conduction band are free
to move in a solid and are responsible for the conductivity. The extent of
conductivity depends upon the energy gap (Eg) between the top of valence
band (EV) and the bottom of the conduction band EC. The electrons from
valence band can be excited by heat, light or electrical energy to the
conduction band and thus, produce a change in the current flowing in a
semiconductor.
11. For insulators Eg > 3 eV, for semiconductors Eg is 0.2 eV to 3 eV, while for
metals Eg 0.
12. p-n junction is the key to all semiconductor devices. When such a junction
is made, a depletion layer is formed consisting of immobile ion-cores devoid
of their electrons or holes. This is responsible for a junction potential barrier.
13. By changing the external applied voltage, junction barriers can be
changed. In forward bias (n-side is connected to negative terminal of the
battery and p-side is connected to the positive), the barrier is decreased while
the barrier increases in reverse bias. Hence, forward bias current is more (mA)
while it is very small (A) in a p-n junction diode.
14. Diodes can be used for rectifying an ac voltage (restricting the ac voltage
to one direction). With the help of a capacitor or a suitable filter, a dc voltage
can be obtained.
15. There are some special purpose diodes.
16. Zener diode is one such special purpose diode. In reverse bias, after a
certain voltage, the current suddenly increases (breakdown voltage) in a Zener
diode. This property has been used to obtain voltage regulation.
17. p-n junctions have also been used to obtain many photonic
or optoelectronic devices where one of the participating entity is photon: (a)
Photodiodes in which photon excitation results in a change of reverse
saturation current which helps us to measure light intensity; (b) Solar cells
which convert photon energy into electricity; (c) Light Emitting Diode and Diode
Laser in which electron excitation by a bias voltage results in the generation of
light.
18. Transistor is an n-p-n or p-n-p junction device. The central block (thin and
lightly doped) is called Base while the other electrodes are Emitter and
Collectors. The emitter-base junction is forward biased while collector-base
junction is reverse biased.
19. The transistors can be connected in such a manner that either C or E or B
is common to both the input and output. This gives the three configurations in
which a transistor is used: Common Emitter (CE), Common Collector (CC) and
Common Base (CB). The plot between IC and VCE for fixed IB is called output
characteristics while the plot between IB and VBE with fixed VCE is called input
characteristics. The important transistor parameters for CE-configuration are:
input resistance,
output resistance,
Exercises
(b) Electrons are minority carriers and pentavalent atoms are the
dopants.
(c) Holes are minority carriers and pentavalent atoms are the
dopants.
(d) Holes are majority carriers and trivalent atoms are the dopants.
(b) is high at high and low frequencies and constant in the middle
frequency range.
(c) is low at high and low frequencies and constant at mid
frequencies.
(d) None of the above.
14.8 In half-wave rectification, what is the output frequency if the
input frequency is 50 Hz. What is the output frequency of a full-
wave rectifier for the same input frequency.
14.9 For a CE-transistor amplifier, the audio signal voltage across
the collected resistance of 2 k is 2 V. Suppose the current
amplification factor of the transistor is 100, find the input signal
voltage and base current, if the base resistance is 1 k.
14.10 Two amplifiers are connected one after the other in series
(cascaded). The first amplifier has a voltage gain of 10 and the
second has a voltage gain of 20. If the input signal is 0.01 volt,
calculate the output ac signal.
14.11 A p-n photodiode is fabricated from a semiconductor with
band gap of 2.8 eV. Can it detect a wavelength of 6000 nm?
Additional Exercises
14.12 The number of silicon atoms per m3 is 5 1028. This is
doped simultaneously with 5 1022 atoms per m3 of Arsenic and 5
1020 per m3 atoms of Indium. Calculate the number of electrons
and holes. Given that ni = 1.5 1016 m3. Is the material n-type or
p-type?
where n0 is a constant.
14.14 In a p-n junction diode, the current I can be expressed as
Figure 14.44
14.16 Write the truth table for a NAND gate connected as given
in Fig. 14.45.
Figure 14.45
Hence identify the exact logic operation carried out by this circuit.
14.17 You are given two circuits as shown in Fig. 14.46, which
consist of NAND gates. Identify the logic operation carried out by
the two circuits.
Figure 14.46
14.18 Write the truth table for circuit given in Fig. 14.47 below
consisting of NOR gates and identify the logic operation (OR,
AND, NOT) which this circuit is performing.
Figure 14.47
Figure 14.48
Chapter Six
6.1 Introduction
6.2 Notions of work and kinetic energy : The work-energy theorem
6.3 Work
6.4 Kinetic energy
6.5 Work done by a variable force
6.6 The work-energy theorem for a variable force
6.7 The concept of potential energy
6.8 The conservation of mechanical energy
6.9 The potential energy of a spring
6.10 Various forms of energy : the law of conservation of energy
6.11 Power
6.12 Collisions
Summary
Points to ponder
Exercises
Additional exercises
Appendix 6.1
6.1 Introduction
The terms work, energy and power are frequently used in everyday
language. A farmer ploughing the field, a construction worker carrying
bricks, a student studying for a competitive examination, an artist
painting a beautiful landscape, all are said to be working. In physics,
however, the word Work covers a definite and precise meaning.
Somebody who has the capacity to work for 14-16 hours a day is said
to have a large stamina or energy. We admire a long distance runner
for her stamina or energy. Energy is thus our capacity to do work. In
Physics too, the term energy is related to work in this sense, but as
said above the term work itself is defined much more precisely. The
word power is used in everyday life with different shades of meaning.
In karate or boxing we talk of powerful punches. These are delivered
at a great speed. This shade of meaning is close to the meaning of the
word power used in physics. We shall find that there is at best a
loose correlation between the physical definitions and the
physiological pictures these terms generate in our minds. The aim of
this chapter is to develop an understanding of these three physical
quantities. Before we proceed to this task, we need to develop a
mathematical prerequisite, namely the scalar product of two vectors.
(6.1b)
From the definition of scalar product and (Eq. 6.1b) we have :
(i)
Or, (6.1c)
since A.A = |A ||A| cos 0 = A2.
(ii) A.B = 0, if A and B are perpendicular.
Answer F.d =
= 3 (5) + 4 (4) + ( 5) (3)
= 16 unit
Now F.F =
= 9 + 16 + 25
= 50 unit
and d.d = d2 =
= 25 + 16 + 9
= 50 unit
cos = ,
= cos1 0.32 t
Fig. 6.1 (a) The scalar product of two vectors A and B is a scalar : A.B = A Bcos .
(b) B cos is the projection of B onto A. (c) A cos is the projection of A onto B.
(6.2a)
where the last step follows from Newtons Second Law. We can
generalise Eq. (6.1) to three dimensions by employing vectors
v2 u2 = 2 a.d
Once again multiplying both sides by m/2 , we obtain
(6.2b)
The above equation provides a motivation for the definitions of work
and kinetic energy. The left side of the equation is the difference in the
quantity half the mass times the square of the speed from its initial
value to its final value. We call each of these quantities the kinetic
energy, denoted by K. The right side is a product of the displacement
and the component of the force along the displacement. This quantity
is called work and is denoted by W. Eq. (6.2b) is then
Kf Ki = W (6.3)
where Ki and Kf are respectively the initial and final kinetic energies of
the object. Work refers to the force and the displacement over which it
acts. Work is done by a force on the body over a certain
displacement.
Equation (6.2) is also a special case of the work-energy (WE) theorem
: The change in kinetic energy of a particle is equal to the work
done on it by the net force. We shall generalise the above derivation
to a varying force in a later section.
Example 6.2 It is well known that a raindrop falls under the influence of the
downward gravitational force and the opposing resistive force. The latter is
known to be proportional to the speed of the drop but is otherwise undetermined.
Consider a drop of mass 1.00 g falling from a height 1.00 km. It hits the ground
with a speed of 50.0 m s-1. (a) What is the work done by the gravitational force ?
What is the work done by the unknown resistive force?
= 1.25 J
where we have assumed that the drop is initially at rest.
Assuming that g is a constant with a value 10 m/s2, the work done by
the gravitational force is,
Wg = mgh
= 10-3 10 103
= 10.0 J
(b) From the work-energy theorem
where Wr is the work done by the resistive force on the raindrop. Thus
Wr = K Wg
= 1.25 10
= 8.75 J
is negative. t
6.3 Work
As seen earlier, work is related to force and the displacement over
which it acts. Consider a constant force F acting on an object of mass
m. The object undergoes a displacement d in the positive x-direction
as shown in Fig. 6.2.
Fig. 6.2 An object undergoes a displacement d under the influence of the force F.
Answer Work done on the cycle by the road is the work done by the
stopping (frictional) force on the cycle due to the road.
(a) The stopping force and the displacement make an angle of 180o (
rad) with each other. Thus, work done by the road,
Wr = Fd cos
= 200 10 cos
= 2000 J
It is this negative work that brings the cycle to a halt in accordance
with WE theorem.
(b) From Newtons Third Law an equal and opposite force acts on the
road due to the cycle. Its magnitude is 200 N. However, the road
undergoes no displacement. Thus, work done by cycle on the road is
zero. t
The lesson of Example 6.3 is that though the force on a body A
exerted by the body B is always equal and opposite to that on B by A
(Newtons Third Law); the work done on A by B is not necessarily
equal and opposite to the work done on B by A.
(6.5)
Kinetic energy is a scalar quantity. The kinetic energy of an object is a
measure of the work an object can do by the virtue of its motion. This
notion has been intuitively known for a long time.
Table 6.2 Typical kinetic energies (K)
The kinetic energy of a fast flowing stream has been used to grind
corn. Sailing ships employ the kinetic energy of the wind. Table 6.2
lists the kinetic energies for various objects.
Answer The initial kinetic energy of the bullet is mv2/2 = 1000 J. It has
a final kinetic energy of 0.11000 = 100 J. If vf is the emergent speed
of the bullet,
= 63.2 m s1
The speed is reduced by approximately 68% (not 90%). t
(6.6)
where the summation is from the initial position xi to the final position
xf.
If the displacements are allowed to approach zero, then the number of
terms in the sum increases without limit, but the sum approaches a
definite value equal to the area under the curve in Fig. 6.3(b). Then
the work done is
lim
(6.7)
where lim stands for the limit of the sum when x tends to zero.
Thus, for a varying force the work done can be expressed as a definite
integral of force over displacement (see also Appendix 3.1).
Fig. 6.3(a)
Fig. 6.3 (a) The shaded rectangle represents the work done by the varying force
F(x), over the small displacement x, W = F(x) x. (b) adding the areas of all the
rectangles we find that for x 0, the area under the curve is exactly equal to the
work done by F(x).
Example 6.5 A woman pushes a trunk on a railway platform which has a rough
surface. She applies a force of 100 N over a distance of 10 m. Thereafter, she
gets progressively tired and her applied force reduces linearly with distance to
50 N. The total distance through which the trunk has been moved is 20 m. Plot
the force applied by the woman and the frictional force, which is 50 N versus
displacement. Calculate the work done by the two forces over 20 m.
Answer
Fig. 6.4 Plot of the force F applied by the woman and the opposing frictional force f
versus displacement.
= 1000 + 750
= 1750 J
The work done by the frictional force is
Wf area of the rectangle AGHI
Wf = (50) 20
= 1000 J
The area on the negative side of the force axis has a negative sign. t
Thus
dK = Fdx
Integrating from the initial position (x i ) to final position ( x f ), we have
or (6.8a)
From Eq. (6.7), it follows that
Kf Ki = W (6.8b)
Thus, the WE theorem is proved for a variable force.
While the WE theorem is useful in a variety of problems, it does not, in
general, incorporate the complete dynamical information of Newtons
second law. It is an integral form of Newtons second law. Newtons
second law is a relation between acceleration and force at any instant
of time. Work-energy theorem involves an integral over an interval of
time. In this sense, the temporal (time) information contained in the
statement of Newtons second law is integrated over and is not
available explicitly. Another observation is that Newtons second law
for two or three dimensions is in vector form whereas the work-energy
theorem is in scalar form. In the scalar form, information with respect
to directions contained in Newtons second law is not present.
Example 6.6 A block of mass m = 1 kg, moving on a horizontal surface with
speed vi = 2 ms1 enters a rough patch ranging from x = 0.10 m to x = 2.01 m.
The retarding force Fr on the block in this range is inversely proportional to x
over this range,
= 2 0.5 ln (20.1)
= 2 1.5 = 0.5 J
Here, note that ln is a symbol for the natural logarithm to the base e
and not the logarithm to the base 10 [ln X = loge X = 2.303 log10 X]. t
m v2 = m g h
which shows that the gravitational potential energy of the object at
height h, when the object is released, manifests itself as kinetic energy
of the object on reaching the ground.
* The variation of g with height is discussed in Chapter 8 on Gravitation.
Physically, the notion of potential energy is applicable only to the class
of forces where work done against the force gets stored up as
energy. When external constraints are removed, it manifests itself as
kinetic energy. Mathematically, (for simplicity, in one dimension) the
potential energy V(x) is defined if the force F(x) can be written as
Fig. 6.5 The conversion of potential energy to kinetic energy for a ball of mass m
dropped from a height H.
The total mechanical energies E0, Eh, and EH of the ball at the
indicated heights zero (ground level), h and H, are
EH = mgH (6.11 a)
(6.11 b)
E0 = (1/2) mvf2 (6.11 c)
The constant force is a special case of a spatially dependent force
F(x). Hence, the mechanical energy is conserved. Thus
EH = E0
or,
a result that was obtained in section 3.7 for a freely falling body.
Further,
EH = Eh
which implies,
(6.11 d)
and is a familiar result from kinematics.
At the height H, the energy is purely potential. It is partially converted
to kinetic at height h and is fully kinetic at ground level. This illustrates
the conservation of mechanical energy.
Fig. 6.6
Answer (i) There are two external forces on the bob : gravity and the
tension (T) in the string. The latter does no work since the
displacement of the bob is always normal to the string. The potential
energy of the bob is thus associated with the gravitational force only.
The total mechanical energy E of the system is conserved. We take
the potential energy of the system to be zero at the lowest point A.
Thus, at A :
(6.12)
or,
(ii) It is clear from Eq. (6.14)
At B, the energy is
Equating this to the energy at A and employing the result from (i),
namely ,
At point C, the string becomes slack and the velocity of the bob is
horizontal and to the left. If the connecting string is cut at this instant,
the bob will execute a projectile motion with horizontal projection akin
to a rock kicked horizontally from the edge of a cliff. Otherwise the bob
will continue on its circular path and complete the revolution. t
(6.16)
Fig. 6.7 Illustration of the spring force with a block attached to the free end of the
spring. (a) The spring force Fs is zero when the displacement x from the
equilibrium position is zero. (b) For the stretched spring x > 0 and Fs < 0 (c) For
the compressed spring x < 0 and Fs > 0.(d) The plot of Fs versus x. The area of
the shaded triangle represents the work done by the spring force. Due to the
(6.17)
Thus the work done by the spring force depends only on the end
points. Specifically, if the block is pulled from xi and allowed to return
to xi ;
= 0 (6.18)
The work done by the spring force in a cyclic process is zero. We
have explicitly demonstrated that the spring force (i) is position
dependent only as first stated by Hooke, (Fs = kx); (ii) does work
which only depends on the initial and final positions, e.g. Eq. (6.17).
Thus, the spring force is a conservative force.
We define the potential energy V(x) of the spring to be zero when
block and spring system is in the equilibrium position. For an
extension (or compression) x the above analysis suggests that
(6.19)
You may easily verify that dV/dx = k x, the spring force. If the block
of mass m in Fig. 6.7 is extended to xm and released from rest, then
its total mechanical energy at any arbitrary point x, where x lies
between xm and + xm, will be given by
or
Note that k/m has the dimensions of [T-2] and our equation is
dimensionally correct. The kinetic energy gets converted to potential
energy and vice versa, however, the total mechanical energy remains
constant. This is graphically depicted in Fig. 6.8.
Fig. 6.8 Parabolic plots of the potential energy V and kinetic energy K of a block
attached to a spring obeying Hookes law. The two plots are complementary, one
decreasing as the other increases. The total mechanical energy E = K + V remains
constant.
Example 6.8 To simulate car accidents, auto manufacturers study the collisions
of moving cars with mounted springs of different spring constants. Consider a
typical simulation with a car of mass 1000 kg moving with a speed 18.0 km/h on
a smooth road and colliding with a horizontally mounted spring of spring
constant 6.25 103 N m1. What is the maximum compression of the spring ?
Answer At maximum compression the kinetic energy of the car is
converted entirely into the potential energy of the spring.
The kinetic energy of the moving car is
K = 1.25 104 J
where we have converted 18 km h1 to 5 m s1 [It is useful to
remember that 36 km h1 = 10 m s1]. At maximum compression xm,
the potential energy V of the spring is equal to the kinetic energy K of
the moving car from the principle of conservation of mechanical
energy.
= 1.25 104 J
We obtain
xm = 2.00 m
We note that we have idealised the situation. The spring is
considered to be massless. The surface has been considered to
possess negligible friction. t
We conclude this section by making a few remarks on conservative
forces.
(i) Information on time is absent from the above discussions. In the
example considered above, we can calculate the compression, but not
the time over which the compression occurs. A solution of Newtons
Second Law for this system is required for temporal information.
(ii) Not all forces are conservative. Friction, for example, is a non-
conservative force. The principle of conservation of energy will have to
be modified in this case. This is illustrated in Example 6.9.
(iii) The zero of the potential energy is arbitrary. It is set according to
convenience. For the spring force we took V(x) = 0, at x = 0, i.e. the
unstretched spring had zero potential energy. For the constant
gravitational force mg, we took V = 0 on the earths surface. In a later
chapter we shall see that for the force due to the universal law of
gravitation, the zero is best defined at an infinite distance from the
gravitational source. However, once the zero of the potential energy is
fixed in a given discussion, it must be consistently adhered to
throughout the discussion. You cannot change horses in midstream !
Example 6.9 Consider Example 6.8 taking the coefficient of friction, , to be 0.5
and calculate the maximum compression of the spring.
Answer In presence of friction, both the spring force and the frictional
force act so as to oppose the compression of the spring as shown in
Fig. 6.9.
We invoke the work-energy theorem, rather than the conservation of
mechanical energy.
The change in kinetic energy is
K = Kf Ki
The work done by the net force is
Equating we have
6.10.1 Heat
We have seen that the frictional force is not a conservative force.
However, work is associated with the force of friction, Example 6.5. A
block of mass m sliding on a rough horizontal surface with speed v0
comes to a halt over a distance x0. The work done by the force of
kinetic friction f over x0 is fx0. By the work-energy theorem
6.11 Power
Often it is interesting to know not only the work done on an object, but
also the rate at which this work is done. We say a person is physically
fit if he not only climbs four floors of a building but climbs them fast.
Power is defined as the time rate at which work is done or energy is
transferred.
The average power of a force is defined as the ratio of the work, W, to
the total time t taken
(6.21)
The work dW done by a force F for a displacement dr is dW = F.dr.
The instantaneous power can also be expressed as
= F.v (6.22)
where v is the instantaneous velocity when the force is F.
Power, like work and energy, is a scalar quantity. Its dimensions are
[ML2T3]. In the SI, its unit is called a watt (W). The watt is 1 J s1. The
unit of power is named after James Watt, one of the innovators of the
steam engine in the eighteenth century.
