Download as pdf or txt
Download as pdf or txt
You are on page 1of 151

Dept.

of EEE
Ch1103 Chemistry Credit: 4
Contact Hours: 4Hrs/ Week

Chemical kinetics: Theories of reaction rates, Monitoring


the progress of a reaction, rate laws and rate constants;
Experimental determination of rate law, order and
molecularity of a chemical reaction, rate laws for reaction
approaching equilibrium, temperature dependence of
reaction of reaction rates; Elementary reaction, steady state
approximation; Analyzing pre-equilibria, unimolecular
reactions.

Chemical Equilibrium: Law of mass action and its


application; Magnitude of equilibrium constant and the
direction of reaction, Chemical equilibrium of
homogeneous and heterogeneous reactions, factors
affecting equilibrium constant, Equilibrium constant,
Effect of pressure on chemical equilibrium; Le-Chateller’s
theorem and application; Solvent and ion exchange
processes, Numerical values of equilibrium constant,
thermodynamic treatment of equilibrium constant, Ionic
equilibrium.
Photochemistry: Definition, Photon, law of photo
chemistry, absorption law and mechanism of
photochemical reaction, fluorescence, phosphorescence
and chemiluminescence, Beer-Lambert law, Quantum
yield, Photosensitized reaction.
Chemistry of polymer: Polymer and polymerization, co-
polymerization, ionic polymerization, living polymer,
structure and properties of macromolecules, plastic and
rubber, conducting polymer.
Crystallography: Crystal symmetry, Miller indices,
different methods for the determination of structure;
Structures of the metallic elements and certain compound
with 3-dimensinal lattices; Defects in solid states.
Periodic Table: Electronic structure of the elements,
Generalization of chemical properties from periodic table.
Inert gases and there importance in industry,
Semiconductors, metallic bond, band theory, hydrogen
bonding, chelate bond.
Electrochemistry: Electrolytes; Nernst’s theory of
electrode potential, type of electrodes and electrode
potentials, EMF measurement, polarization and over
potentials; Origin of EMF, Free energy and EMF,
Electrical double layer, Factor affecting electrode reaction
and current, Modes of Mass transfer, Lithium ion battery.
Transport number; pH value and its determination;
Electrode potentials and corrosion, Electroplating and
galvanizing.
Nuclear chemistry, Nuclear reaction, Nuclear binding
energy, Radioactivity, patterns of nuclear stability;
Nuclear transmutations, Nuclear model, energy changes in
nuclear reactions; Nuclear Fission, Nuclear Fusion;
Nuclear reactor, Nuclear force, Methods of separation of
isotopes, applications of isotopes, nuclear hazard.

Reference Book
1. Essential of physical chemistry by Bahl & Tuli
2. Radiochemistry & nuclear method of analysis by W. D.
Ehmann & D. E. Vance
3. Principles of physical chemistry by Md. Yousuf Ali
Mollah
4. Concise inorganic chemistry by J D Lee

Nuclear chemistry
Nuclear Reactions: Nuclear reactions are interactions
between an incoming fundamental particle, proton,
neutrino, or multinucleon nucleus, and a target nucleus.
The reaction may result in the scattering of the projectile,
the excitation of the target nucleus, or excitation followed
by nuclear transformations of the target to another nuclide
by gain or loss of subatomic particles.

A nuclear reaction is generally represented by an equation


indicating the nuclear characteristics of the reactants and
products as

Where X stands for the target nucleus, a for the projectile


effecting the reaction, b for the particle ejected and y for
the product nucleus. Q is the energy balance of the
reaction. Except when the nuclear charge conservation is
to be emphasized (Z1 + Z2 = Z3+ Z4), the Zs are omitted in
the equation. This reaction can also be written as

when the particles a and b are uniquely distinguishable.

A shorter form of writing a nuclear reaction is due to


Bethe, on which notation the above reaction can be
represented as A1X(a, b)A4Y
For example,

Q-value of Nuclear Reactions and Its Significance


The Q value of a nuclear reaction refers to the
amount of energy released or absorbed in the
reaction. This is analogous to the thermodynamic
quantity of the enthalpy of a chemical reaction, H.