There is another unit of power, namely the horse-power (hp)
1 hp = 746 W
This unit is still used to describe the output of automobiles,
motorbikes, etc.
We encounter the unit watt when we buy electrical goods such as
bulbs, heaters and refrigerators. A 100 watt bulb which is on for 10
hours uses 1 kilowatt hour (kWh) of energy.
100 (watt) 10 (hour)
= 1000 watt hour
=1 kilowatt hour (kWh)
= 103 (W) 3600 (s)
= 3.6 106 J
Our electricity bills carry the energy consumption in units of kWh. Note
that kWh is a unit of energy and not of power.
Example 6.11 An elevator can carry a maximum load of 1800 kg
(elevator + passengers) is moving up with a constant speed of 2 m s
1. The frictional force opposing the motion is 4000 N. Determine the
minimum power delivered by the motor to the elevator in watts as well
as in horse power.
Answer The downward force on the elevator is
F = m g + Ff = (1800 10) + 4000 = 22000 N
The motor must supply enough power to balance this force. Hence,
P = F. v = 22000 2 = 44000 W = 59 hp t
6.12 Collisions
In physics we study motion (change in position). At the same time, we
try to discover physical quantities, which do not change in a physical
process. The laws of momentum and energy conservation are typical
examples. In this section we shall apply these laws to a commonly
encountered phenomena, namely collisions. Several games such as
billiards, marbles or carrom involve collisions.We shall study the
collision of two masses in an idealised form.
Consider two masses m1 and m2. The particle m1 is moving with
speed v1i , the subscript i implying initial. We can cosider m2 to be at
rest. No loss of generality is involved in making such a selection. In
this situation the mass m1 collides with the stationary mass m2 and
this is depicted in Fig. 6.10.
(6.23)
The loss in kinetic energy on collision is
the bounce (by using ). Hence you will get the coefficient of
restitution.
Now take the big ball and a small ball and hold them in your hands one over the
other, with the heavier ball below the lighter one, as shown here. Drop them
together, taking care that they remain together while falling, and see what
happens. You will find that the heavier ball rises less than when it was dropped
alone, while the lighter one shoots up to about 3 m. With practice, you will be
able to hold the ball properly so that the lighter ball rises vertically up and does
not fly sideways. This is head-on collision.
You can try to find the best combination of balls which gives you the best effect.
You can measure the masses on a standard balance. We leave it to you to think
how you can determine the initial and final velocities of the balls.
(6.25)
From Eqs. (6.24) and (6.25) it follows that,
or,
Hence, (6.26)
Substituting this in Eq. (6.24), we obtain
(6.27)
and (6.28)
Thus, the unknowns {v1f, v2f} are obtained in terms of the knowns
{m1, m2, v1i}. Special cases of our analysis are interesting.
Case I : If the two masses are equal
v1f = 0
v2f = v1i
The first mass comes to rest and pushes off the second mass with its
initial speed on collision.
Case II : If one mass dominates, e.g. m2 > > m1 v1f ~ v1i v2f ~ 0
The heavier mass is undisturbed while the lighter mass reverses its
velocity.
Example 6.12 Slowing down of neutrons: In a nuclear reactor a
neutron of high speed (typically 107 m s1) must be slowed to 103 m s
1 so that it can have a high probability of interacting with isotope
and causing it to fission. Show that a neutron can lose most of its
kinetic energy in an elastic collision with a light nuclei like deuterium or
carbon which has a mass of only a few times the neutron mass. The
material making up the light nuclei, usually heavy water (D2O) or
graphite, is called a moderator.
Answer The initial kinetic energy of the neutron is
while the fractional kinetic energy gained by the moderating nuclei K2f
/K1i is
f2 = 1 f1 (elastic collision)
One can also verify this result by substituting from Eq. (6.28).
For deuterium m2 = 2m1 and we obtain f1 = 1/9 while f2 = 8/9. Almost
90% of the neutrons energy is transferred to deuterium. For carbon f1
= 71.6% and f2 = 28.4%. In practice, however, this number is smaller
since head-on collisions are rare. t
If the initial velocities and final velocities of both the bodies are along
the same straight line, then it is called a one-dimensional collision, or
head-on collision. In the case of small spherical bodies, this is
possible if the direction of travel of body 1 passes through the centre
of body 2 which is at rest. In general, the collision is two-dimensional,
where the initial velocities and the final velocities lie in a plane.
6.12.3 Collisions in Two Dimensions
Fig. 6.10 also depicts the collision of a moving mass m1 with the
stationary mass m2. Linear momentum is conserved in such a
collision. Since momentum is a vector this implies three equations for
the three directions {x, y, z}. Consider the plane determined by the
final velocity directions of m1 and m2 and choose it to be the x-y plane.
The conservation of the z-component of the linear momentum implies
that the entire collision is in the x-y plane. The x- and y-component
equations are
m1v1i = m1v1f cos 1 + m2v2f cos 2 (6.29)
0 = m1v1f sin 1 m2v2f sin 2 (6.30)
One knows {m1, m2, v1i} in most situations. There are thus four
unknowns {v1f , v2f , 1 and 2}, and only two equations. If 1 = 2 = 0,
we regain Eq. (6.24) for one dimensional collision.
If, further the collision is elastic,
(6.31)
We obtain an additional equation. That still leaves us one equation
short. At least one of the four unknowns, say 1, must be made
known for the problem to be solvable. For example, 1 can be
determined by moving a detector in an angular fashion from the x to
the y axis. Given {m1, m2, v1i , 1} we can determine {v1f , v2f , 2} from
Eqs. (6.29)-(6.31).
Example 6.13 Consider the collision depicted in Fig. 6.10 to be
between two billiard balls with equal masses m1 = m2. The first ball is
called the cue while the second ball is called the target. The billiard
player wants to sink the target ball in a corner pocket, which is at an
angle 2 = 37. Assume that the collision is elastic and that friction and
rotational motion are not important. Obtain 1.
Answer From momentum conservation, since the masses are equal
or
(6.32)
Since the collision is elastic and m1 = m2 it follows from conservation
or
4. The principle of conservation of mechanical energy states that the total
mechanical energy of a body remains constant if the only forces that act on the
body are conservative.
5. The gravitational potential energy of a particle of mass m at a height x about
the earths surface is
V(x) = m g x
where the variation of g with height is ignored.
6. The elastic potential energy of a spring of force constant k and extension x is
7. The scalar or dot product of two vectors A and B is written as A.B and is a
scalar quantity given by : A.B = AB cos , where is the angle between A and
B. It can be positive, negative or zero depending upon the value of . The scalar
product of two vectors can be interpreted as the product of magnitude of one
vector and component of the other vector along the first vector. For unit vectors :
and
Scalar products obey the commutative and the distributive laws.
POINTS TO PONDER
1. The phrase calculate the work done is incomplete. We should refer (or
imply clearly by context) to the work done by a specific force or a group of
forces on a given body over a certain displacement.
2. Work done is a scalar quantity. It can be positive or negative unlike mass
and kinetic energy which are positive scalar quantities. The work done by the
friction or viscous force on a moving body is negative.
3. For two bodies, the sum of the mutual forces exerted between them is zero
from Newtons Third Law,
F12 +F21= 0
But the sum of the work done by the two forces need not always cancel, i.e.
W12+ W210
However, it may sometimes be true.
4. The work done by a force can be calculated sometimes even if the exact
nature of the force is not known. This is clear from Example 6.2 where the WE
theorem is used in such a situation.
5. The WE theorem is not independent of Newtons Second Law. The WE
theorem may be viewed as a scalar form of the Second Law. The principle of
conservation of mechanical energy may be viewed as a consequence of the
WE theorem for conservative forces.
6. The WE theorem holds in all inertial frames. It can also be extended to non-
inertial frames provided we include the pseudoforces in the calculation of the
net force acting on the body under consideration.
7. The potential energy of a body subjected to a conservative force is always
undetermined upto a constant. For example, the point where the potential
energy is zero is a matter of choice. For the gravitational potential energymgh,
the zero of the potential energy is chosen to be the ground. For the spring
potential energy kx2/2, the zero of the potential energy is the equilibrium
position of the oscillating mass.
8. Every force encountered in mechanics does not have an associated
potential energy. For example, work done by friction over a closed path is not
zero and no potential energy can be associated with friction.
9. During a collision : (a) the total linear momentum is conserved at each
instant of the collision ; (b) the kinetic energy conservation (even if the
collision is elastic) applies after the collision is over and does not hold at every
instant of the collision. In fact the two colliding objects are deformed and may
be momentarily at rest with respect to each other.
Exercises
]Fig. 6.13
where are unit vectors along the x-, y- and z-axis of the
system respectively. What is the work done by this force in moving
the body a distance of 4 m along the z-axis ?
6.12 An electron and a proton are detected in a cosmic ray
experiment, the first with kinetic energy 10 keV, and the second
with 100 keV. Which is faster, the electron or the proton ? Obtain
the ratio of their speeds. (electron mass = 9.1110-31 kg, proton
mass = 1.671027 kg, 1 eV = 1.60 1019 J).
6.13 A rain drop of radius 2 mm falls from a height of 500 m above
the ground. It falls with decreasing acceleration (due to viscous
resistance of the air) until at half its original height, it attains its
maximum (terminal) speed, and moves with uniform speed
thereafter. What is the work done by the gravitational force on the
drop in the first and second half of its journey ? What is the work
done by the resistive force in the entire journey if its speed on
reaching the ground is 10 m s1 ?
6.14 A molecule in a gas container hits a horizontal wall with
speed 200 m s1 and angle 30 with the normal, and rebounds
with the same speed. Is momentum conserved in the collision ? Is
the collision elastic or inelastic ?
6.15 A pump on the ground floor of a building can pump up water
to fill a tank of volume 30 m3 in 15 min. If the tank is 40 m above
the ground, and the efficiency of the pump is 30%, how much
electric power is consumed by the pump ?
6.16 Two identical ball bearings in contact with each other and
resting on a frictionless table are hit head-on by another ball
bearing of the same mass moving initially with a speed V. If the
collision is elastic, which of the following (Fig. 6.14) is a possible
result after collision ?
Fig. 6.14
Additional Exercises
6.24 A bullet of mass 0.012 kg and horizontal speed 70 m s1
strikes a block of wood of mass 0.4 kg and instantly comes to rest
with respect to the block. The block is suspended from the ceiling
by means of thin wires. Calculate the height to which the block
rises. Also, estimate the amount of heat produced in the block.
6.25 Two inclined frictionless tracks, one gradual and the other
steep meet at A from where two stones are allowed to slide down
from rest, one on each track (Fig. 6.16). Will the stones reach the
bottom at the same time? Will they reach there with the same
speed? Explain. Given 1 = 300, 2 = 600, and h = 10 m, what are
the speeds and times taken by the two stones ?
Fig. 6.16
6.26 A 1 kg block situated on a rough incline is connected to a
spring of spring constant 100 N m1 as shown in Fig. 6.17. The
block is released from rest with the spring in the unstretched
position. The block moves 10 cm down the incline before coming
to rest. Find the coefficient of friction between the block and the
incline. Assume that the spring has a negligible mass and the
pulley is frictionless.
Fig. 6.17
6.27 A bolt of mass 0.3 kg falls from the ceiling of an elevator
moving down with an uniform speed of 7 m s1. It hits the floor of
the elevator (length of the elevator = 3 m) and does not rebound.
What is the heat produced by the impact ? Would your answer be
different if the elevator were stationary ?
6.28 A trolley of mass 200 kg moves with a uniform speed of 36
km/h on a frictionless track. A child of mass 20 kg runs on the
trolley from one end to the other (10 m away) with a speed of 4 m
s1 relative to the trolley in a direction opposite to the its motion,
and jumps out of the trolley. What is the final speed of the trolley ?
How much has the trolley moved from the time the child begins to
run ?
6.29 Which of the following potential energy curves in Fig. 6.18
cannot possibly describe the elastic collision of two billiard balls ?
Here r is the distance between centres of the balls.
Fig. 6.18
Fig. 6.20 An illustration of a single stride in walking. While the first leg is maximally
off the round, the second leg is on the ground and vice-versa
(6.34)
Assuming ml = 10 kg and slow running of a nine-minute mile which
translates to 3 m s-1 in SI units, we obtain
= 270 W
We must bear in mind that this is a lower estimate since several
avenues of power loss (e.g. swinging of hands, air resistance etc.)
have been ignored. The interesting point is that we did not worry about
the forces involved. The forces, mainly friction and those exerted on
the leg by the muscles of the rest of the body, are hard to estimate.
Static friction does no work and we bypassed the impossible task of
estimating the work done by the muscles by taking recourse to the
work-energy theorem. We can also see the advantage of a wheel. The
wheel permits smooth locomotion without the continual starting and
stopping in mammalian locomotion.
Chapter Nine
9.1 Introduction
Nature has endowed the human eye (retina) with the sensitivity to
detect electromagnetic waves within a small range of the
electromagnetic spectrum. Electromagnetic radiation belonging to this
region of the spectrum (wavelength of about 400 nm to 750 nm) is
called light. It is mainly through light and the sense of vision that we
know and interpret the world around us.
There are two things that we can intuitively mention about light from
common experience. First, that it travels with enormous speed and
second, that it travels in a straight line. It took some time for people to
realise that the speed of light is finite and measurable. Its presently
accepted value in vacuum is c = 2.99792458 108 m s1. For many
purposes, it suffices to take c = 3 108 m s1. The speed of light in
vacuum is the highest speed attainable in nature.
The intuitive notion that light travels in a straight line seems to
contradict what we have learnt in Chapter 8, that light is an
electromagnetic wave of wavelength belonging to the visible part of
the spectrum. How to reconcile the two facts? The answer is that the
wavelength of light is very small compared to the size of ordinary
objects that we encounter commonly (generally of the order of a few
cm or larger). In this situation, as you will learn in Chapter 10, a light
wave can be considered to travel from one point to another, along a
straight line joining them. The path is called a ray of light, and a
bundle of such rays constitutes a beam of light.
In this chapter, we consider the phenomena of reflection, refraction
and dispersion of light, using the ray picture of light. Using the basic
laws of reflection and refraction, we shall study the image formation by
plane and spherical reflecting and refracting surfaces. We then go on
to describe the construction and working of some important optical
instruments, including the human eye.
We are familiar with the laws of reflection. The angle of reflection (i.e.,
the angle between reflected ray and the normal to the reflecting
surface or the mirror) equals the angle of incidence (angle between
incident ray and the normal). Also that the incident ray, reflected ray
and the normal to the reflecting surface at the point of incidence lie in
the same plane
(Fig. 9.1). These laws are valid at each point on any reflecting surface
whether plane or curved. However, we shall restrict our discussion to
the special case of curved surfaces, that is, spherical surfaces. The
normal in this case is to be taken as normal to the tangent to surface
at the point of incidence. That is, the normal is along the radius, the
line joining the centre of curvature of the mirror to the point of
incidence.
Figure 9.1 The incident ray, reflected ray and the normal to the reflecting surface
lie in the same plane.
=2
or, FD = (9.2)
Now, for small , the point D is very close to the point P. Therefore,
FD = f and CD = R. Equation (9.2) then gives
f = R/2 (9.3)
(iv) The ray incident at any angle at the pole. The reflected ray follows
laws of reflection.
Figure 9.5 shows the ray diagram considering three rays. It shows the
image AB (in this case, real) of an object AB formed by a concave
mirror. It does not mean that only three rays emanate from the point A.
An infinite number of rays emanate from any source, in all directions.
Thus, point A is image point of A if every ray originating at point A and
falling on the concave mirror after reflection passes through the point
A.
We now derive the mirror equation or the relation between the object
distance (u), image distance (v) and the focal length (f ).
From Fig. 9.5, the two right-angled triangles ABF and MPF are
similar. (For paraxial rays, MP can be considered to be a straight line
perpendicular to CP.) Therefore,
or ( PM = AB) (9.4)
Since APB = APB, the right angled triangles ABP and ABP are
also similar. Therefore,
(9.5)
Comparing Eqs. (9.4) and (9.5), we get
(9.6)
or
(9.7)
m= (9.8)
h and h will be taken positive or negative in accordance with the
accepted sign convention. In triangles ABP and ABP, we have,
With the sign convention, this becomes
so that
m= (9.9)
We have derived here the mirror equation, Eq. (9.7), and the
magnification formula, Eq. (9.9), for the case of real, inverted image
formed by a concave mirror. With the proper use of sign convention,
these are,in fact, valid for all the cases of reflection by a spherical
mirror (concave or convex) whether the image formed is real or virtual.
Figure 9.6 shows the ray diagrams for virtual image formed by a
concave and convex mirror. You should verify that Eqs. (9.7) and (9.9)
are valid for these cases as well.
Figure 9.6 Image formation by (a) a concave mirror with object between
P and F, and (b) a convex mirror.
Example 9.1 Suppose that the lower half of the concave mirrors reflecting
surface in Fig. 9.5 is covered with an opaque (non-reflective) material. What
effect will this have on the image of an object placed in front of the mirror?
Solution You may think that the image will now show only half of the object,
but taking the laws of reflection to be true for all points of the remaining part of
the mirror, the image will be that of the whole object. However, as the area of
the reflecting surface has been reduced, the intensity of the image will be low
(in this case, half).
Example 9.2 A mobile phone lies along the principal axis of a concave mirror,
as shown in Fig. 9.7. Show by suitable diagram, the formation of its image.
Explain why the magnification is not uniform. Will the distortion of image
depend on the location of the phone with respect to the mirror?
Figure 9.7
Solution The ray diagram for the formation of the image of the phone is shown
in Fig. 9.7. The image of the part which is on the plane perpendicular to
principal axis will be on the same plane. It will be of the same size, i.e., BC =
BC. You can yourself realise why the image is distorted.
Example 9.3 An object is placed at (i) 10 cm, (ii) 5 cm in front of a concave
mirror of radius of curvature 15 cm. Find the position, nature, and
magnification of the image in each case.
Solution The focal length f = 15/2 cm = 7.5 cm
(i) The object distance u = 10 cm. Then Eq. (9.7) gives
or = 30 cm
The image is 30 cm from the mirror on the same side as the object.
Also, magnification m =
or
Magnification m =
Since the jogger moves at a constant speed of 5 m s1, after 1 s the position of
the image v (for u = 39 + 5 = 34) is (34/35 )m.
The shift in the position of image in 1 s is
Therefore, the average speed of the image when the jogger is between
39 m and 34 m from the mirror, is (1/280) m s1
Similarly, it can be seen that for u = 29 m, 19 m and 9 m, the speed with
which the image appears to move is
respectively.
Although the jogger has been moving with a constant speed, the speed of
his/her image appears to increase substantially as he/she moves closer to the
mirror. This phenomenon can be noticed by any person sitting in a stationary
car or a bus. In case of moving vehicles, a similar phenomenon could be
observed if the vehicle in the rear is moving closer with a constant speed.
9.3 Refraction
(9.10)
where n21 is a constant, called the refractive index of the second
medium with respect to the first medium. Equation (9.10) is the well-
known Snells law of refraction. We note that n21 is a characteristic of
the pair of media (and also depends on the wavelength of light), but is
independent of the angle of incidence.
Figure 9.8 Refraction and reflection of light.
From Eq. (9.10), if n21 > 1, r < i, i.e., the refracted ray bends towards
the normal. In such a case medium 2 is said to be optically denser (or
denser, in short) than medium 1. On the other hand, if n21 <1, r > i, the
refracted ray bends away from the normal. This is the case when
incident ray in a denser medium refracts into a rarer medium.
Note: Optical density should not be confused with mass density, which
is mass per unit volume. It is possible that mass density of an optically
denser medium may be less than that of an optically rarer medium
(optical density is the ratio of the speed of light in two media). For
example, turpentine and water. Mass density of turpentine is less than
that of water but its optical density is higher.
If n21 is the refractive index of medium 2 with respect to medium 1
and n12 the refractive index of medium 1 with respect to medium 2,
then it should be clear that
(9.11)
It also follows that if n32 is the refractive index of medium 3 with
respect to medium 2 then n32 = n31 n12, where n31 is the refractive
index of medium 3 with respect to medium 1.
Example 9.5 The earth takes 24 h to rotate once about its axis. How much
time does the sun take to shift by 1 when viewed from
the earth?
Solution
Time taken for 360 shift = 24 h
Time taken for 1 shift = 24/360 h = 4 min.
9.4 Total Internal Reflection
Figure 9.11 Advance sunrise and delayed sunset due to atmospheric refraction.
To make this time minimum, one has to differentiate it (with respect to the
coordinate of X) and find the point X when t is a minimum. On doing all this
algebra (which we skip here), we find that the guard should enter water at a
point where Snells law is satisfied. To understand this, draw a perpendicular
LM to side SR at X. Let GXM = i and CXL = r. Then it can be seen that t is
minimum when
In the case of light v1/v2, the ratio of the velocity of light in vacuum to that in
the medium, is the refractive index n of the medium.
In short, whether it is a wave or a particle or a human being, whenever two
mediums and two velocities are involved, one must follow Snells law if one
wants to take the shortest time.
All optical phenomena can be demonstrated very easily with the use
of a laser torch or pointer, which is easily available nowadays. Take a
glass beaker with clear water in it. Stir the water a few times with a
piece of soap, so that it becomes a little turbid. Take a laser pointer
and shine its beam through the turbid water. You will find that the path
of the beam inside the water shines brightly.
Shine the beam from below the beaker such that it strikes at the upper
water surface at the other end. Do you find that it undergoes partial
reflection (which is seen as a spot on the table below) and partial
refraction [which comes out in the air and is seen as a spot on the
roof; Fig. 9.13(a)]? Now direct the laser beam from one side of the
beaker such that it strikes the upper surface of water more obliquely
[Fig. 9.13(b)]. Adjust the direction of laser beam until you find the
angle for which the refraction above the water surface is totally absent
and the beam is totally reflected back to water. This is total internal
reflection at its simplest.
Pour this water in a long test tube and shine the laser light from top,
as shown in Fig. 9.13(c). Adjust the direction of the laser beam such
that it is totally internally reflected every time it strikes the walls of the
tube. This is similar to what happens in optical fibres.
Take care not to look into the laser beam directly and not to point it at
anybodys face.
(i) Mirage: On hot summer days, the air near the ground becomes
hotter than the air at higher levels. The refractive index of air
increases with its density. Hotter air is less dense, and has smaller
refractive index than the cooler air. If the air currents are small, that is,
the air is still, the optical density at different layers of air increases with
height. As a result, light from a tall object such as a tree, passes
through a medium whose refractive index decreases towards the
ground. Thus, a ray of light from such an object successively bends
away from the normal and undergoes total internal reflection, if the
angle of incidence for the air near the ground exceeds the critical
angle. This is shown in Fig. 9.14(b). To a distant observer, the light
appears to be coming from somewhere below the ground. The
observer naturally assumes that light is being reflected from the
ground, say, by a pool of water near the tall object. Such inverted
images of distant tall objects cause an optical illusion to the observer.
This phenomenon is called mirage. This type of mirage is especially
common in hot deserts. Some of you might have noticed that while
moving in a bus or a car during a hot summer day, a distant patch of
road, especially on a highway, appears to be wet. But, you do not find
any evidence of wetness when you reach that spot. This is also due to
mirage.
Figure 9.13 Observing total internal reflection in water with a laser beam (refraction
due to glass of beaker neglected being very thin).
Figure 9.15 Prisms designed to bend rays by 90 and 180 or to invert image
without changing its size make use of total internal reflection.
A bundle of optical fibres can be put to several uses. Optical fibres are
extensively used for transmitting and receiving electrical signals which
are converted to light by suitable transducers. Obviously, optical fibres
can also be used for transmission of optical signals. For example,
these are used as a light pipe to facilitate visual examination of
internal organs like esophagus, stomach and intestines. You might
have seen a commonly available decorative lamp with fine plastic
fibres with their free ends forming a fountain like structure. The other
end of the fibres is fixed over an electric lamp. When the lamp is
switched on, the light travels from the bottom of each fibre and
appears at the tip of its free end as a dot of light. The fibres in such
decorative lamps are optical fibres.
The main requirement in fabricating optical fibres is that there should
be very little absorption of light as it travels for long distances inside
them. This has been achieved by purification and special preparation
of materials such as quartz. In silica glass fibres, it is possible to
transmit more than 95% of the light over a fibre length of 1 km.
(Compare with what you expect for a block of ordinary window glass 1
km thick.)
tan NOM =
tan NCM =
tan NIM =
i= (9.13)
Similarly,
r = NCM NIM
i.e., r = (9.14)
Now, by Snells law
n1 sin i = n2 sin r
(9.15)
Here, OM, MI and MC represent magnitudes of distances. Applying
the Cartesian sign convention,
OM = u, MI = +v, MC = +R
Substituting these in Eq. (9.15), we get
(9.16)
Equation (9.16) gives us a relation between object and image distance
in terms of refractive index of the medium and the radius of curvature
of the curved spherical surface. It holds for any curved spherical
surface.
Example 9.6 Light from a point source in air falls on a spherical glass surface
(n = 1.5 and radius of curvature = 20 cm). The distance of the light source from
the glass surface is 100 cm. At what position the image is formed?
Solution
We use the relation given by Eq. (9.16). Here
u = 100 cm, v = ?, R = + 20 cm, n1 = 1, and n2 = 1.5.
We then have
or v = +100 cm
The image is formed at a distance of 100 cm from the glass surface, in the
direction of incident light.
(9.17)
A similar procedure applied to the second interface* ADC gives,
(9.18)
* Note that now the refractive index of the medium on the right side of
ADC is n1 while on its left it is n2. Further DI1 is negative as the
distance is measured against the direction of incident light.
For a thin lens, BI1 = DI1. Adding
Eqs. (9.17) and (9.18), we get
(9.19)
Suppose the object is at infinity, i.e.,
OB and DI = f, Eq. (9.19) gives
(9.20)
The point where image of an object placed at infinity is formed is
called the focus F, of the lens and the distance f gives its focal length.
A lens has two foci, F and F, on either side of it (Fig. 9.19). By the
sign convention,
BC1 = + R1,
DC2 = R2
(9.21)
Equation (9.21) is known as the lens makers formula. It is useful to
design lenses of desired focal length using surfaces of suitable radii of
curvature. Note that the formula is true for a concave lens also. In that
case R1is negative, R2 positive and therefore, f is negative.
Figure 9.18 (a) The position of object, and the image formed by a double convex
lens,
(b) Refraction at the first spherical surface and
(c) Refraction at the second spherical surface.
(9.22)
Again, in the thin lens approximation, B and D are both close to the
optical centre of the lens. Applying the sign convention,
BO = u, DI = +v, we get
(9.23)
Equation (9.23) is the familiar thin lens formula. Though we derived it
for a real image formed by a convex lens, the formula is valid for both
convex as well as concave lenses and for both real and virtual
images.
It is worth mentioning that the two foci, F and F, of a double convex or
concave lens are equidistant from the optical centre. The focus on the
side of the (original) source of light is called the first focal point,
whereas the other is called the second focal point.
To find the image of an object by a lens, we can, in principle, take any
two rays emanating from a point on an object; trace their paths using
the laws of refraction and find the point where the refracted rays meet
(or appear to meet). In practice, however, it is convenient to choose
any two of the following rays:
(i) A ray emanating from the object parallel to the principal axis of the
lens after refraction passes through the second principal focus F (in a
convex lens) or appears to diverge (in a concave lens) from the first
principal focus F.
(ii) A ray of light, passing through the optical centre of the lens,
emerges without any deviation after refraction.
(iii) A ray of light passing through the first principal focus (for a convex
lens) or appearing to meet at it (for a concave lens) emerges parallel
to the principal axis after refraction.
Figure 9.19 Tracing rays through (a) convex lens (b) concave lens.
Figures 9.19(a) and (b) illustrate these rules for a convex and a
concave lens, respectively. You should practice drawing similar ray
diagrams for different positions of the object with respect to the lens
and also verify that the lens formula, Eq. (9.23), holds good for all
cases.
m= = (9.24)
When we apply the sign convention, we see that, for erect (and virtual)
image formed by a convex or concave lens, m is positive, while for an
inverted (and real) image, m is negative.
P= (9.25)
The SI unit for power of a lens is dioptre (D): 1D = 1m1. The power of
a lens of focal length of 1 metre is one dioptre. Power of a lens is
positive for a converging lens and negative for a diverging lens. Thus,
when an optician prescribes a corrective lens of power + 2.5 D, the
required lens is a convex lens of focal length + 40 cm. A lens of power
of 4.0 D means a concave lens of focal length 25 cm.
Figure 9.20 Power of a lens.
Example 9.8 (i) If f = 0.5 m for a glass lens, what is the power of the lens? (ii)
The radii of curvature of the faces of a double convex lens are 10 cm and 15
cm. Its focal length is 12 cm. What is the refractive index of glass? (iii) A
convex lens has 20 cm focal length in air. What is focal length in water?
(Refractive index of air-water = 1.33, refractive index for air-glass = 1.5.)
Solution
(i) Power = +2 dioptre.
(ii) Here, we have f = +12 cm, R1 = +10 cm, R2 = 15 cm.
(9.26)
Combining these two equations, we find f = + 78.2 cm.
(9.28)
Adding Eqs. (9.27) and (9.28), we get
(9.29)
If the two lens-system is regarded as equivalent to a single lens of
focal length f, we have
so that we get
(9.30)
(9.31)
In terms of power, Eq. (9.31) can be written as
P = P1 + P2 + P3 + (9.32)
where P is the net power of the lens combination. Note that the sum in
Eq. (9.32) is an algebraic sum of individual powers, so some of the
terms on the right side may be positive (for convex lenses) and some
negative (for concave lenses). Combination of lenses helps to obtain
diverging or converging lenses of desired magnification. It also
enhances sharpness of the image. Since the image formed by the first
lens becomes the object for the second, Eq. (9.25) implies that the
total magnification m of the combination is a product of magnification
(m1, m2, m3,...) of individual lenses
m = m1 m2 m3 ... (9.33)
Example 9.9 Find the position of the image formed by the lens combination
given in the Fig. 9.22.
Figure 9.22
or v1 = 15 cm
The image formed by the first lens serves as the object for the second. This is
at a distance of (15 5) cm = 10 cm to the right of the second lens. Though
the image is real, it serves as a virtual object for the second lens, which means
that the rays appear to come from it for the second lens.
or v2 =
The virtual image is formed at an infinite distance to the left of the second lens.
This acts as an object for the third lens.
or
or v3 = 30 cm
The final image is formed 30 cm to the right of the third lens.
that is,
= i + e A (9.35)
Thus, the angle of deviation depends on the angle of incidence. A plot
between the angle of deviation and angle of incidence is shown in Fig.
9.24.
You can see that, in general, any given value of , except for i = e,
corresponds to two values i and hence of e. This, in fact, is expected
from the symmetry of i and e in Eq. (9.35), i.e., remains the same if i
and e are interchanged. Physically, this is related to the fact that the
path of ray in Fig. 9.23 can be traced back, resulting in the same angle
of deviation. At the minimum deviation Dm, the refracted ray inside the
prism becomes parallel to its base. We have
= Dm, i = e which implies r1 = r2.
(9.38)
The angles A and Dm can be measured experimentally. Equation
(9.38) thus provides a method of determining refractive index of the
material of the prism.
For a small angle prism, i.e., a thin prism, Dm is also very small, and
we get
Dm = (n211)A
It has been known for a long time that when a narrow beam of
sunlight, usually called white light, is incident on a glass prism, the
emergent light is seen to be consisting of several colours. There is
actually a continuous variation of colour, but broadly, the different
component colours that appear in sequence are: violet, indigo, blue,
green, yellow, orange and red (given by the acronym VIBGYOR). The
red light bends the least, while the violet light bends the most (Fig.
9.25).
Figure 9.25 Dispersion of sunlight or white light on passing through a glass prism.
The relative deviation of different colours shown is highly exaggerated.
In a classic experiment known for its simplicity but great significance,
Isaac Newton settled the issue once for all. He put another similar
prism, but in an inverted position, and let the emergent beam from the
first prism fall on the second prism (Fig. 9.26). The resulting emergent
beam was found to be white light. The explanation was clear the
first prism splits the white light into its component colours, while the
inverted prism recombines them to give white light. Thus, white light
itself consists of light of different colours, which are separated by the
prism.
It must be understood here that a ray of light, as defined
mathematically, does not exist. An actual ray is really a beam of many
rays of light. Each ray splits into component colours when it enters the
glass prism. When those coloured rays come out on the other side,
they again produce a white beam.
Figure 9.29 (a) shows the eye. Light enters the eye through a curved
front surface, the cornea. It passes through the pupil which is the
central hole in the iris. The size of the pupil can change under control
of muscles. The light is further focussed by the eye lens on the retina.
The retina is a film of nerve fibres covering the curved back surface of
the eye. The retina contains rods and cones which sense light
intensity and colour, respectively, and transmit electrical signals via
the optic nerve to the brain which finally processes this information.
The shape (curvature) and therefore the focal length of the lens can
be modified somewhat by the ciliary muscles. For example, when the
muscle is relaxed, the focal length is about 2.5 cm and objects at
infinity are in sharp focus on the retina. When the object is brought
closer to the eye, in order to maintain the same image-lens distance (
2.5 cm), the focal length of the eye lens becomes shorter by the action
of the ciliary muscles. This property of the eye is called
accommodation. If the object is too close to the eye, the lens cannot
curve enough to focus the image on to the retina, and the image is
blurred. The closest distance for which the lens can focus light on the
retina is called the least distance of distinct vision, or the near point.
The standard value for normal vision is taken as 25 cm. (Often the
near point is given the symbol D.) This distance increases with age,
because of the decreasing effectiveness of the ciliary muscle and the
loss of flexibility of the lens. The near point may be as close as about
7 to 8 cm in a child ten years of age, and may increase to as much as
200 cm at 60 years of age. Thus, if an elderly person tries to read a
book at about 25 cm from the eye, the image appears blurred. This
condition (defect of the eye) is called presbyopia. It is corrected by
using a converging lens for reading.
Thus, our eyes are marvellous organs that have the capability to
interpret incoming electromagnetic waves as images through a
complex process. These are our greatest assets and we must take
proper care to protect them. Imagine the world without a pair of
functional eyes. Yet many amongst us bravely face this challenge by
effectively overcoming their limitations to lead a normal life. They
deserve our appreciation for their courage and conviction.
In spite of all precautions and proactive action, our eyes may develop
some defects due to various reasons. We shall restrict our discussion
to some common optical defects of the eye. For example, the light
from a distant object arriving at the eye-lens may get converged at a
point in front of the retina. This type of defect is called
nearsightedness or myopia. This means that the eye is producing too
much convergence in the incident beam. To compensate this, we
interpose a concave lens between the eye and the object, with the
diverging effect desired to get the image focussed on the retina [Fig.
9.29(b)].
Figure 9.29 (a) The structure of the eye; (b) shortsighted or myopic eye and its
correction;
(c) farsighted or hypermetropic eye and its correction; and (d) astigmatic eye and
its correction.
Example 9.10 What focal length should the reading spectacles have for a
person for whom the least distance of distinct vision is 50 cm?
Solution The distance of normal vision is 25 cm. So if a book is at
u = 25 cm, its image should be formed at v = 50 cm. Therefore, the desired
focal length is given by
or
or f = + 50 cm (convex lens).
Example 9.11
(a) The far point of a myopic person is 80 cm in front of the eye. What is the
power of the lens required to enable him to see very distant objects clearly?
(b) In what way does the corrective lens help the above person? Does the lens
magnify very distant objects? Explain carefully.
(c) The above person prefers to remove his spectacles while reading a book.
Explain why?
Solution
(a) Solving as in the previous example, we find that the person should use a
concave lens of focal length = 80 cm, i.e., of power = 1.25 dioptres.
(b) No. The concave lens, in fact, reduces the size of the object, but the angle
subtended by the distant object at the eye is the same as the angle subtended
by the image (at the far point) at the eye. The eye is able to see distant objects
not because the corrective lens magnifies the object, but because it brings the
object (i.e., it produces virtual image of the object) at the far point of the eye
which then can be focussed by the eye-lens on the retina.
(c) The myopic person may have a normal near point, i.e., about
25 cm (or even less). In order to read a book with the spectacles, such a
person must keep the book at a distance greater than
25 cm so that the image of the book by the concave lens is produced not
closer than 25 cm. The angular size of the book (or its image) at the greater
distance is evidently less than the angular size when the book is placed at 25
cm and no spectacles are needed. Hence, the person prefers to remove the
spectacles while reading.
Example 9.12 (a) The near point of a hypermetropic person is 75 cm from the
eye. What is the power of the lens required to enable the person to read
clearly a book held at 25 cm from the eye? (b) In what way does the corrective
lens help the above person? Does the lens magnify objects held near the eye?
(c) The above person prefers to remove the spectacles while looking at the
sky. Explain why?
Solution
(a) u = 25 cm, v = 75 cm
1/f = 1/25 1/75, i.e., f = 37.5 cm.
The corrective lens needs to have a converging power of +2.67 dioptres.
(b) The corrective lens produces a virtual image (at 75 cm) of an object at 25
cm. The angular size of this image is the same as that of the object. In this
sense the lens does not magnify the object but merely brings the object to the
near point of the hypermetric eye, which then gets focussed on the retina.
However, the angular size is greater than that of the same object at the near
point (75 cm) viewed without the spectacles.
(c) A hypermetropic eye may have normal far point i.e., it may have enough
converging power to focus parallel rays from infinity on the retina of the
shortened eyeball. Wearing spectacles of converging lenses (used for near
vision) will amount to more converging power than needed for parallel rays.
Hence the person prefers not to use the spectacles for far objects.
(9.39)
Since D is about 25 cm, to have a magnification of six, one needs a
convex lens of focal length, f = 5 cm.
Note that m = h/h where h is the size of the object and h the size of
the image. This is also the ratio of the angle subtended by the image
to that subtended by the object, if placed at D for comfortable viewing.
(Note that this is not the angle actually subtended by the object at the
eye, which is h/u.) What a single-lens simple magnifier achieves is
that it allows the object to be brought closer to the eye than D.
Figure 9.30 A simple microscope; (a) the magnifying lens is located such that the
image is at the near point, (b) the angle subtanded by the object, is the same as
that at the near point, and (c) the object near the focal point of the lens; the image
is far off but closer than infinity.