The Q value for a nuclear reaction can be calculated


by
Q = (Massesreactants - Massesproducts) × 931.5 MeV

A positive value of Q means that the masses of the


products are less than those of the reactants, energy
has been released, and the reaction is exoergic.
Conversely, a negative Q value indicates energy has
been absorbed an endoergic reaction.

[Note: The sign convention of Q being reverse for


enthalpy, H in thermodynamics.]
Differences between nuclear reactions and
chemical reactions

Nuclear Reactions Chemical Reactions


1. Proceed by redistribution 1. Proceed by the
of nuclear particles. rearrangement of
extranuclear electrons.
2. One element may be 2. No new element can be
converted into another. produced.
3. Often accompanied by 3. Accompanied by release
release or absorption of or absorption of relatively
enormous amount of small amount of energy.
energy.

4. Rate of reaction is 4. Rate of reaction is


unaffected by external influenced by external
factors such as factors.
concentration, temperature,
pressure and catalyst.
5. Nuclear reaction is 5. There is breaking of old
independent of such
bonds and formation of
factors. new chemical bonds in a
chemical reaction.
6. Nuclear reactions are 6. Chemical reactions can
mostly irreversible either be reversible or
irreversible.

Nuclear Fission
When the nucleus of an atom splits into lighter nuclei
through a nuclear reaction the process is termed as nuclear
fission. This decay can be natural spontaneous splitting by
radioactive decay, or can actually be simulated in a lab by
achieving necessary conditions (bombarding with
neutrinos.
Examples of Nuclear Fission
1. An example of nuclear fission is the splitting of
Uranium-235. The equation of the reaction has
been given below:
2. The other example of nuclear fission is the
splitting of Uranium-233. The equation of the
reaction has been given below:

3. The splitting of Plutonium-239 is the other


example of nuclear fission given below:

Nuclear Fusion
Nuclear fusion is a reaction through which two or
more light nuclei collide to form a heavier nucleus.
The nuclear fusion process occurs in elements that
have a low atomic number, such as hydrogen.
Example: The energy released by the sun results from a
series of nuclear fusion reactions.
Nuclear fission Nuclear Fusion
1. A bigger (heavier nucleus 1. Lighter nuclei fuse
splits into smaller (lighter) together to form the heavier
nuclei. nucleus.
2. It does not require high 2. Extremely high
temperature. temperature is required for
fusion to take place.
3. A chain reaction sets in. 3. It is not a chain reaction.
4. It can be controlled and 4. It cannot be controlled
energy released can be used and energy released cannot
for peaceful purposes. be used properly.
5. The products of the 5. The products of a fusion
reaction are radioactive in reaction are nonradioactive
nature. in nature.
6. At the end of the reaction 6. No nuclear waste is left at
nuclear waste is left behind. the end of fusion reaction.
MCQ: 38, 53.
5. How many α and β particles will be emitted by an
element 218 206
84𝐴 in changing to a stable isotope of 82𝐵 ?
Mention the symbol of A and B. (Tuli-139)
15(b) The mass number and atomic number of a
radioactive element Actinium are 227 and 89 respectively.
Calculate the number of α and β particles emitted, if the
mass number and atomic number of the new element lead
are 207 and 82 respectively.
Answer: 5 α and 3 β
44. 238
92𝑈 undergoes a series of changes emitting α and β
particles and finally 206 82𝑃𝑏 is formed. Calculate the
number of α and β particles which must have been ejected
during the series. (Tuli-142)
Answer: 8 α and 6 β
Practice problem: 19, 24, 35(a)
MCQ: 48, 51
Atomic Nucleus: The positively charged core of an
atom. The nucleus contains protons and neutrons
and accounts for most of the mass of the atom.

The present day particle physicist no longer


considers protons and neutrons as elementary but as
composite particles. According to Dirac,
1 1
1𝑝→ 0𝑛 + 0
+1𝑒 +
1
0𝑛 → 1
1𝑝 + 0
−1𝑒 +
̅

where, p = proton, e+ = positron (an anti- electron), 


= neutrino, n = neutron, e- = electron, and

̅=antineutrino.
Neutrino: A particle of zero charge and a mass close
to zero.
Antiparticle: The antiparticle has the same mass and
spin number as the particle but differs in electric
charge, color or flavor.

Elementary particles are those with no known internal


structure.

Nuclear Mass: The nuclear mass is slightly less than


the sum of the masses of protons and neutrons. This
little difference in mass is spent in binding the
nucleons in the nucleus.

Mass Defect: The mass defect or mass excess is


equal to the difference between the nuclidic mass
and the mass number. This quantity is symbolized by
, i.e.,  = M - A where, M = nuclidic mass, and A =
mass number.

Packing Fraction: It is defined as (M - A)/A, i.e., the


mass excess per nucleon. Its value is sometimes
positive or zero but more often negative. A large
positive packing fraction value means that the
nucleons are loosely bound.

Binding Energy: The binding energy of a nucleus is


the energy needed to break a nucleus into its
individual protons and neutrons. It is denoted BE and
thus the binding energy of the nucleus is

BE = m x 931.5 MeV (MeV = Million electron volts)

where, m = Zmp, + Nmn, - m


Z = no. of protons, mp = mass of proton, N = no. of
neutron, mn = mass of neutron, and m = atomic mass.

The greater the value of m, the greater is the binding


energy of the nucleus. The binding energy of a
nucleus divided by the total number of nucleons (i.e.
neutrons + protons) is called the binding energy per
nucleon in the atom, i.e.,
Binding energy
Binding energy per nucleon =
Nucleons number

Nuclear Binding Energy


The energy required to break down a nucleus into its component nucleons is called
the nuclear binding energy.
63
Cu + Energy 29 p+ + 34 no
Nuclear binding energies are usually expressed in terms of kJ/mole of nuclei or
MeV's/nucleon. Calculation of the nuclear binding energy involves the following three
steps:

 Determining the Mass Defect


 Conversion of Mass Defect into Energy
 Expressing Nuclear Binding Energy as Energy per Mole of Atoms, or
as Energy per Nucleon

Determining the Mass Defect

The difference between the mass of a nucleus and the sum of the masses of the
nucleons of which it is composed is called the mass defect. Three things need to be
known in order to calculate the mass defect:

 the actual mass of the nucleus,


 the composition of the nucleus (number of protons and of neutrons),
 the masses of a proton and of a neutron.

To calculate the mass defect:

 add up the masses of each proton and of each neutron that make up the
nucleus,
 subtract the actual mass of the nucleus from the combined mass of the
components to obtain the mass defect.

Example: Find the mass defect of a copper-63 nucleus if the actual mass of a
copper-63 nucleus is 62.91367 amu.

 Find the composition of the copper-63 nucleus and determine the combined
mass of its components.

Copper has 29 protons and copper-63 also has (63 - 29) 34 neutrons.
The mass of a proton is 1.00728 amu and a neutron is 1.00867 amu.
The combined mass is calculated:

29 protons(1.00728 amu/proton) + 34 neutrons(1.00867 amu/neutron)


or
63.50590 amu

 Calculate the mass defect.

m = 63.50590 amu - 62.91367 amu = 0.59223 amu

Top
Conversion of Mass Defect into Energy

To convert the mass defect into energy:

 Convert the mass defect into kilograms (1 amu = 1.6606 x 10-27 kg)
 Convert the mass defect into its energy equivalent using Einstein's equation.

Example: Determine the binding energy of the copper-63 atom.

 Convert the mass defect (calculated in the previous example) into kg.

(0.59223 amu/nucleus)(1.6606 x 10-27 kg/amu) = 9.8346 x 10-28 kg/nucleus

 Convert this mass into energy using E = mc2, where c = 2.9979 x 108 m/s.