We will now find the magnification when the image is at infinity. In this
case we will have to obtained the angular magnification. Suppose the
object has a height h. The maximum angle it can subtend, and be
clearly visible (without a lens), is when it is at the near point, i.e., a
distance D. The angle subtended is then given by
tan o (9.40)
We now find the angle subtended at the eye by the image when the
object is at u. From the relations
(9.41)
as is clear from Fig. 9.29(c). The angular magnification is, therefore
(9.42)
This is one less than the magnification when the image is at the near
point, Eq. (9.39), but the viewing is more comfortable and the
difference in magnification is usually small. In subsequent discussions
of optical instruments (microscope and telescope) we shall assume
the image to be at infinity.
A simple microscope has a limited maximum magnification ( 9) for
realistic focal lengths. For much larger magnifications, one uses two
lenses, one compounding the effect of the other. This is known as a
compound microscope. A schematic diagram of a compound
microscope is shown in Fig. 9.31. The lens nearest the object, called
the objective, forms a real, inverted, magnified image of the object.
This serves as the object for the second lens, the eyepiece, which
functions essentially like a simple microscope or magnifier, produces
the final image, which is enlarged and virtual. The first inverted image
is thus near (at or within) the focal plane of the eyepiece, at a distance
appropriate for final image formation at infinity, or a little closer for
image formation at the near point. Clearly, the final image is inverted
with respect to the original object.
We now obtain the magnification due to a compound microscope. The
ray diagram of
Fig. 9.31 shows that the (linear) magnification due to the objective,
namely h/h, equals
(9.43)
where we have used the result
Here h is the size of the first image, the object size being h and fo
being the focal length of the objective. The first image is formed near
the focal point of the eyepiece. The distance L, i.e., the distance
between the second focal point of the objective and the first focal point
of the eyepiece (focal length fe) is called the tube length of the
compound microscope.
Figure 9.31 Ray diagram for the formation of image by a compound microscope.
As the first inverted image is near the focal point of the eyepiece, we
use the result from the discussion above for the simple microscope to
obtain the (angular) magnification me due to it [Eq. (9.39)], when the
final image is formed at the near point, is
[9.44(a)]
(9.45)
Clearly, to achieve a large magnification of a small object (hence the
name microscope), the objective and eyepiece should have small
focal lengths. In practice, it is difficult to make the focal length much
smaller than 1 cm. Also large lenses are required to make L large.
For example, with an objective with fo = 1.0 cm, and an eyepiece with
focal length fe = 2.0 cm, and a tube length of 20 cm, the magnification
is
9.9.3 Telescope
(9.46)
Summary
1. Reflection is governed by the equation i = r and refraction by the Snells
law, sini/sinr = n, where the incident ray, reflected ray, refracted ray and
normal lie in the same plane. Angles of incidence, reflection and refraction are
i, r and r, respectively.
2. The critical angle of incidence ic for a ray incident from a denser to rarer
medium, is that angle for which the angle of refraction is 90. For
i > ic, total internal reflection occurs. Multiple internal reflections in diamond (ic
24.4), totally reflecting prisms and mirage, are some examples of total
internal reflection. Optical fibres consist of glass fibres coated with a thin layer
of material of lower refractive index. Light incident at an angle at one end
comes out at the other, after multiple internal reflections, even if the fibre is
bent.
3. Cartesian sign convention: Distances measured in the same direction as the
incident light are positive; those measured in the opposite direction are
negative. All distances are measured from the pole/optic centre of the
mirror/lens on the principal axis. The heights measured upwards above x-axis
and normal to the principal axis of the mirror/lens are taken as positive. The
heights measured downwards are taken as negative.
4. Mirror equation:
where u and v are object and image distances, respectively and f is the focal
length of the mirror. f is (approximately) half the radius of curvature R. f is
negative for concave mirror; f is positive for a convex mirror.
5. For a prism of the angle A, of refractive index n2 placed in a medium of
refractive index n1,
R1 and R2 are the radii of curvature of the lens surfaces. f is positive for a
converging lens; f is negative for a diverging lens. The power of a lens P = 1/f.
The total power of a combination of several lenses is
P = P1 + P2 + P3 +
7. Dispersion is the splitting of light into its constituent colours.
8. The Eye: The eye has a convex lens of focal length about 2.5 cm. This focal
length can be varied somewhat so that the image is always formed on the
retina. This ability of the eye is called accommodation. In a defective eye, if the
image is focussed before the retina (myopia), a diverging corrective lens is
needed; if the image is focussed beyond the retina (hypermetropia), a
converging corrective lens is needed. Astigmatism is corrected by using
cylindrical lenses.
9. Magnifying power m of a simple microscope is given by m = 1 + (D/f), where
D = 25 cm is the least distance of distinct vision and f is the focal length of the
convex lens. If the image is at infinity, m = D/f. For a compound microscope,
the magnifying power is given by
m = me m0 where me = 1 + (D/fe), is the magnification due to the eyepiece
and mo is the magnification produced by the objective. Approximately,
where fo and fe are the focal lengths of the objective and eyepiece,
respectively, and L is the distance between their focal points.
10. Magnifying power m of a telescope is the ratio of the angle subtended at
the eye by the image to the angle subtended at the eye by the object.
where f0 and fe are the focal lengths of the objective and eyepiece,
respectively.
Points to Ponder
1. The laws of reflection and refraction are true for all surfaces and pairs of
media at the point of the incidence.
2. The real image of an object placed between f and 2f from a convex lens can
be seen on a screen placed at the image location. If the screen is removed, is
the image still there? This question puzzles many, because it is difficult to
reconcile ourselves with an image suspended in air without a screen. But the
image does exist. Rays from a given point on the object are converging to an
image point in space and diverging away. The screen simply diffuses these
rays, some of which reach our eye and we see the image. This can be seen by
the images formed in air during a laser show.
3. Image formation needs regular reflection/refraction. In principle, all rays from
a given point should reach the same image point. This is why you do not see
your image by an irregular reflecting object, say the page of a book.
4. Thick lenses give coloured images due to dispersion. The variety in colour
of objects we see around us is due to the constituent colours of the light
incident on them. A monochromatic light may produce an entirely different
perception about the colours on an object as seen in white light.
5. For a simple microscope, the angular size of the object equals the angular
size of the image. Yet it offers magnification because we can keep the small
object much closer to the eye than 25 cm and hence have it subtend a large
angle. The image is at 25 cm which we can see. Without the microscope, you
would need to keep the small object at 25 cm which would subtend a very
small angle.
Exercises
9.3 A tank is filled with water to a height of 12.5 cm. The apparent
depth of a needle lying at the bottom of the tank is measured by a
microscope to be 9.4 cm. What is the refractive index of water? If
water is replaced by a liquid of refractive index 1.63 up to the
same height, by what distance would the microscope have to be
moved to focus on the needle again?
9.4 Figures 9.34(a) and (b) show refraction of a ray in air incident
at 60 with the normal to a glass-air and water-air interface,
respectively. Predict the angle of refraction in glass when the
angle of incidence in water is 45 with the normal to a water-glass
interface [Fig. 9.34(c)].
Figure 9.34
9.16 A small pin fixed on a table top is viewed from above from a
distance of 50cm. By what distance would the pin appear to be
raised if it is viewed from the same point through a 15cm thick
glass slab held parallel to the table? Refractive index of glass =
1.5. Does the answer depend on the location of the slab?
9.17 (a) Figure 9.35 shows a cross-section of a light pipe made of
a glass fibre of refractive index 1.68. The outer covering of the
pipe is made of a material of refractive index 1.44. What is the
range of the angles of the incident rays with the axis of the pipe for
which total reflections inside the pipe take place, as shown in the
figure.
Figure 9.35
9.18 Answer the following questions:
(a) You have learnt that plane and convex mirrors produce virtual
images of objects. Can they produce real images under some
circumstances? Explain.
(b) A virtual image, we always say, cannot be caught on a screen.
Yet when we see a virtual image, we are obviously bringing it on
to the screen (i.e., the retina) of our eye. Is there a contradiction?
(c) A diver under water, looks obliquely at a fisherman standing on
the bank of a lake. Would the fisherman look taller or shorter to the
diver than what he actually is?
(d) Does the apparent depth of a tank of water change if viewed
obliquely? If so, does the apparent depth increase or decrease?
(e) The refractive index of diamond is much greater than that of
ordinary glass. Is this fact of some use to a diamond cutter?
9.19 The image of a small electric bulb fixed on the wall of a room
is to be obtained on the opposite wall 3m away by means of a
large convex lens. What is the maximum possible focal length of
the lens required for the purpose?
9.20 A screen is placed 90cm from an object. The image of the
object on the screen is formed by a convex lens at two different
locations separated by 20cm. Determine the focal length of the
lens.
9.21 (a) Determine the effective focal length of the combination of
the two lenses in Exercise 9.10, if they are placed 8.0cm apart
with their principal axes coincident. Does the answer depend on
which side of the combination a beam of parallel light is incident?
Is the notion of effective focal length of this system useful at all?
(b) An object 1.5 cm in size is placed on the side of the convex
lens in the arrangement (a) above. The distance between the
object and the convex lens is 40cm. Determine the magnification
produced by the two-lens system, and the size of the image.
9.24 For a normal eye, the far point is at infinity and the near point
of distinct vision is about 25cm in front of the eye. The cornea of
the eye provides a converging power of about 40 dioptres, and the
least converging power of the eye-lens behind the cornea is about
20 dioptres. From this rough data estimate the range of
accommodation (i.e., the range of converging power of the eye-
lens) of a normal eye.
9.25 Does short-sightedness (myopia) or long-sightedness (hyper-
metropia) imply necessarily that the eye has partially lost its ability
of accommodation? If not, what might cause these defects of
vision?
9.30 (a) At what distance should the lens be held from the figure in
Exercise 9.29 in order to view the squares distinctly with the
maximum possible magnifying power?
(b) What is the magnification in this case?
(c) Is the magnification equal to the magnifying power in this case?
Explain.
9.31 What should be the distance between the object in Exercise
9.30 and the magnifying glass if the virtual image of each square
in the figure is to have an area of 6.25 mm2. Would you be able to
see the squares distinctly with your eyes very close to the
magnifier?
[Note: Exercises 9.29 to 9.31 will help you clearly understand the
difference between magnification in absolute size and the angular
magnification (or magnifying power) of an instrument.]
9.32 Answer the following questions:
(b) the final image is formed at the least distance of distinct vision
(25cm)?
9.35 (a) For the telescope described in Exercise 9.34 (a), what is
the separation between the objective lens and the eyepiece?
(b) If this telescope is used to view a 100 m tall tower 3 km away,
what is the height of the image of the tower formed by the
objective lens?
(c) What is the height of the final image of the tower if it is formed
at 25cm?
Figure 9.36
Figure 9.37
Chapter Ten
Wave Optics
10.1 Introduction
Light travels in a straight line in Class VI; it does not do so in Class XII and
beyond! Surprised, arent you?
In school, you are shown an experiment in which you take three cardboards
with pinholes in them, place a candle on one side and look from the other side.
If the flame of the candle and the three pinholes are in a straight line, you can
see the candle. Even if one of them is displaced a little, you cannot see the
candle. This proves, so your teacher says, that light travels in a straight line.
In the present book, there are two consecutive chapters, one on ray optics and
the other on wave optics. Ray optics is based on rectilinear propagation of
light, and deals with mirrors, lenses, reflection, refraction, etc. Then you come
to the chapter on wave optics, and you are told that light travels as a wave,
that it can bend around objects, it can diffract and interfere, etc.
In optical region, light has a wavelength of about half a micrometre. If it
encounters an obstacle of about this size, it can bend around it and can be
seen on the other side. Thus a micrometre size obstacle will not be able to
stop a light ray. If the obstacle is much larger, however, light will not be able to
bend to that extent, and will not be seen on the other side.
This is a property of a wave in general, and can be seen in sound waves too.
The sound wave of our speech has a wavelength of about 50cm to 1 m. If it
meets an obstacle of the size of a few metres, it bends around it and reaches
points behind the obstacle. But when it comes across a larger obstacle of a
few hundred metres, such as a hillock, most of it is reflected and is heard as
an echo.
Then what about the primary school experiment? What happens there is that
when we move any cardboard, the displacement is of the order of a few
millimetres, which is much larger than the wavelength of light. Hence the
candle cannot be seen. If we are able to move one of the cardboards by a
micrometer or less, light will be able to diffract, and the candle will still be seen.
One could add to the first sentence in this box: It learns how to bend as it
grows up!
We will now use Huygens principle to derive the laws of refraction. Let
PP represent the surface separating medium 1 and medium 2, as
shown in Fig. 10.4.
sin i = (10.1)
and
sin r = (10.2)
where i and r are the angles of incidence and refraction, respectively.
Thus we obtain
(10.3)
From the above equation, we get the important result that if r < i (i.e., if
the ray bends toward the normal), the speed of the light wave in the
second medium (v2) will be less then the speed of the light wave in the
first medium (v1). This prediction is opposite to the prediction from the
corpuscular model of light and as later experiments showed, the
prediction of the wave theory is correct. Now, if c represents the speed
of light in vacuum, then,
(10.4)
and
n2 = (10.5)
are known as the refractive indices of medium 1 and medium 2,
respectively. In terms of the refractive indices, Eq. (10.3) can
be written as
n1 sin i = n2 sin r (10.6)
This is the Snells law of refraction. Further, if 1 and 2 denote the
wavelengths of light in medium 1 and medium 2, respectively and if
the distance BC is equal to 1 then the distance AE will be equal to
2 (because if the crest from B has reached C in time , then the crest
from A should have also reached E in time ); thus,
or
(10.7)
The above equation implies that when a wave gets refracted into a
denser medium (v1 > v2) the wavelength and the speed of propagation
decrease but the frequency (= v/) remains the same.
(10.8)
Thus, if i = ic then sin r = 1 and r = 90. Obviously, for i > ic, there can
not be any refracted wave. The angle ic is known as the critical angle
and for all angles of incidence greater than the critical angle, we will
not have any refracted wave and the wave will undergo what is known
as total internal reflection. The phenomenon of total internal reflection
and its applications was discussed in Section 9.4.
BC = v
Figure 10.7 Refraction of a plane wave by (a) a thin prism, (b) a convex lens. (c)
Reflection of a plane wave by a concave mirror.
(10.9)
The formula given above is valid only when the speed of the source is
small compared to that of light. A more accurate formula for the
Doppler effect which is valid even when the speeds are close to that of
light, requires the use of Einsteins special theory of relativity. The
Doppler effect for light is very important in astronomy. It is the basis
for the measurements of the radial velocities of distant galaxies.
or, vradial
= 306 km/s
Therefore, the galaxy is moving away from us.
Example 10.2
(a) When monochromatic light is incident on a surface separating two media,
the reflected and refracted light both have the same frequency as the incident
frequency. Explain why?
(b) When light travels from a rarer to a denser medium, the speed decreases.
Does the reduction in speed imply a reduction in the energy carried by the light
wave?
(c) In the wave picture of light, intensity of light is determined by the square of
the amplitude of the wave. What determines the intensity of light in the photon
picture of light.
Solution
(a) Reflection and refraction arise through interaction of incident light with the
atomic constituents of matter. Atoms may be viewed as
oscillators, which take up the frequency of the external agency (light) causing
forced oscillations. The frequency of light emitted by a charged oscillator
equals its frequency of oscillation. Thus, the frequency of scattered light equals
the frequency of incident light.
(b) No. Energy carried by a wave depends on the amplitude of the wave, not
on the speed of wave propagation.
(c) For a given frequency, intensity of light in the photon picture is determined
by the number of photons crossing an unit area per unit time.
10.4 Coherent and Incoherent Addition of Waves
(a)
(b)
Figure 10.8 (a) Two needles oscillating in phase in water represent two coherent
sources. (b) The pattern of displacement of water molecules at an instant on the
surface of water showing nodal N (no displacement) and antinodal A (maximum
displacement) lines.
y2 = a cos t
Thus, the resultant of displacement at P would be given by
y = y1 + y2 = 2 a cos t
Since the intensity is the proportional to the square of the amplitude,
the resultant intensity will be given by
I = 4 I0
where I0 represents the intensity produced by each one of the
individual sources; I0 is proportional to a2. In fact at any point on the
perpendicular bisector of S1S2, the intensity will be 4I0. The two
sources are said to interfere constructively and we have what is
referred to as constructive interference. We next consider a point Q
[Fig. 10.9(a)]
for which
S2Q S1Q = 2
The waves emanating from S1 will arrive exactly two cycles earlier
than the waves from S2 and will again be in phase [Fig. 10.9(a)]. Thus,
if the displacement produced by S1 is given by
y1 = a cos t
then the displacement produced by S2 will be given by
y2 = a cos (t 4) = a cos t
where we have used the fact that a path difference of 2 corresponds
to a phase difference of 4. The two displacements are once again in
phase and the intensity will again be 4 I0 giving rise to constructive
interference. In the above analysis we have assumed that the
distances S1Q and S2Q are much greater than d (which represents
the distance between S1 and S2) so that although S1Q and S2Q are
not equal, the amplitudes of the displacement produced by each wave
are very nearly the same.
We next consider a point R [Fig. 10.9(b)] for which
S2R S1R = 2.5
The waves emanating from S1 will arrive exactly two and a half cycles
later than the waves from S2 [Fig. 10.10(b)]. Thus if the displacement
produced by S1 is given by
y1 = a cos t
Figure 10.10 Locus of points for which S1P S2P is equal to zero, , 2, 3.
y1 = a cos t
then, the displacement produced by S2 would be
y2 = a cos (t + )
and the resultant displacement will be given by
y = y1 + y2
= a [cos t + cos (t +)]
= 2 a cos (/2) cos (t + /2)
The amplitude of the resultant displacement is 2a cos (/2) and
therefore the intensity at that point will be
I = 4 I0 cos2 (/2) (10.12)
If = 0, 2 , 4 , which corresponds to the condition given by
Eq. (10.10) we will have constructive interference leading to maximum
intensity. On the other hand, if = , 3, 5 [which
corresponds to the condition given by Eq. (10.11)] we will have
destructive interference leading to zero intensity. Now if the two
sources are coherent (i.e., if the two needles are going up and down
regularly) then the phase difference at any point will not change with
time and we will have a stable interference pattern; i.e., the positions
of maxima and minima will not change with time. However, if the two
needles do not maintain a constant phase difference, then the
interference pattern will also change with time and, if the phase
difference changes very rapidly with time, the positions of maxima and
minima will also vary rapidly with time and we will see a time-
averaged intensity distribution. When this happens, we will observe
an average intensity that will be given by
(10.13)
where angular brackets represent time averaging. Indeed it is shown
in Section 7.2 that if (t) varies randomly with time, the time-averaged
quantity < cos2 (/2) > will be 1/2. This is also intuitively obvious
because the function cos2 (/2) will randomly vary between 0 and 1
and the average value will be 1/2. The resultant intensity will be given
by
I = 2 I0 (10.14)
at all points.
Figure 10.11 If two sodium lamps illuminate two pinholes S1 and S2, the intensities
will add up and no interference fringes will be observed on the screen.
If x, d<<D then negligible error will be introduced if S2P + S1P (in the
denominator) is replaced by 2D. For example, for d = 0.1 cm, D = 100
cm, OP = 1 cm (which correspond to typical values for an interference
experiment using light waves), we have S2P + S1P = [(100)2 +
(1.05)2] + [(100)2 + (0.95)2]
200.01 cm
Thus if we replace S2P + S1P by 2 D, the error involved is about
0.005%. In this approximation, Eq. (10.16) becomes
x = xn = ; n = 0, 1, 2, ... (10.18)
x = xn = (n+ ) (10.19)
Thus dark and bright bands appear on the screen, as shown in Fig.