E = (9.8346 x 10-28 kg/nucleus)(2.9979 x 108 m/s)2 = 8.8387 x 10-11 J/nucleus

Top

Expressing Nuclear Binding Energy as Energy per Mole of Atoms, or as


Energy per Nucleon

The energy calculated in the previous example is the nuclear binding


energy. However, nuclear binding energy is often expressed as kJ/mol of nuclei or
as MeV/nucleon.

 To convert the energy to kJ/mol of nuclei we will simply employ the


conversion factors for converting joules into kilojoules (1 kJ = 1000 J) and for
converting individual particles into moles of particles (Avogadro's Number).

(8.8387 x 10-11 J/nucleus)(1 kJ/1000 J)(6.022 x 1023 nuclei/mol) = 5.3227 x


1010 kJ/mol of nuclei

 To convert the binding energy to MeV (megaelectron volts) per


nucleon we will employ the conversion factor for converting joules into MeV (1
MeV = 1.602 x 10-13 J) and the number of nucleons (protons and neutrons)
which make up the nucleus.

(8.8387 x 10-11 J/nucleus)[1 MeV/(1.602 x 10-13 J)](1 nucleus/63 nucleons) = 8.758


MeV/nucleon

Top
PP: 13(a), 36(a), 38(a)
MCQ: 54, 55, 56
Nuclear Stability: It is found that the nuclei of some
isotopes are stable for an indefinite period of time
whereas others can decay in one or more of several
different ways, with half lives that range from small
fractions of a second to billion of years. A study of the
characteristics of a large number of stable and
radioactive nuclides shows the following to be some
of the important factors involved.
(i) The Magic Number of Protons or Neutrons
The magic number in nuclear physics is defined as a
number of nucleons consisting of either neutrons or
protons separately arranged in a manner of complete
shells within their atomic nucleus. Atomic nuclei
consisting of a magic number of neutrons and
protons are considered much more stable than any
other nuclei.
Proton: 2, 8, 20, 28, 50, 82, 114
Neutron: 2, 8, 20, 28, 50, 82, 126, 184
The protons usually correspond with the following
elements called oxygen, helium, nickel, calcium, tin,
and lead, along with the hypothetical unbihexium, but
126 is so far to be considered as a magic number in
terms of neutrons. The atomic nuclei consisting of a
magical number in the nucleons are generally
occupied with average binding energy, which is very
high per nucleon than the one expected based upon
the predictions like the semi-empirical mass formula
and thus is considered to have much more stability
against its nuclear decay. The isotopes with magic
numbers have exceptional stability, which means that
the transuranium elements can be theoretically
created with extremely larger-sized nuclei and still
will not be subjected to any extremely rapid
radioactive decays, which are typically associated
with high atomic numbers. Thus it is assumed that
the significant isotopes consisting of magic numbers
exist in an island of stability.

(ii) The Binding Energy


The binding energy is another important
characteristic of each nuclide, whether stable or not,
by virtue of which it holds the nucleons together in the
nucleus. A plot of binding energy per nucleon vs.
mass number reaches a maximum at 56Fe. This
implies that the most stable nuclei lie in mass number
reaches a maximum at 56Fe. This implies that the
most stable nuclei lie in the vicinity of 56Fe. Nuclei
with low or high mass number become more stable
by acquiring mass number of about 56.
Thus heavy nuclei are expected to undergo fission
nuclei. Thus 235U undergoes fission into lighter and
more stable isotopes as 141Ba and 92Kr with the
release of energy. On the other hand, the lighter
nuclei are anticipated to perform fusion. Similarly two
or more lighter nuclei (2H, 3H) with lower binding
energy per nucleon combine or fuse together into a
heavier and more stable nucleus. This is also
accompanied by release of energy. Both fission and
fusion processes are accompanied by the release of
energy.
(iii) Packing Fraction
Packing fraction is a very important quantity and is a
measure of the stability of the nucleus. In general, the
lower the packing fraction of an element, the greater
is the stability of its nucleus.