10.13. Such bands are called fringes. Equations (10.18) and (10.19)
show that dark and bright fringes are equally spaced and the distance
between two consecutive bright and dark fringes is given by
= xn+1 xn
or = (10.20)
which is the expression for the fringe width. Obviously, the central
point O (in Fig. 10.12) will be bright because S1O = S2O and it will
correspond to n = 0. If we consider the line perpendicular to the plane
of the paper and passing through O [i.e., along the y-axis] then all
points on this line will be equidistant from S1 and S2 and we will have
a bright central fringe which is a straight line as shown in Fig. 10.13. In
order to determine the shape of the interference pattern on the screen
we note that a particular fringe would correspond to the locus of points
with a constant value of S2P S1P. Whenever this constant is an
integral multiple of , the fringe will be bright and whenever it is an odd
integral multiple of /2 it will be a dark fringe. Now, the locus of the
point P lying in the x-y plane such that S2P S1P (= ) is a constant,
is a hyperbola. Thus the fringe pattern will strictly be a hyperbola;
however, if the distance D is very large compared to the fringe width,
the fringes will be very nearly straight lines as shown in Fig. 10.13.
The wave nature of light was demonstrated convincingly for the first
time in 1801 by Thomas Young by a wonderfully simple experiment.
He let a ray of sunlight into a dark room, placed a dark screen in front
of it, pierced with two small pinholes, and beyond this, at some
distance, a white screen. He then saw two darkish lines at both sides
of a bright line, which gave him sufficient encouragement to repeat the
experiment, this time with spirit flame as light source, with a little salt in
it to produce the bright yellow sodium light. This time he saw a number
of dark lines, regularly spaced; the first clear proof that light added to
light can produce darkness. This phenomenon is called interference.
Thomas Young had expected it because he believed in the wave
theory of light.
Figure 10.13 Computer generated fringe pattern produced by two point source
S1 and S2 on the screen GG(Fig. 10.12); (a) and (b) correspond to d = 0.005 mm
and 0.025 mm, respectively (both figures correspond to D= 5 cm and = 5 10
5 cm.) (Adopted from OPTICS by A. Ghatak, Tata McGraw Hill Publishing Co. Ltd.,
We should mention here that the fringes are straight lines although S1
and S2 are point sources. If we had slits instead of the point sources
(Fig. 10.14), each pair of points would have produced straight line
fringes resulting in straight line fringes with increased intensities.
* Dennis Gabor received the 1971 Nobel Prize in Physics for
discovering the principles of holography.
Example 10.3 Two slits are made one millimetre apart and the screen is
placed one metre away. What is the fringe separation when blue-green light of
wavelength 500 nm is used?
Figure 10.14 Photograph and the graph of the intensity distribution in Youngs double-slit
experiment.
10.6 DIFFRACTION
If we look clearly at the shadow cast by an opaque object, close to the
region of geometrical shadow, there are alternate dark and bright
regions just like in interference. This happens due to the phenomenon
of diffraction. Diffraction is a general characteristic exhibited by all
types of waves, be it sound waves, light waves, water waves or matter
waves. Since the wavelength of light is much smaller than the
dimensions of most obstacles; we do not encounter diffraction effects
of light in everyday observations. However, the finite resolution of our
eye or of optical instruments such as telescopes or microscopes is
limited due to the phenomenon of diffraction. Indeed the colours that
you see when a CD is viewed is due to diffraction effects. We will now
discuss the phenomenon of diffraction.
The path difference NP LP between the two edges of the slit can be
calculated exactly as for Youngs experiment. From Fig. 10.15,
NP LP = NQ
= a sin
a (10.21)
Similarly, if two points M1 and M2 in the slit plane are separated by y,
the path difference M2P M1P y. We now have to sum up equal,
coherent contributions from a large number of sources, each with a
different phase. This calculation was made by Fresnel using integral
calculus, so we omit it here. The main features of the diffraction
pattern can be understood by simple arguments.
At the central point C on the screen, the angle is zero. All path
differences are zero and hence all the parts of the slit contribute in
phase. This gives maximum intensity at C. Experimental observation
shown in Fig. 10.15 indicates that the intensity has a central maximum
at = 0 and other secondary maxima at l (n+1/2) /a, and has
minima (zero intensity) at l n/a,
n = 1, 2, 3, .... It is easy to see why it has minima at these values
of angle. Consider first the angle where the path difference a is .
Then,
. (10.22)
Now, divide the slit into two equal halves LM and MN each of size a/2.
For every point M1 in LM, there is a point M2 in MN such that M1M2 =
a/2. The path difference between M1 and M2 at P = M2P M1P = a/2
= /2 for the angle chosen. This means that the contributions from M1
and M2 are 180 out of phase and cancel in the direction = /a.
Contributions from the two halves of the slit LM and MN, therefore,
cancel each other. Equation (10.22) gives the angle at which the
intensity falls to zero. One can similarly show that the intensity is zero
for = n/a, with n being any integer (except zero!). Notice that the
angular size of the central maximum increases when the slit width a
decreases.
Figure 10.15 The geometry of path differences for diffraction by a single slit.
(10.23)
The first two-thirds of the slit can therefore be divided into two halves
which have a /2 path difference. The contributions of these two
halves cancel in the same manner as described earlier. Only the
remaining one-third of the slit contributes to the intensity at a point
between the two minima. Clearly, this will be much weaker than the
central maximum (where the entire slit contributes in phase). One can
similarly show that there are maxima at (n + 1/2) /a with n = 2, 3, etc.
These become weaker with increasing n, since only one-fifth, one-
seventh, etc., of the slit contributes in these cases. The photograph
and intensity pattern corresponding to it is shown in Fig. 10.16.
Solution We want
Notice that the wavelength of light and distance of the screen do not
enter in the calculation of a.
Figure 10.17 The actual double-slit interference pattern. The envelope shows the
single slit diffraction.
(i) The interference pattern has a number of equally spaced bright and
dark bands. The diffraction pattern has a central bright maximum
which is twice as wide as the other maxima. The intensity falls as we
go to successive maxima away from the centre, on either side.
(ii) We calculate the interference pattern by superposing two waves
originating from the two narrow slits. The diffraction pattern is a
superposition of a continuous family of waves originating from each
point on a single slit.
(iii) For a single slit of width a, the first null of the interference pattern
occurs at an angle of /a. At the same angle of /a, we get a maximum
(not a null) for two narrow slits separated by a distance a.
Keep the slit parallel to the filament, right in front of the eye. Use
spectacles if you normally do. With slight adjustment of the width of
the slit and the parallelism of the edges, the pattern should be seen
with its bright and dark bands. Since the position of all the bands
(except the central one) depends on wavelength, they will show some
colours. Using a filter for red or blue will make the fringes clearer. With
both filters available, the wider fringes for red compared to blue can be
seen.
In this experiment, the filament plays the role of the first slit S in
Fig. 10.16. The lens of the eye focuses the pattern on the screen (the
retina of the eye).
Figure 10.18 Holding two blades to form a single slit. A bulb filament viewed
through this shows clear diffraction bands.
With some effort, one can cut a double slit in an aluminium foil with a
blade. The bulb filament can be viewed as before to repeat Youngs
experiment. In daytime, there is another suitable bright source
subtending a small angle at the eye. This is the reflection of the Sun in
any shiny convex surface (e.g., a cycle bell). Do not try direct sunlight
it can damage the eye and will not give fringes anyway as the Sun
subtends an angle of (1/2).
In interference and diffraction, light energy is redistributed. If it reduces
in one region, producing a dark fringe, it increases in another region,
producing a bright fringe. There is no gain or loss of energy, which is
consistent with the principle of conservation of energy.
10.6.3 Resolving power of optical instruments
In Chapter 9 we had discussed about telescopes. The angular
resolution of the telescope is determined by the objective of the
telescope. The stars which are not resolved in the image produced by
the objective cannot be resolved by any further magnification
produced by the eyepiece. The primary purpose of the eyepiece is to
provide magnification of the image produced by the objective.
(10.24)
where f is the focal length of the lens and 2a is the diameter of the
circular aperture or the diameter of the lens, whichever is smaller.
Typically if
0.5 m, f 20 cm and a 5 cm
we have
r0 1.2 m
Although the size of the spot is very small, it plays an important role in
determining the limit of resolution of optical instruments like a
telescope or a microscope. For the two stars to be just resolved
implying
(10.25)
Thus will be small if the diameter of the objective is large. This
implies that the telescope will have better resolving power if a is large.
It is for this reason that for better resolution, a telescope must have a
large diameter objective.
Example 10.6 Assume that light of wavelength 6000 is coming from a star.
What is the limit of resolution of a telescope whose objective has a diameter of
100 inch?
Solution A 100 inch telescope implies that 2a = 100 inch
= 254 cm. Thus if,
6000 = 6105 cm
then
radians
(10.27)
Two objects whose images are closer than this distance will not be
resolved, they will be seen as one. The corresponding minimum
separation, dmin, in the object plane is given by
dmin =
=
= (10.28)
(10.29)
If the medium between the object and the objective lens is not air but
a medium of refractive index n, Eq. (10.29) gets modified to
(10.30)
The product n sin is called the numerical aperture and is sometimes
marked on the objective.
Now watch the pattern, preferably with one eye. By moving away or closer to
the wall, find the position where you can just see some two black stripes as
separate stripes. All the black stripes to the left of this stripe would merge into
one another and would not be distinguishable. On the other hand, the black
stripes to the right of this would be more and more clearly visible. Note the
width d of the white stripe which separates the two regions, and measure the
distance D of the wall from your eye. Then d/D is the resolution of your eye.
You have watched specks of dust floating in air in a sunbeam entering through
your window. Find the distance (of a speck) which you can clearly see and
distinguish from a neighbouring speck. Knowing the resolution of your eye and
the distance of the speck, estimate the size of the speck of dust.
(10.31)
We define a quantity zF called the Fresnel distance by the following
equation
Equation (10.31) shows that for distances much smaller than zF , the
spreading due to diffraction is smaller compared to the size of the
beam. It becomes comparable when the distance is approximately zF.
For distances much greater than zF, the spreading due to diffraction
dominates over that due to ray optics (i.e., the size a of the aperture).
Equation (10.31) also shows that ray optics is valid in the limit of
wavelength tending to zero.
Example 10.7 For what distance is ray optics a good approximation when the
aperture is 3 mm wide and the wavelength is 500 nm?
Solution
This example shows that even with a small aperture, diffraction spreading can
be neglected for rays many metres in length. Thus, ray optics is valid in many
common situations.
10.7 POLARISATION
Consider holding a long string that is held horizontally, the other end
of which is assumed to be fixed. If we move the end of the string up
and down in a periodic manner, we will generate a wave propagating
in the +x direction (Fig. 10.21). Such a wave could be described by the
following equation
y (x,t) = a sin (kx t) (10.32)
where a and (= 2) represent the amplitude and the angular
frequency of the wave, respectively; further,
(10.33)
represents the wavelength associated with the wave. We had
discussed propagation of such waves in Chapter 15 of Class XI
textbook. Since the displacement (which is along the y direction) is at
right angles to the direction of propagation of the wave, we have what
is known as a transverse wave. Also, since the displacement is in the
y direction, it is often referred to as a y-polarised wave. Since each
point on the string moves on a straight line, the wave is also referred
to as a linearly polarised wave. Further, the string always remains
confined to the x-y plane and therefore it is also referred to as a plane
polarised wave.
Figure 10.21 (a) The curves represent the displacement of a string at t = 0 and
at t = t, respectively when a sinusoidal wave is propagating in the +x-direction. (b)
The curve represents the time variation of the displacement at x = 0 when a
sinusoidal wave is propagating in the +x-direction. At x = x, the time variation of
the displacement will be slightly displaced to the right.
Example 10.8 Discuss the intensity of transmitted light when a polaroid sheet
is rotated between two crossed polaroids?
Solution Let I0 be the intensity of polarised light after passing through the first
polariser P1. Then the intensity of light after passing through second polariser
P2 will be
,
where is the angle between pass axes of P1 and P2. Since P1 and P3 are
crossed the angle between the pass axes of P2 and P3 will be (/2). Hence
the intensity of light emerging from P3 will be
Figure 10.23 (a) Polarisation of the blue scattered light from the sky. The incident
sunlight is unpolarised (dots and arrows). A typical molecule is shown. It scatters
light by 90 polarised normal to the plane of the paper (dots only). (b) Polarisation
of light reflected from a transparent medium at the Brewster angle (reflected ray
perpendicular to refracted ray).
(10.36)
This is known as Brewsters law.
Summary
Points to Ponder
1. Waves from a point source spread out in all directions, while light was seen
to travel along narrow rays. It required the insight and experiment of Huygens,
Young and Fresnel to understand how a wave theory could explain all aspects
of the behaviour of light.
2. The crucial new feature of waves is interference of amplitudes from different
sources which can be both constructive and destructive, as shown in Youngs
experiment.
3. Even a wave falling on single slit should be regarded as a large number of
sources which interefere constructively in the forward direction ( = 0), and
destructively in other directions.
4. Diffraction phenomena define the limits of ray optics. The limit of the ability
of microscopes and telescopes to distinguish very close objects is set by the
wavelength of light.
5. Most interference and diffraction effects exist even for longitudinal waves
like sound in air. But polarisation phenomena are special to transverse waves
like light waves.
* Richand Feynman was one of the recipients of the 1965 Nobel Prize
in Physics for his fundamental work in quantum electrodynamics.
Exercises
10.1 Monochromatic light of wavelength 589 nm is incident from
air on a water surface. What are the wavelength, frequency and
speed of
(a) reflected, and (b) refracted light? Refractive index of water is
1.33.
10.2 What is the shape of the wavefront in each of the following
cases:
(a) Light diverging from a point source.
(b) Light emerging out of a convex lens when a point source is
placed at its focus.
(c) The portion of the wavefront of light from a distant star
intercepted by the Earth.
10.3 (a) The refractive index of glass is 1.5. What is the speed of
light in glass? (Speed of light in vacuum is 3.0 108 m s1)
10.15 For sound waves, the Doppler formula for frequency shift
differs slightly between the two situations: (i) source at rest;
observer moving, and (ii) source moving; observer at rest. The
exact Doppler formulas for the case of light waves in vacuum are,
however, strictly identical for these situations. Explain why this
should be so. Would you expect the formulas to be strictly identical
for the two situations in case of light travelling in a medium?
10.16 In double-slit experiment using light of wavelength 600 nm,
the angular width of a fringe formed on a distant screen is 0.1.
What is the spacing between the two slits?
10.17 Answer the following questions:
(a) In a single slit diffraction experiment, the width of the slit is
made double the original width. How does this affect the size and
intensity of the central diffraction band?
(b) In what way is diffraction from each slit related to the
interference pattern in a double-slit experiment?
(c) When a tiny circular obstacle is placed in the path of light from
a distant source, a bright spot is seen at the centre of the shadow
of the obstacle. Explain why?
(d) Two students are separated by a 7 m partition wall in a room
10 m high. If both light and sound waves can bend around
obstacles, how is it that the students are unable to see each other
even though they can converse easily.
(e) Ray optics is based on the assumption that light travels in a
straight line. Diffraction effects (observed when light propagates
through small apertures/slits or around small obstacles) disprove
this assumption. Yet the ray optics assumption is so commonly
used in understanding location and several other properties of
images in optical instruments. What is the justification?
10.18 Two towers on top of two hills are 40 km apart. The line
joining them passes 50 m above a hill halfway between the
towers. What is the longest wavelength of radio waves, which can
be sent between the towers without appreciable diffraction effects?
10.19 A parallel beam of light of wavelength 500 nm falls on a
narrow slit and the resulting diffraction pattern is observed on a
screen 1 m away. It is observed that the first minimum is at a
distance of 2.5 mm from the centre of the screen. Find the width of
the slit.
10.20 Answer the following questions:
(a) When a low flying aircraft passes overhead, we sometimes
notice a slight shaking of the picture on our TV screen. Suggest a
possible explanation.
(b) As you have learnt in the text, the principle of linear
superposition of wave displacement is basic to understanding
intensity distributions in diffraction and interference patterns. What
is the justification of this principle?
10.21 In deriving the single slit diffraction pattern, it was stated that
the intensity is zero at angles of n/a. Justify this by suitably
dividing the slit to bring out the cancellation.
Chapter Eleven
11.1 Introduction
11.7 Calorimetry
Summary
Points to ponder
Exercises
11.1 INTRODUCTION
We all have common-sense notions of heat and temperature.
Temperature is a measure of hotness of a body. A kettle with boiling
water is hotter than a box containing ice. In physics, we need to define
the notion of heat, temperature, etc., more carefully. In this chapter,
you will learn what heat is and how it is measured, and study the
various proceses by which heat flows from one body to another. Along
the way, you will find out why blacksmiths heat the iron ring before
fitting on the rim of a wooden wheel of a bullock cart and why the wind
at the beach often reverses direction after the sun goes down. You will
also learn what happens when water boils or freezes, and its
temperature does not change during these processes even though a
great deal of heat is flowing into or out of it.
Fig. 11.1 A plot of Fahrenheit temperature (tF) versus Celsius temperature (tc).
(11.1)
Fig. 11.2 Pressure versus temperature of a low density gas kept at constant
volume.
This relationship is known as ideal gas law. It can be written in a more
general form that applies not just to a given quantity of a single gas
but to any quantity of any dilute gas and is known as IDEAL-GAS
EQUATION:
or PV = RT (11.2)
where, is the number of moles in the sample of gas and R is called
universal gas constant:
R = 8.31 J mol1 K1
In Eq. 11.2, we have learnt that the pressure and volume are directly
proportional to temperature : PV T. This relationship allows a gas to
be used to measure temperature in a constant volume gas
thermometer. Holding the volume of a gas constant, it gives P T.
Thus, with a constant-volume gas thermometer, temperature is read in
terms of pressure. A plot of pressure versus temperature gives a
straight line in this case, as shown in Fig. 11.2.
However, measurements on real gases deviate from the values
predicted by the ideal gas law at low temperature. But the relationship
is linear over a large temperature range, and it looks as though the
pressure might reach zero with decreasing temperature if the gas
continued to be a gas. The absolute minimum temperature for an ideal
gas, therefore, inferred by extrapolating the straight line to the axis, as
in Fig. 11.3. This temperature is found to be 273.15 C and is
designated as ABSOLUTE ZERO. Absolute zero is the foundation of
the Kelvin temperature scale or absolute scale temperature named
after the British scientist Lord Kelvin. On this scale, 273.15 C is
taken as the zero point, that is 0 K (Fig. 11.4).
Fig. 11.3 A plot of pressure versus temperature and extrapolation of lines for low
density gases indicates the same absolute zero temperature.
Fig. 11.4 Comparision of the Kelvin, Celsius and Fahrenheit temperature scales.
The size of the unit for Kelvin temperature is the same celsius degree,
so temperature on these scales are related by
T = tC + 273.15 (11.3)
11.5 THERMAL EXPANSION
You may have observed that sometimes sealed bottles with metallic
lids are so tightly screwed that one has to put the lid in hot water for
sometime to open the lid. This would allow the metallic cover to
expand, thereby loosening it to unscrew easily. In case of liquids, you
may have observed that mercury in a thermometer rises, when the
thermometer is put in a slightly warm water. If we take out the
thermometer from the warm water the level of mercury falls again.
Similarly, in the case of gases, a balloon partially inflated in a cool
room may expand to full size when placed in warm water. On the
other hand, a fully inflated balloon when immersed in cold water would
start shrinking due to contraction of the air inside.
It is our common experience that most substances expand on heating
and contract on cooling. A change in the temperature of a body
causes change in its dimensions. The increase in the dimensions of a
body due to the increase in its temperature is called thermal
expansion. The expansion in length is called LINEAR EXPANSION.
The expansion in area is called AREA EXPANSION. The expansion in
volume is called VOLUME EXPANSION (Fig. 11.5).
If the substance is in the form of a long rod, then for small change in
temperature, T, the fractional change in length, l/l, is directly
proportional to T.
(11.4)
where 1 is known as the COEFFICIENT OF LINEAR EXPANSION
and is characteristic of the material of the rod. In Table 11.1 are given
typical average values of the coefficient of linear expansion for some
materials in the temperature range 0 C to 100 C. From this Table,
compare the value of l for glass and copper. We find that copper
expands about five times more than glass for the same rise in
temperature. Normally, metals expand more and have relatively high
values of l.
Table 11.1 Values of coefficient of linear expansion for some materials
OF VOLUME EXPANSION, as
(11.5)
PV = RT
At constant pressure
PV = R T
(a) (b)
Fig. 11.7 Thermal expansion of water.
In equation (11.7), terms in (l)2 and (l)3 have been neglected since
l is small compared to l. So
(11.8)
which gives
v = 3l (11.9)
What happens by preventing the thermal expansion of a rod by fixing
its ends rigidly? Clearly, the rod acquires a compressive strain due to
the external forces provided by the rigid support at the ends. The
corresponding stress set up in the rod is called THERMAL STRESS.
For example, consider a steel rail of length 5 m and area of cross
section 40 cm2 that is prevented from expanding while the
temperature rises by 10 C. The coefficient of linear expansion of steel
Answer
Fig. 11.8
Hence,
Example 11.2 A blacksmith fixes iron ring on the rim of the wooden wheel of a
bullock cart. The diameter of the rim and the iron ring are 5.243 m and 5.231
m respectively at 27 C. To what temperature should the ring be heated so as
to fit the rim of the wheel?
Answer
Given, T1 = 27 C
LT1 = 5.231 m
LT2 = 5.243 m
So,
In the second step, now suppose you take double the amount of water
and heat it, using the same heating arrangement, to raise the
temperature by 20 C, you will find the time taken is again twice that
required in the first step.
In the third step, in place of water, now heat the same quantity of
some oil, say mustard oil, and raise the temperature again by 20 C.
Now note the time by the same stopwatch. You will find the time taken
will be shorter and therefore, the quantity of heat required would be
less than that required by the same amount of water for the same rise
in temperature.
(11.10)
where Q is the amount of heat supplied to the substance to change
its temperature from T to T + T.
(11.11)
The SPECIFIC HEAT CAPACITY is the property of the substance
which determines the change in the temperature of the substance
(undergoing no phase change) when a given quantity of heat is
absorbed (or rejected) by it. It is defined as the amount of heat per unit
mass absorbed or rejected by the substance to change its
temperature by one unit. It depends on the nature of the substance
and its temperature. The SI unit of specific heat capacity is J kg1 K1.
If the amount of substance is specified in terms of moles , instead of
mass m in kg, we can define heat capacity per mole of the substance
by
(11.12)
where C is known as MOLAR SPECIFIC HEAT CAPACITY of the
substance. Like S, C also depends on the nature of the substance and
its temperature. The SI unit of molar specific heat capacity is J mol1
K1.
11.7 CALORIMETRY
A system is said to be isolated if no exchange or transfer of heat
occurs between the system and its surroundings. When different parts
of an isolated system are at different temperature, a quantity of heat
transfers from the part at higher temperature to the part at lower
temperature. The heat lost by the part at higher temperature is equal
to the heat gained by the part at lower temperature.
Calorimetry means measurement of heat. When a body at higher
temperature is brought in contact with another body at lower
temperature, the heat lost by the hot body is equal to the heat gained
by the colder body, provided no heat is allowed to escape to the
surroundings. A device in which heat measurement can be made is
called a CALORIMETER. It consists a metallic vessel and stirrer of the
same material like copper or alumiunium. The vessel is kept inside a
wooden jacket which contains heat insulating materials like glass wool
etc. The outer jacket acts as a heat shield and reduces the heat loss
from the inner vessel. There is an opening in the outer jacket through
which a mercury thermometer can be inserted into the calorimeter.
The following example provides a method by which the specific heat
capacity of a given solid can be determinated by using the principle,
heat gained is equal to the heat lost.
Fig. 11.9 A plot of temperature versus time showing the changes in the state of ice
on heating (not to scale).
The change of state from solid to liquid is called MELTING and from
liquid to solid is called FUSION. It is observed that the temperature
remains constant until the entire amount of the solid substance melts.
That is, BOTH THE SOLID AND LIQUID STATES OF THE
SUBSTANCE COEXIST IN THERMAL EQUILIBRIUM DURING THE
CHANGE OF STATES FROM SOLID TO LIQUID. The temperature at
which the solid and the liquid states of the substance in thermal
equilibrium with each other is called its MELTING POINT. It is
characteristic of the substance. It also depends on pressure. The
melting point of a substance at standard atomspheric pressure is
called its NORMAL MELTING POINT. Let us do the following activity
to understand the process of melting of ice.
Take a slab of ice. Take a metallic wire and fix two blocks, say 5 kg
each, at its ends. Put the wire over the slab as shown in Fig. 11.10.
You will observe that the wire passes through the ice slab. This
happens due to the fact that just below the wire, ice melts at lower
temperature due to increase in pressure. When the wire has passed,
water above the wire freezes again. Thus the wire passes through the
slab and the slab does not split. This phenomenon of refreezing is
called REGELATION. Skating is possible on snow due to the
formation of water below the skates. Water is formed due to the
increase of pressure and it acts as a lubricant.
Fig. 11.10
After the whole of ice gets converted into water and as we continue
further heating, we shall see that temperature begins to rise. The
temperature keeps on rising till it reaches nearly 100 C when it again
becomes steady. The heat supplied is now being utilised to change
water from liquid state to vapour or gaseous state.
The change of state from liquid to vapour (or gas) is called
VAPORISATION. It is observed that the temperature remains constant
until the entire amount of the liquid is converted into vapour. That is,
both the liquid and vapour states of the substance coexist in thermal
equilibrium, during the change of state from liquid to vapour. The
temperature at which the liquid and the vapour states of the substance
coexist is called its BOILING POINT. Let us do the following activity to
understand the process of boiling of water.
Triple Point
(a) (b)
Pressure-temperature phase diagrams for (a) water and (b) CO2 (not to the scale).
Take a round-bottom flask, more than half filled with water. Keep it
over a burner and fix a thermometer and steam outlet through the cork
of the flask (Fig. 11.11). As water gets heated in the flask, note first
that the air, which was dissolved in the water, will come out as small
bubbles. Later, bubbles of steam will form at the bottom but as they
rise to the cooler water near the top, they condense and disappear.
Finally, as the temperature of the entire mass of the water reaches
100 C, bubbles of steam reach the surface and boiling is said to
occur. The steam in the flask may not be visible but as it comes out of
the flask, it condenses as tiny droplets of water, giving a foggy
appearance.
If now the steam outlet is closed for a few seconds to increase the
pressure in the flask, you will notice that boiling stops. More heat
would be required to raise the temperature (depending on the
increase in pressure) before boiling begins again. Thus boiling point
increases with increase in pressure.
Let us now remove the burner. Allow water to cool to about 80 C.
Remove the thermometer and steam outlet. Close the flask with the
airtight cork. Keep the flask turned upside down on the stand. Pour
ice-cold water on the flask. Water vapours in the flask condense
reducing the pressure on the water surface inside the flask. Water
begins to boil again, now at a lower temperature. Thus boiling point
decreases with decrease in pressure.
This explains why cooking is difficult on hills. At high altitudes,
atmospheric pressure is lower, reducing the boiling point of water as
compared to that at sea level. On the other hand, boiling point is
increased inside a pressure cooker by increasing the pressure. Hence
cooking is faster. The boiling point of a substance at standard
atmospheric pressure is called its NORMAL BOILING POINT.
However, all substances do not pass through the three states: solid-
liquid-gas. There are certain substances which normally pass from the
solid to the vapour state directly and vice versa. The change from
solid state to vapour state without passing through the liquid state is
called SUBLIMATION, and the substance is said to sublime. Dry ice
(solid CO2) sublimes, so also iodine. During the sublimation process
both the solid and vapour states of a substance coexist in thermal
equilibrium.
Fig. 11.12 Temperature versus heat for water at 1 atm pressure (not to scale).
Note that when heat is added (or removed) during a change of state,
the temperature remains constant. Note in Fig. 11.12 that the slopes
of the phase lines are not all the same, which indicates that specific
heats of the various states are not equal. For water, the latent heat of
fusion and vaporisation are Lf = 3.33 105 J kg1 and Lv = 22.6 105
J kg1 respectively. That is 3.33 105 J of heat are needed to melt 1
kg of ice at 0 C, and 22.6 105 J of heat are needed to convert 1 kg
of water to steam at 100 C. So, steam at 100 C carries 22.6 105 J
kg1 more heat than water at 100 C. This is why burns from steam
are usually more serious than those from boiling water.
Answer
Heat lost by water = msw (fi)w
= (0.30 kg) (4186 J kg1 K1) (50.0 C 6.7 C)
= 54376.14 J
Heat required to melt ice = m2Lf = (0.15 kg) Lf
Heat required to raise temperature of ice water to final temperature =
mIsw (fi)I
= (0.15 kg) (4186 J kg1 K 1) (6.7 C 0 C)
= 4206.93 J
Heat lost = heat gained
= 75600J + 1005000 J
+ 1255800 J + 6768000 J
= 9.1106 J t
11.9.1 Conduction
Conduction is the mechanism of transfer of heat between two adjacent
parts of a body because of their temperature difference. Suppose one
end of a metallic rod is put in a flame, the other end of the rod will
soon be so hot that you cannot hold it by your bare hands. Here heat
transfer takes place by conduction from the hot end of the rod through
its different parts to the other end. Gases are poor thermal conductors
while liquids have conductivities intermediate between solids and
gases.
Heat conduction may be described quantitatively as the time rate of
heat flow in a material for a given temperature difference. Consider a
metallic bar of length L and uniform cross section A with its two ends
maintained at different temperatures. This can be done, for example,
by putting the ends in thermal contact with large reservoirs at
temperatures, say, TC and TD respectively (Fig. 11.14). Let us assume
the ideal condition that the sides of the bar are fully insulated so that
no heat is exchanged between the sides and the surroundings.
After sometime, a steady state is reached; the temperature of the bar
decreases uniformly with distance from TC to TD; (TC>TD). The
reservoir at C supplies heat at a constant rate, which transfers through
the bar and is given out at the same rate to the reservoir at D. It is
found experimentally that in this steady state, the rate of flow of heat
(or heat current) H is proportional to the temperature difference (TC
TD) and the area of cross section A and is inversely proportional to the
length L :
H = KA (11.14)
The constant of proportionality K is called the THERMAL
CONDUCTIVITY of the material. The greater the value of K for a
material, the more rapidly will it conduct heat. The SI unit of K is J S1
m1 K1 or W m1 K1. The thermal conductivities of various
substances are listed in Table 11.5. These values vary slightly with
temperature, but can be considered to be constant over a normal
temperature range.
Fig. 11.14 Steady state heat flow by conduction in a bar with its two ends
maintained at temperatures TC and TD; (TC > TD).
Fig. 11.15
Answer The insulating material around the rods reduces heat loss
from the sides of the rods. Therefore, heat flows only along the length
of the rods. Consider any cross section of the rod. In the steady state,
heat flowing into the element must equal the heat flowing out of it;
otherwise there would be a net gain or loss of heat by the element and
its temperature would not be steady. Thus in the steady state, rate of
heat flowing across a cross section of the rod is the same at every
point along the length of the combined steel-copper rod. Let T be the
temperature of the steel-copper junction in the steady state. Then,
where 1 and 2 refer to the steel and copper rod respectively. For A1 =
2 A2, L1 = 15.0 cm, L2 = 10.0 cm, K1 = 50.2 J s1 m1 K 1, K2 = 385 J
s1 m1 K 1, we have
Example 11.7 An iron bar (L1 = 0.1 m, A1 = 0.02 m2, K1 = 79 W m1 K1) and
a brass bar (L2 = 0.1 m, A2 = 0.02 m2, K2 = 109 W m1K1) are soldered end
to end as shown in Fig. 11.16. The free ends of the iron bar and brass bar are
maintained at 373 K and 273 K respectively. Obtain expressions for and hence
compute (i) the temperature of the junction of the two bars, (ii) the equivalent
thermal conductivity of the compound bar, and (iii) the heat current through the
compound bar.
Answer
Fig 11.16
=
For A1 = A2 = A and L1 = L2 = L, this equation leads to
K1 (T1 T0) = K2 (T0 T2)
Thus the junction temperature T0 of the two bars is
T0 =
Using this equation, the heat current H through either bar is
H=
Using these equations, the heat current H through the compound bar
of length L1 + L2 = 2L and the equivalent thermal conductivity K, of the
compound bar are given by
(i)
= 315 K
(ii)
=
= 91.6 W m1 K1
(iii)
= 916.1 W t
11.9.2 Convection
Convection is a mode of heat transfer by actual motion of matter. It is
possible only in fluids. Convection can be natural or forced. In natural
convection, gravity plays an important part. When a fluid is heated
from below, the hot part expands and, therefore, becomes less dense.
Because of buoyancy, it rises and the upper colder part replaces it.
This again gets heated, rises up and is replaced by the colder part of
the fluid. The process goes on. This mode of heat transfer is evidently
different from conduction. Convection involves bulk transport of
different parts of the fluid. In forced convection, material is forced to
move by a pump or by some other physical means. The common
examples of forced convection systems are forced-air heating systems
in home, the human circulatory system, and the cooling system of an
automobile engine. In the human body, the heart acts as the pump
that circulates blood through different parts of the body, transferring
heat by forced convection and maintaining it at a uniform temperature.
Natural convection is responsible for many familiar phenomena.
During the day, the ground heats up more quickly than large bodies of
water do. This occurs both because the water has a greater specific
heat and because mixing currents disperse the absorbed heat
throughout the great volume of water. The air in contact with the warm
ground is heated by conduction. It expands, becoming less dense
than the surrounding cooler air. As a result, the warm air rises (air
currents) and other air moves (winds) to fill the space-creating a sea
breeze near a large body of water. Cooler air descends, and a thermal
convection cycle is set up, which transfers heat away from the land. At
night, the ground loses its heat more quickly, and the water surface is
warmer than the land. As a result, the cycle is reveresed (Fig. 11.17).
Take some water, say 300 ml, in a calorimeter with a stirrer and cover
it with two holed lid. Fix a thermometer through a hole in the lid and
make sure that the bulb of thermometer is immersed in the water.
Note the reading of the thermometer. This reading T1 is the
temperature of the surroundings. Heat the water kept in the
calorimeter till it attains a temperature, say, 40 C above room
temperature (i.e., temperature of the surroundings). Then stop heating
the water by removing the heat source. Start the stop-watch and note
the reading of the thermometer after fixed interval of time, say after
every one minute of stirring gently with the stirrer. Continue to note the
temperature (T2) of water till it attains a temperature about 5 C above
that of the surroundings. Then plot a graph by taking each value of
temperature T = T2 T1 along y axis and the coresponding value of t
along x-axis (Fig. 11.18).
From the graph you will infer how the cooling of hot water depends on
the difference of its temperature from that of the surroundings. You will
also notice that initially the rate of cooling is higher and decreases as
the temperature of the body falls.
The above activity shows that a hot body loses heat to its
surroundings in the form of heat radiation. The rate of loss of heat
depends on the difference in temperature between the body and its
surroundings. Newton was the first to study, in a systematic manner,
the relation between the heat lost by a body in a given enclosure and
its temperature.
According to Newtons law of cooling, the rate of loss of heat, dQ/dt
of the body is directly proportional to the difference of temperature T
= (T2T1) of the body and the surroundings. The law holds good only
for small difference of temperature. Also, the loss of heat by radiation
depends upon the nature of the surface of the body and the area of
the exposed surface. We can write
(11.15)
where k is a positive constant depending upon the area and nature of
the surface of the body. Suppose a body of mass m and specific heat
capacity s is at temperature T2. Let T1 be the temperature of the
surroundings. If the temperature falls by a small amount dT2 in time dt,
then the amount of heat lost is
dQ = ms dT2
Rate of loss of heat is given by
(11.16)
From Eqs. (11.15) and (11.16) we have
(11.17)
where K = k/m s
On integrating,
loge (T2 T1) = K t + c (11.18)
or T2 = T1 + C eKt; where C = ec (11.19)
Equation (11.19) enables you to calculate the time of cooling of a body
through a particular range of temperature.
For small temperature differences, the rate of cooling, due to
conduction, convection, and radiation combined, is proportional to the
difference in temperature. It is a valid approximation in the transfer of
heat from a radiator to a room, the loss of heat through the wall of a
room, or the cooling of a cup of tea on the table.
= K (50 C)
When we divide above two equations, we have
SUMMARY
1. Heat is a form of energy that flows between a body and its surrounding
medium by virtue of temperature difference between them. The degree of
hotness of the body is quantitatively represented by temperature.
2. A temperature-measuring device (thermometer) makes use of some
measurable property (called thermometric property) that changes with
temperature. Different thermometers lead to different temperature scales. To
construct a temperature scale, two fixed points are chosen and assigned some
arbitrary values of temperature. The two numbers fix the origin of the scale
and the size of its unit.
3. The Celsius temperature (tC) and the Farenheit temperare (tF)are related by
tF = (9/5) tC + 32
4. The ideal gas equation connecting pressure (P), volume (V) and absolute
temperature (T) is :
PV = RT
5. In the absolute temperature scale, the zero of the scale is the absolute zero
of temperature the temperature where every substance in nature has the
least possible molecular activity. The Kelvin absolute temperature scale (T )
has the same unit size as the Celsius scale (Tc ), but differs in the origin :
TC = T 273.15
where l and V denote the change in length l and volume V for a change of
temperature T. The relation between them is :
v = 3 l
7. The specific heat capacity of a substance is defined by
where m is the mass of the substance and Q is the heat required to change
its temperature by T. The molar specific heat capacity of a substance is
defined by
8. The latent heat of fusion (Lf) is the heat per unit mass required to change a
substance from solid into liquid at the same temperature and pressure. The
latent heat of vaporisation (Lv) is the heat per unit mass required to change a
substance from liquid to the vapour state without change in the temperature
and pressure.
9. The three modes of heat transfer are conduction, convection and radiation.
11. Newtons Law of Cooling says that the rate of cooling of a body is
proportional to the excess temperature of the body over the surroundings :
POINTS TO PONDER
T = tc + 273.15
and the assignment T = 273.16 K for the triple point of water are exact
relations (by choice). With this choice, the Celsius temperature of the melting
point of water and boiling point of water (both at 1 atm pressure) are very
close to, but not exactly equal to 0 C and 100 C respectively. In the original
Celsius scale, these latter fixed points were exactly at 0 C and 100 C (by
choice), but now the triple point of water is the preferred choice for the fixed
point, because it has a unique temperature.
2. A liquid in equilibrium with vapour has the same pressure and temperature
throughout the system; the two phases in equilibrium differ in their molar
volume (i.e. density). This is true for a system with any number of phases in
equilibrium.
3. Heat transfer always involves temperature difference between two systems
or two parts of the same system. Any energy transfer that does not involve
temperature difference in some way is not heat.
EXERCISES
11.1 The triple points of neon and carbon dioxide are 24.57 K and
216.55 K respectively. Express these temperatures on the Celsius
and Fahrenheit scales.
11.2 Two absolute scales A and B have triple points of water
defined to be 200 A and 350 B. What is the relation between TA
and TB ?
Thermodynamics
12.1 Introduction
Summary
Points to ponder
Exercises
12.1 INTRODUCTION
In previous chapter we have studied thermal properties of matter. In
this chapter we shall study laws that govern thermal energy. We shall
study the processes where work is converted into heat and vice versa.
In winter, when we rub our palms together, we feel warmer; here work
done in rubbing produces the heat. Conversely, in a steam engine,
the heat of the steam is used to do useful work in moving the pistons,
which in turn rotate the wheels of the train.
Fig. 12.1 (a) Systems A and B (two gases) separated by an adiabatic wall an
insulating wall that does not allow flow of heat. (b) The same systems A and B
separated by a diathermic wall a conducting wall that allows heat to flow from
one to another. In this case, thermal equilibrium is attained in due course.
The Zeroth Law clearly suggests that when two systems A and B, are
in thermal equilibrium, there must be a physical quantity that has the
same value for both. This thermodynamic variable whose value is
equal for two systems in thermal equilibrium is called temperature (T ).
Thus, if A and B are separately in equilibrium with C, TA = TC and TB =
TC. This implies that TA = TB i.e. the systems A and B are also in
thermal equilibrium.
We have arrived at the concept of temperature formally via the Zeroth
Law. The next question is :how to assign numerical values to
temperatures of different bodies ? In other words, how do we
construct a scale of temperature ? Thermometry deals with this basic
question to which we turn in the next section.
Fig. 12.2 (a) Systems A and B are separated by an adiabatic wall, while each is in
contact with a third system C via a conducting wall. (b) The adiabatic wall between
A and B is replaced by a conducting wall, while C is insulated from A and B by an
adiabatic wall.
Fig. 12.3 (a) Internal energy U of a gas is the sum of the kinetic and potential
energies of its molecules when the box is at rest. Kinetic energy due to various
types of motion (translational, rotational, vibrational) is to be included in U. (b) If
the same box is moving as a whole with some velocity, the kinetic energy of the
box is not to be included in U.
Fig. 12.4 Heat and work are two distinct modes of energy transfer to a system that
results in change in its internal energy. (a) Heat is energy transfer due to
temperature difference between the system and the surroundings. (b) Work is
energy transfer brought about by means (e.g. moving the piston by raising or
lowering some weight connected to it) that do not involve such a temperature
difference.
Q = U + W (12.1)
i.e. the energy (Q) supplied to the system goes in partly to increase
the internal energy of the system (U) and the rest in work on the
environment (W). Equation (12.1) is known as the FIRST LAW OF
THERMODYNAMICS. It is simply the general law of conservation of
energy applied to any system in which the energy transfer from or to
the surroundings is taken into account.
Let us put Eq. (12.1) in the alternative form
Q W = U (12.2)
Now, the system may go from an initial state to the final state in a
number of ways. For example, to change the state of a gas from (P1,
V1) to (P2, V2), we can first change the volume of the gas from V1 to
V2, keeping its pressure constant i.e. we can first go the state (P1, V2)
and then change the pressure of the gas from P1 to P2, keeping
volume constant, to take the gas to (P2, V2). Alternatively, we can first
keep the volume constant and then keep the pressure constant. Since
U is a state variable, U depends only on the initial and final states
and not on the path taken by the gas to go from one to the other.
However, Q and W will, in general, depend on the path taken to go
from the initial to final states. From the First Law of Thermodynamics,
Eq. (12.2), it is clear that the combination Q W, is however, path
independent. This shows that if a system is taken through a process in
which U = 0 (for example, isothermal expansion of an ideal gas, see
section 12.8),
Q = W
i.e., heat supplied to the system is used up entirely by the system in
doing work on the environment.
If the system is a gas in a cylinder with a movable piston, the gas in
moving the piston does work. Since force is pressure times area, and
area times displacement is volume, work done by the system against
a constant pressure P is
W = P V
where V is the change in volume of the gas. Thus, for this case, Eq.
(12.1) gives
Q = U + P V (12.3)
As an application of Eq. (12.3), consider the change in internal energy
for 1 g of water when we go from its liquid to vapour phase. The
measured latent heat of water is 2256 J/g. i.e., for 1 g of water Q =
2256 J. At atmospheric pressure, 1 g of water has a volume 1 cm3 in
liquid phase and 1671 cm3 in vapour phase.
Therefore,
(12.4)
We expect Q and, therefore, heat capacity S to be proportional to the
mass of the substance. Further, it could also depend on the
temperature, i.e., a different amount of heat may be needed for a unit
rise in temperature at different temperatures. To define a constant
characteristic of the substance and independent of its amount, we
divide S by the mass of the substance m in kg :
(12.5)
s is known as the specific heat capacity of the substance. It depends
on the nature of the substance and its temperature. The unit of
specific heat capacity is J kg1 K1.
If the amount of substance is specified in terms of moles (instead of
mass m in kg ), we can define heat capacity per mole of the substance
by
(12.6)
C is known as molar specific heat capacity of the substance. Like s, C
is independent of the amount of substance. C depends on the nature
of the substance, its temperature and the conditions under which heat
is supplied. The unit of C is J mo11 K1. As we shall see later (in
connection with specific heat capacity of gases), additional conditions
may be needed to define C or s. The idea in defining C is that simple
predictions can be made in regard to molar specific heat capacities.
Table 12.1 lists measured specific and molar heat capacities of solids
at atmospheric pressure and ordinary room temperature.
(12.7)
Cp Cv = R (12.8)
where Cp and Cv are molar specific heat capacities of an ideal gas at
constant pressure and volume respectively and R is the universal gas
constant. To prove the relation, we begin with Eq. (12.3) for 1 mole of
the gas :
Q = U + P V
If Q is absorbed at constant volume, V = 0
(12.9)
(12.10)
The subscript p can be dropped from the first term since U of an ideal
gas depends only on T. Now, for a mole of an ideal gas
PV = RT
which gives
(12.11)
Equations (12.9) to (12.11) give the desired relation, Eq. (12.8).
Isothermal process
For an isothermal process (T fixed), the ideal gas equation gives
PV = constant
i.e., pressure of a given mass of gas varies inversely as its volume.
This is nothing but Boyles Law.
Suppose an ideal gas goes isothermally (at temperature T ) from its
initial state (P1, V1) to the final state (P2, V2). At any intermediate
stage with pressure P and volume change from V to
V + V (V small)
W = P V
Taking (V 0) and summing the quantity W over the entire
process,
(12.12)
where in the second step we have made use of the ideal gas equation
PV = RT and taken the constants out of the integral. For an ideal
gas, internal energy depends only on temperature. Thus, there is no
change in the internal energy of an ideal gas in an isothermal process.
The First Law of Thermodynamics then implies that heat supplied to
the gas equals the work done by the gas : Q = W. Note from Eq.
(12.12) that for V2 > V1, W > 0; and for V2 < V1, W < 0. That is, in an
isothermal expansion, the gas absorbs heat and does work while in an
isothermal compression, work is done on the gas by the environment
and heat is released.
Adiabatic process
In an adiabatic process, the system is insulated from the surroundings
and heat absorbed or released is zero. From Eq. (12.1), we see that
work done by the gas results in decrease in its internal energy (and
hence its temperature for an ideal gas). We quote without proof (the
result that you will learn in higher courses) that for an adiabatic
process of an ideal gas.
P V = const (12.13)
where is the ratio of specific heats (ordinary or molar) at constant
pressure and at constant volume.
Figure12.8 shows the P-V curves of an ideal gas for two adiabatic
processes connecting two isotherms.
Fig. 12.8 P-V curves for isothermal and adiabatic processes of an ideal gas.
= (12.15)
From Eq. (12.34), the constant is P1V1 or P2V2
(12.16)
As expected, if work is done by the gas in an adiabatic process (W >
0), from Eq. (12.16), T2 < T1. On the other hand, if work is done on the
gas (W < 0), we get T2 > T1 i.e., the temperature of the gas rises.
Isochoric process
In an isochoric process, V is constant. No work is done on or by the
gas. From Eq. (12.1), the heat absorbed by the gas goes entirely to
change its internal energy and its temperature. The change in
temperature for a given amount of heat is determined by the specific
heat of the gas at constant volume.
Isobaric process
In an isobaric process, P is fixed. Work done by the gas is
W = P (V2 V1) = R (T2 T1) (12.17)
Since temperature changes, so does internal energy. The heat
absorbed goes partly to increase internal energy and partly to do
work. The change in temperature for a given amount of heat is
determined by the specific heat of the gas at constant pressure.
Cyclic process
In a cyclic process, the system returns to its initial state. Since internal
energy is a state variable, U = 0 for a cyclic process. From Eq.
(12.1), the total heat absorbed equals the work done by the system.
Fig. 12.9 Schematic representation of a heat engine. The engine takes heat Q1
from a hot reservoir at temperature T1, releases heat Q2 to a cold reservoir at
temperature T2 and delivers work W to the surroundings.
The cycle is repeated again and again to get useful work for some
purpose. The discipline of thermodynamics has its roots in the study of
heat engines. A basic question relates to the efficiency of a heat
engine. The efficiency () of a heat engine is defined by
(12.18)
where Q1 is the heat input i.e., the heat absorbed by the system in one
complete cycle and W is the work done on the environment in a cycle.
In a cycle, a certain amount of heat (Q2) may also be rejected to the
environment. Then, according to the First Law of Thermodynamics,
over one complete cycle,
W = Q1 Q2 (12.19)
i.e.,
(12.20)
For Q2 = 0, = 1, i.e., the engine will have 100% efficiency in
converting heat into work. Note that the First Law of Thermodynamics
i.e., the energy conservation law does not rule out such an engine. But
experience shows that such an ideal engine with = 1 is never
possible, even if we can eliminate various kinds of losses associated
with actual heat engines. It turns out that there is a fundamental limit
on the efficiency of a heat engine set by an independent principle of
nature, called the Second Law of Thermodynamics (section 12.11).
The mechanism of conversion of heat into work varies for different
heat engines. Basically, there are two ways : the system (say a gas or
a mixture of gases) is heated by an external furnace, as in a steam
engine; or it is heated internally by an exothermic chemical reaction as
in an internal combustion engine. The various steps involved in a
cycle also differ from one engine to another.
Pioneers of Thermodynamics
Lord Kelvin (William Thomson) (1824-1907), born in Belfast, Ireland,
is among the foremost British scientists of the nineteenth century.
Thomson played a key role in the development of the law of conservation of
energy suggested by the work of James Joule (1818-1889), Julius Mayer
(1814-1878) and Hermann Helmholtz (1821-1894). He collaborated with Joule
on the so-called Joule-Thomson effect : cooling of a gas when it expands into
vacuum. He introduced the notion of the absolute zero of temperature and
proposed the absolute temperature scale, now called the Kelvin scale in his
honour. From the work of Sadi
Carnot
(1796-1832), Thomson arrived at a form of the Second Law of
Thermodynamics. Thomson was a versatile physicist, with notable
contributions to electromagnetic theory and hydrodynamics.
where Q2 is the heat extracted from the cold reservoir and W is the
work done on the systemthe refrigerant. ( for heat pump is defined
as Q1/W) Note that while by definition can never exceed 1, can be
greater than 1. By energy conservation, the heat released to the hot
reservoir is
Q1 = W + Q2
i.e., (12.22)
In a heat engine, heat cannot be fully converted to work; likewise a
refrigerator cannot work without some external work done on the
system, i.e., the coefficient of performance in Eq. (12.21) cannot be
infinite.
CLAUSIUS STATEMENT
No process is possible whose sole result is the transfer of heat from a
colder object to a hotter object.
It can be proved that the two statements above are completely
equivalent.
Fig. 12.11 Carnot cycle for a heat engine with an ideal gas as the working
substance.
(b) Step 2 3 Adiabatic expansion of the gas from (P2, V2, T1) to (P3,
V3, T2) Work done by the gas, using Eq. (12.16), is
(12.24)
(c) Step 3 4 Isothermal compression of the gas from (P3, V3, T2) to
(P4, V4, T2).
(12.25)
(d) Step 4 1 Adiabatic compression of the gas from (P4, V4, T2) to
(P1,V1, T1).
Work done on the gas, [using Eq.(12.16)], is
(12.26)
From Eqs. (12.23) to (12.26) total work done by the gas in one
complete cycle is
W = W1 2 + W2 3 W3 4 W4 1
i.e. (12.29)
Similarly, since step 4 1 is an adiabatic process
i.e. (12.30)
From Eqs. (12.29) and (12.30),
(12.31)
Using Eq. (12.31) in Eq. (12.28), we get
Fig. 12.12 An irreversible engine (I) coupled to a reversible refrigerator (R). If W >
W, this would amount to extraction of heat W W from the sink and its full
conversion to work, in contradiction with the Second Law of Thermodynamics.
(12.33)
Q = U + W
where Q is the heat supplied to the system, W is the work done by the
system and U is the change in internal energy of the system.
where m is the mass of the substance and Q is the heat required to change
its temperature by T. The molar specific heat capacity of a substance is
defined by
where is the number of moles of the substance. For a solid, the law of
equipartition of energy gives
C=3R
Calorie is the old unit of heat. 1 calorie is the amount of heat required to raise
the temperature of 1 g of water from 14.5 C to 15.5 C. 1 cal = 4.186 J.
5. For an ideal gas, the molar specific heat capacities at constant pressure and
volume satisfy the relation
Cp Cv = R
Q = W = R T ln
9. In an adiabatic process of an ideal gas
PV = constant
where
Work done by an ideal gas in an adiabatic change of state from (P1, V1, T1) to
(P2, V2, T2) is
11. In a refrigerator or a heat pump, the system extracts heat Q2 from the cold
reservoir and releases Q1 amount of heat to the hot reservoir, with work W
done on the system. The co-efficient of performance of a refrigerator is given
by
Kelvin-Planck statement
Clausius statement
No process is possible whose sole result is the transfer of heat from a colder
object to a hotter object.
Put simply, the Second Law implies that no heat engine can have efficiency
equal to 1 or no refrigerator can have co-efficient of performance equal to
infinity.
13. A process is reversible if it can be reversed such that both the system and
the surroundings return to their original states, with no other change anywhere
else in the universe. Spontaneous processes of nature are irreversible. The
idealised reversible process is a quasi-static process with no dissipative
factors such as friction, viscosity, etc.
(Carnot engine)
4. Heat capacity, in general, depends on the process the system goes through
when heat is supplied.
12.1 A geyser heats water flowing at the rate of 3.0 litres per
minute from 27 C to 77 C. If the geyser operates on a gas
burner, what is the rate of consumption of the fuel if its heat of
combustion is 4.0 104 J/g ?
12.2 What amount of heat must be supplied to 2.0 102 kg of
nitrogen (at room temperature) to raise its temperature by 45 C at
constant pressure ? (Molecular mass of N2 = 28; R = 8.3 J mol1
K1.)
12.3 Explain why
(a) Two bodies at different temperatures T1 and T2 if brought in
thermal contact do not necessarily settle to the mean temperature
(T1 + T2 )/2.
(b) The coolant in a chemical or a nuclear plant (i.e., the liquid
used to prevent the different parts of a plant from getting too hot)
should have high specific heat.
(c) Air pressure in a car tyre increases during driving.
Kinetic Theory
13.1 Introduction
Summary
Points to ponder
Exercises
ADDITIONAL EXERCISES
13.1 INTRODUCTION
Boyle discovered the law named after him in 1661. Boyle, Newton and
several others tried to explain the behaviour of gases by considering
that gases are made up of tiny atomic particles. The actual atomic
theory got established more than 150 years later. Kinetic theory
explains the behaviour of gases based on the idea that the gas
consists of rapidly moving atoms or molecules. This is possible as the
inter-atomic forces, which are short range forces that are important for
solids and liquids, can be neglected for gases. The kinetic theory was
developed in the nineteenth century by Maxwell, Boltzmann and
others. It has been remarkably successful. It gives a molecular
interpretation of pressure and temperature of a gas, and is consistent
with gas laws and Avogadros hypothesis. It correctly explains specific
heat capacities of many gases. It also relates measurable properties
of gases such as viscosity, conduction and diffusion with molecular
parameters, yielding estimates of molecular sizes and masses. This
chapter gives an introduction to kinetic theory.
In ancient Greece, Democritus (Fourth century B.C.) is best known for his
atomic hypothesis. The word atom means indivisible in Greek. According to
him, atoms differ from each other physically, in shape, size and other
properties and this resulted in the different properties of the substances
formed by their combination. The atoms of water were smooth and round and
unable to hook on to each other, which is why liquid /water flows easily. The
atoms of earth were rough and jagged, so they held together to form hard
substances. The atoms of fire were thorny which is why it caused painful
burns. These fascinating ideas, despite their ingenuity, could not evolve much
further, perhaps because they were intuitive conjectures and speculations not
tested and modified by quantitative experiments - the hallmark of modern
science.
To explain the laws Dalton suggested, about 200 years ago, that the
smallest constituents of an element are atoms. Atoms of one element
are identical but differ from those of other elements. A small number of
atoms of each element combine to form a molecule of the compound.
Gay Lussacs law, also given in early 19th century, states: When
gases combine chemically to yield another gas, their volumes are in
the ratios of small integers. Avogadros law (or hypothesis) says:
Equal volumes of all gases at equal temperature and pressure have
the same number of molecules. Avogadros law, when combined with
Daltons theory explains Gay Lussacs law. Since the elements are
often in the form of molecules, Daltons atomic theory can also be
referred to as the molecular theory of matter. The theory is now well
accepted by scientists. However even at the end of the nineteenth
century there were famous scientists who did not believe in atomic
theory !
From many observations, in recent times we now know that molecules
(made up of one or more atoms) constitute matter. Electron
microscopes and scanning tunnelling microscopes enable us to even
-10
see them. The size of an atom is about an angstrom (10 m). In
solids, which are tightly packed, atoms are spaced about a few
angstroms (2 ) apart. In liquids the separation between atoms is also
about the same. In liquids the atoms are not as rigidly fixed as in
solids, and can move around. This enables a liquid to flow. In gases
the interatomic distances are in tens of angstroms. The average
distance a molecule can travel without colliding is called the mean free
path. The mean free path, in gases, is of the order of thousands of
angstroms. The atoms are much freer in gases and can travel long
distances without colliding. If they are not enclosed, gases disperse
away. In solids and liquids the closeness makes the interatomic force
important. The force has a long range attraction and a short range
repulsion. The atoms attract when they are at a few angstroms but
repel when they come closer. The static appearance of a gas is
misleading. The gas is full of activity and the equilibrium is a dynamic
one. In dynamic equilibrium, molecules collide and change their
speeds during the collision. Only the average properties are constant.
Atomic theory is not the end of our quest, but the beginning. We now
know that atoms are not indivisible or elementary. They consist of a
nucleus and electrons. The nucleus itself is made up of protons and
neutrons. The protons and neutrons are again made up of quarks.
Even quarks may not be the end of the story. There may be string like
elementary entities. Nature always has surprises for us, but the search
for truth is often enjoyable and the discoveries beautiful. In this
chapter, we shall limit ourselves to understanding the behaviour of
gases (and a little bit of solids), as a collection of moving molecules in
incessant motion.
13.3 Behaviour of Gases
Properties of gases are easier to understand than those of solids and
liquids. This is mainly because in a gas, molecules are far from each
other and their mutual interactions are negligible except when two
molecules collide. Gases at low pressures and high temperatures
much above that at which they liquefy (or solidify) approximately
satisfy a simple relation between their pressure, temperature and
volume given by (see Ch. 11)
PV = KT (13.1)
for a given sample of the gas. Here T is the temperature in kelvin or
(absolute) scale. K is a constant for the given sample but varies with
the volume of the gas. If we now bring in the idea of atoms or
molecules then K is proportional to the number of molecules, (say) N
in the sample. We can write K = N k . Observation tells us that this k is
same for all gases. It is called Boltzmann constant and is denoted by
kB.
As = constant = kB (13.2)
if P, V and T are same, then N is also same for all gases. This is
Avogadros hypothesis, that the number of molecules per unit volume
is same for all gases at a fixed temperature and pressure. The number
in 22.4 litres of any gas is 6.02 10 23 . This is known as Avogadro
number and is denoted by N A . The mass of 22.4 litres of any gas is
equal to its molecular weight in grams at S.T.P (standard temperature
273 K and pressure 1 atm). This amount of substance is called a mole
(see Chapter 2 for a more precise definition). Avogadro had guessed
the equality of numbers in equal volumes of gas at a fixed temperature
and pressure from chemical reactions. Kinetic theory justifies this
hypothesis.
The perfect gas equation can be written as
PV = RT (13.3)
where is the number of moles and R = N AkB is a universal constant.
The temperature T is absolute temperature. Choosing kelvin scale for
absolute temperature, R = 8.314 J mol1 K1 .
Here
(13.4)
He was an English chemist. When different types of atoms combine, they obey
certain simple laws. Daltons atomic theory explains these laws in a simple
way. He also gave a theory of colour blindness.
He made a brilliant guess that equal volumes of gases have equal number of
molecules at the same temperature and pressure. This helped in
understanding the combination of different gases in a very simple way. It is
now called Avogadros hypothesis (or law). He also suggested that the
smallest constituent of gases like hydrogen, oxygen and nitrogen are not
atoms but diatomic molecules.
P (atm)
Fig.13.1 Real gases approach ideal gas behaviour at low pressures and high
temperatures.
where n is the number density, i.e. number of molecules per unit
volume. kB is the Boltzmann constant introduced above. Its value in SI
units is 1.38 1023JK1.
Another useful form of Eq. (13.3) is
(13.5)
where is the mass density of the gas.
PV = constant (13.6)
i.e., keeping temperature constant, pressure of a given mass of gas
varies inversely with volume. This is the famous Boyles law. Fig. 13.2
shows comparison between experimental P-V curves and the
theoretical curves predicted by Boyles law. Once again you see that
the agreement is good at high temperatures and low pressures. Next,
if you fix P, Eq. (13.1) shows that V Ti.e., for a fixed pressure, the
volume of a gas is proportional to its absolute temperature T (Charles
law). See Fig. 13.3.
Fig.13.2 Experimental P-V curves (solid lines) for steam at three temperatures
compared with Boyles law (dotted lines). P is in units of 22 atm and V in units of
0.09 litres.
i.e. (13.8)
= P1 + P2 + ... (13.9)
Clearly P1 = 1RT/V is the pressure gas 1 would exert at the same
conditions of volume and temperature if no other gases were present.
This is called the partial pressure of the gas. Thus, the total pressure
of a mixture of ideal gases is the sum of partial pressures. This is
Daltons law of partial pressures.
Fig. 13.3 Experimental T-V curves (solid lines) for CO2 at three pressures
compared with Charles law (dotted lines). T is in units of 300 K and V in units of
0.13 litres.
Example 13.1 The density of water is 1000 kg m3. The density of water
vapour at 100 C and 1 atm pressure is 0.6 kg m3. The volume of a molecule
multiplied by the total number gives ,what is called, molecular volume.
Estimate the ratio (or fraction) of the molecular volume to the total volume
occupied by the water vapour under the above conditions of temperature and
pressure.
Example 13.2 Estimate the volume of a water molecule using the data in
Example 13.1.
Answer In the liquid (or solid) phase, the molecules of water are quite
closely packed. The density of water molecule may therefore, be
regarded as roughly equal to the density of bulk water = 1000 kg m3.
To estimate the volume of a water molecule, we need to know the
mass of a single water molecule. We know that 1 mole of water has a
mass approximately equal to (2 + 16)g = 18 g = 0.018 kg. Since 1
mole contains about 6 1023 molecules (Avogadros number), the
mass of a molecule of water is (0.018)/(6 1023) kg = 3 1026 kg.
Therefore, a rough estimate of the volume of a water molecule is as
follows : Volume of a water molecule = (3 1026 kg)/ (1000 kg m3) =
3 1029 m3 = (4/3) (Radius)3 Hence, Radius 2 1010 m = 2
Example 13.3 What is the average distance between atoms (interatomic
distance) in water? Use the data given in Examples 13.1 and 13.2.
Fig. 13.4 Elastic collision of a gas molecule with the wall of the container.
of area A (= l2). Since the collision is elastic, the molecule rebounds
with the same velocity; its y and z components of velocity do not
change in the collision but the x-component reverses sign. That is, the
velocity after collision is (-vx , vy , vz) . The change in momentum of
the molecule is : mvx (mvx ) = 2mvx . By the principle of
conservation of momentum, the momentum imparted to the wall in the
collision = 2mvx .
To calculate the force (and pressure) on the wall, we need to calculate
momentum imparted to the wall per unit time. In a small time interval
t, a molecule with x-component of velocity vx will hit the wall if it is
within the distance vx t from the wall. That is, all molecules within the
volume Avx t only can hit the wall in time t. But, on the average, half
of these are moving towards the wall and the other half away from the
wall. Thus the number of molecules with velocity (vx , vy , vz) hitting
the wall in time t is 12A vx t n where n is the number of molecules
per unit volume. The total momentum transferred to the wall by these
molecules in time t is :
Q = (2mvx ) (12 n A vx t ) (13.10)
The force on the wall is the rate of momentum transfer Q/t and
pressure is force per unit area :
P = Q /(A t) = n m vx2 (3.11)
Actually, all molecules in a gas do not have the same velocity; there is
a distribution in velocities. The above equation therefore, stands for
pressure due to the group of molecules with speed vx in the x-direction
and n stands for the number density of that group of molecules. The
total pressure is obtained by summing over the contribution due to all
groups:
P=nm (13.12)
where is the average of vx2 . Now the gas is isotropic, i.e. there is
no preferred direction of velocity of the molecules in the vessel.
Therefore, by symmetry,
= =
where v is the speed and denotes the mean of the squared speed.
Thus
P = (1/3) n m (13.14)
Some remarks on this derivation. First, though we choose the
container to be a cube, the shape of the vessel really is immaterial.
For a vessel of arbitrary shape, we can always choose a small
infinitesimal (planar) area and carry through the steps above. Notice
that both A and t do not appear in the final result. By Pascals law,
given in Ch. 10, pressure in one portion of the gas in equilibrium is the
same as anywhere else. Second, we have ignored any collisions in
the derivation. Though this assumption is difficult to justify rigorously,
we can qualitatively see that it will not lead to erroneous results. The
number of molecules hitting the wall in time t was found to be 12 n
Avx t. Now the collisions are random and the gas is in a steady state.
Thus, if a molecule with velocity (vx , vy , vz ) acquires a different
velocity due to collision with some molecule, there will always be
some other molecule with a different initial velocity which after a
collision acquires the velocity (vx , vy , vz). If this were not so, the
distribution of velocities would not remain steady. In any case we are
finding . Thus, on the whole, molecular collisions (if they are not
too frequent and the time spent in a collision is negligible compared to
time between collisions) will not affect the calculation above.
PV = (1/3) nV m (13.15a)
PV = (2/3) N x 12 m (13.15b)
where N (= nV ) is the number of molecules in the sample.
E = N (1/2) m (13.16)
Equation (13.15) then gives :
PV = (2/3) E (13.17)
We are now ready for a kinetic interpretation of temperature.
Combining Eq. (13.17) with the ideal gas Eq. (13.3), we get
E = (3/2) kBNT (13.18)
12 m1 = 12 m2 = (3/2) kBT
so that
P = (n1 + n2 +... ) kBT (13.21)
which is Daltons law of partial pressures.
From Eq. (13.19), we can get an idea of the typical speed of
molecules in a gas. At a temperature T = 300 K, the mean square
speed of a molecule in nitrogen gas is :
kg.
= 3 kB T / m = (516)2 m2s-2
The square root of is known as root mean square (rms) speed and
is denoted by vrms,
Example 13.5 A flask contains argon and chlorine in the ratio of 2:1 by mass.
The temperature of the mixture is 27 C. Obtain the ratio of (i) average kinetic
energy per molecule, and (ii) root mean square speed vrms of the molecules
of the two gases. Atomic mass of argon = 39.9 u; Molecular mass of chlorine =
70.9 u.
= =1.77
where M denotes the molecular mass of the gas. (For argon, a
molecule is just an atom of argon.)
Taking square root of both sides,
= 1.33
You should note that the composition of the mixture by mass is quite
irrelevant to the above calculation. Any other proportion by mass of
argon and chlorine would give the same answers to (i) and (ii),
provided the temperature remains unaltered.
*E denotes the translational part of the internal energy U that may include energies
due to other degrees of freedom also. See section 13.5.
In a given mass of gas, the velocities of all molecules are not the same, even
when bulk parameters like pressure, volume and temperature are fixed.
Collisions change the direction and the speed of molecules. However in a
state of equilibrium, the distribution of speeds is constant or fixed.
Distributions are very important and useful when dealing with systems
containing large number of objects. As an example consider the ages of
different persons in a city. It is not feasible to deal with the age of each
individual. We can divide the people into groups: children up to age 20 years,
adults between ages of 20 and 60, old people above 60. If we want more
detailed information we can choose smaller intervals, 0-1, 1-2,..., 99-100 of
age groups. When the size of the interval becomes smaller, say half year, the
number of persons in the interval will also reduce, roughly half the original
number in the one year interval. The number of persons dN(x) in the age
interval x and x+dx is proportional to dx or dN(x) = nx dx. We have used nx to
denote the number of persons at the value of x.
Maxwell distribution of molecular speeds
Example 13.6 Uranium has two isotopes of masses 235 and 238 units. If both
are present in Uranium hexafluoride gas which would have the larger average
speed ? If atomic mass of fluorine is 19 units, estimate the percentage
difference in speeds at any temperature.
2
Answer At a fixed temperature the average energy = m <v > is
constant. So smaller the mass of the molecule, faster will be the
speed. The ratio of speeds is inversely proportional to the square root
of the ratio of the masses. The masses are 349 and 352 units. So
v349 / v352 = ( 352/ 349)1/2 = 1.0044 .
Example 13.7 (a) When a molecule (or an elastic ball) hits a ( massive) wall, it
rebounds with the same speed. When a ball hits a massive bat held firmly, the
same thing happens. However, when the bat is moving towards the ball, the
ball rebounds with a different speed. Does the ball move faster or slower?
(Ch.6 will refresh your memory on elastic collisions.)
(b) When gas in a cylinder is compressed by pushing in a piston, its
temperature rises. Guess at an explanation of this in terms of kinetic theory
using (a) above.
(c) What happens when a compressed gas pushes a piston out and expands.
What would you observe ?
(d) Sachin Tendulkar uses a heavy cricket bat while playing. Does it help him
in anyway ?
Answer (a) Let the speed of the ball be u relative to the wicket behind
the bat. If the bat is moving towards the ball with a speed V relative to
the wicket, then the relative speed of the ball to bat is V + u towards
the bat. When the ball rebounds (after hitting the massive bat) its
speed, relative to bat, is V + u moving away from the bat. So relative
to the wicket the speed of the rebounding ball is V + (V + u) = 2V + u,
moving away from the wicket. So the ball speeds up after the collision
with the bat. The rebound speed will be less than u if the bat is not
massive. For a molecule this would imply an increase in temperature.
You should be able to answer (b) (c) and (d) based on the answer to
(a).
molecule ball.) t
(13.22)
For a gas in thermal equilibrium at temperature T the average value of
(13.23)
Since there is no preferred direction, Eq. (13.23) implies
, ,
(13.24)
(13.25)
Fig. 13.6 The two independent axes of rotation of a diatomic molecule
where 1 and 2 are the angular speeds about the axes 1 and 2 and
I1, I2 are the corresponding moments of inertia. Note that each
rotational degree of freedom contributes a term to the energy that
contains square of a rotational variable of motion.
We have assumed above that the O2 molecule is a rigid rotator, i.e.
the molecule does not vibrate. This assumption, though found to be
true (at moderate temperatures) for O2, is not always valid. Molecules
like CO even at moderate temperatures have a mode of vibration, i.e.
its atoms oscillate along the interatomic axis like a one-dimensional
oscillator, and contribute a vibrational energy term v to the total
energy:
(13.26)
where k is the force constant of the oscillator and y the vibrational co-
ordinate.
Once again the vibrational energy terms in Eq. (13.26) contain
squared terms of vibrational variables of motion y and dy/dt .
(13.27)
The molar specific heat at constant volume, Cv, is
* Rotation along the line joining the atoms has very small moment of
inertia and does not come into play for quantum mechanical reasons.
See end of section 13.6.
Cp = R (13.30)
R (13.35)
i.e. Cv = (3 + f ) R, Cp = (4 + f ) R, (13.36)
Note that Cp Cv = R is true for any ideal gas, whether mono, di or
polyatomic.
Table 13.1 summarises the theoretical predictions for specific heats of
gases ignoring any vibrational modes of motion. The values are in
good agreement with experimental values of specific heats of several
gases given in Table 13.2. Of course, there are discrepancies
between predicted and actual values of specific heats of several other
gases (not shown in the table), such as Cl2, C2H6 and many other
polyatomic gases. Usually, the experimental values for specific heats
of these gases are greater than the predicted values given in
Table13.1 suggesting that the agreement can be improved by
including vibrational modes of motion in the calculation. The law of
equipartition of energy is thus well verified experimentally at ordinary
temperatures.
Example 13.8 A cylinder of fixed capacity 44.8 litres contains helium gas at
standard temperature and pressure. What is the amount of heat needed to
raise the temperature of the gas in the cylinder by 15.0 C ? (R = 8.31 J mo11
K1).
Answer Using the gas law PV = RT, you can easily show that 1 mol
of any (ideal) gas at standard temperature (273 K) and pressure
Table 13.1 Predicted values of specific heat capacities of gases
(ignoring vibrational modes),
U = 3 kBT NA = 3 RT
Now at constant pressure Q = U + PV
= U, since for a solid V is negligible. Hence,
(13.37)
Can one see atoms rushing about. Almost but not quite. One can see pollen
grains of a flower being pushed around by molecules of water. The size of the
grain is ~ 10-5 m. In 1827, a Scottish botanist Robert Brown, while examining,
under a microscope, pollen grains of a flower suspended in water noticed that
they continuously moved about in a zigzag, random fashion.
When the object is sufficiently small but still visible under a microscope, the
difference in molecular hits from different directions is not altogether negligible,
i.e. the impulses and the torques given to the suspended object through
continuous bombardment by the molecules of the medium (water or some
other fluid) do not exactly sum to zero. There is a net impulse and torque in
this or that direction. The suspended object thus, moves about in a zigzag
manner and tumbles about randomly. This motion called now Brownian
motion is a visible proof of molecular activity. In the last 50 years or so
molecules have been seen by scanning tunneling and other special
microscopes.
Fig. 13.7 The volume swept by a molecule in time t in which any molecule will
collide with it.
(13.40)
Let us estimate l and for air molecules with average speeds <v> = (
485m/s). At STP
n=
= 2.7 10 25 m 3.
Taking, d = 2 1010 m,
= 6.1 1010 s
and l = 2.9 107 m 1500d (13.41)
Example 13.9 Estimate the mean free path for a water molecule in water
vapour at 373 K. Use information from Exercises 13.1 and Eq. (13.41) above.
Answer The d for water vapour is same as that of air. The number
density is inversely proportional to absolute temperature.
So
SUMMARY
1. The ideal gas equation connecting pressure (P), volume (V) and absolute
temperature (T ) is
PV = RT = kB NT
Real gases satisfy the ideal gas equation only approximately, more so at low
pressures and high temperatures.
is the mean of squared speed. Combined with the ideal gas equation it
yields a kinetic interpretation of temperature.
This tells us that the temperature of a gas is a measure of the average kinetic
energy of a molecule, independent of the nature of the gas or molecule. In a
mixture of gases at a fixed temperature the heavier molecule has the lower
average speed.
E= kB NT.
PV = E
2 kB T = kB T.
5. Using the law of equipartition of energy, the molar specific heats of gases
can be determined and the values are in agreement with the experimental
values of specific heats of several gases. The agreement can be improved by
including vibrational modes of motion.
POINTS TO PONDER
1. Pressure of a fluid is not only exerted on the wall. Pressure exists
everywhere in a fluid. Any layer of gas inside the volume of a container is in
equilibrium because the pressure is the same on both sides of the layer.
3. The law of equipartition of energy is stated thus: the energy for each degree
of freedom in thermal equilibrium is kB T. Each quadratic term in the total
energy expression of a molecule is to be counted as a degree of freedom.
Thus, each vibrational mode gives 2 (not 1) degrees of freedom (kinetic and
potential energy modes), corresponding to the energy 2 kB T = kB T.
4. Molecules of air in a room do not all fall and settle on the ground (due to
gravity) because of their high speeds and incessant collisions. In equilibrium,
there is a very slight increase in density at lower heights (like in the
atmosphere). The effect is small since the potential energy (mgh) for ordinary
heights is much less than the average kinetic energy mv2 of the molecules.
5. < v2 > is not always equal to ( < v >)2. The average of a squared quantity is
not necessarily the square of the average. Can you find examples for this
statement.
EXERCISES
13.1 Estimate the fraction of molecular volume to the actual
volume occupied by oxygen gas at STP. Take the diameter of an
oxygen molecule to be 3 .
13.2 Molar volume is the volume occupied by 1 mol of any (ideal)
gas at standard temperature and pressure (STP : 1 atmospheric
pressure, 0 C). Show that it is 22.4 litres.
13.3 Figure 13.8 shows plot of PV/T versus P for 1.00103 kg of
oxygen gas at two different temperatures.
Fig. 13.8
13.14 Given below are densities of some solids and liquids. Give
rough estimates of the size of their atoms :