A negative packing fraction implies exceptional


nuclear stability. If a nuclide has a zero or low positive
value for its packing fraction, then the nuclide would
be stable. The nuclei having very high positive values
of packing fraction are unstable and hence such
nuclei undergo spontaneous disintegration.

The plot in Figure shows how the magnitude of


packing changes with the change of mass number. It
reaches a minimum at mass number 45. This implies
that transition elements like 55Mn, 56Fe, etc. having
the mass numbers in the neighborhood of 45 have
the lowest negative values for their packing fraction
and hence the nuclei of these isotopes are highly
stable.
(iv) Even-Odd Number of Protons and Neutrons
The stability of a nucleus depends on the number of
protons and neutrons. Table shows that the number
of stable nucleus is maximum when both Z and N are
even numbers suggesting a tendency to form p-p and
n-n pairs as conducive to stability. On the other hand
the number of stable nucleus in which either the Z or
N is odd is about a third of those where both are even.
By comparison, only 5 stable isotopes have an odd
number of protons and an odd number of neutrons.
Table : Distribution of Stable and Unstable Isotopes based on
Neutron and Proton Numbers

# of stable
Proton number (Z) Neutron Number
Isotopes

Even Even 164

Even Odd 55

Odd Even 50

Odd Odd 5

(v) The Neutron to Proton Ratio, N/Z


When a graph is made of N vs. Z, the stable nuclei
are found to lie within a narrow region called the
stability band. Stability apparently requires the N/Z
ratio to be about 1.0 for the lightest nuclei and to
increase up to about 1.5 for the heaviest nuclei. The
need for excess neutrons to accompany protons in
heavy nuclei suggests that neutrons play an
important role in counteracting proton-proton
repulsion.

A nucleus outside the stability band will decay by a


mechanism that brings its N/Z ratio into the stable
region. Nuclides to the left of the band usually decay
by  emission whereas those to the right usually
decay by  emission or electron capture. Nuclides of
Z > 83 often decay by emission.

Radioactive Decay: The process in which a nucleus


spontaneously disintegrates, giving off radiation.
There are five common types of radioactive decay:

(1) Alpha Emission (): Emission of a 42𝐻𝑒 nucleus


or  particle, from an unstable nucleus.
226
88𝑅𝑎  222
86𝑅𝑛 + 4
2𝐻𝑒
The product nucleus has an atomic number that is
two less, and a mass number that is four less, than
that of the original nucleus.

(ii) Beta Emission (): Beta Decay is a type of


radioactive decay in which a neutron is altered into a
proton inside the nucleus of the radioactive sample.
Emission of a high-speed electron from an unstable
nucleus. e.g. emission is equivalent to the
conversion of a neutron to a proton. The decay
converts the initial nuclide into an isobar of that
nuclide.
0𝑛→ 1𝑝 + −1𝑒 + 
1 1 0
̅
6𝐶  7𝑁 + −1𝑒
14 14 0

The product nucleus has an atomic number that is


one more than that of the original nucleus. The mass
number remains the same.
(iii) Positron Emission (): Beta Decay is a type of
radioactive decay in which a proton is transformed
into a neutron inside the nucleus of the radioactive
sample. Emission of a positron (a particle identical to
an electron in mass but having a positive instead of a
negative charge) from an unstable nucleus, e.g.
emission is equivalent to the conversion of a proton
to a neutron.
1𝑝→ 0𝑛 + +1𝑒 + 
1 1 0

95
43𝑇𝑐  42
95
𝑀𝑜 + 0
+1𝑒
The product nucleus has an atomic number that is
one less than that of the original nucleus. The mass
number remains the same.

(iv) Electron Capture (EC): The decay of an


unstable nucleus by capturing or picking up, an
electron from an inner orbital of an atom.
19𝐾 + −1𝑒  18𝐴𝑟
40 0 40

In effect, a proton is changed to a neutron, as in


positron emission,
1𝑝 + −1𝑒  0𝑛
1 0 1

The product nucleus has an atomic number that is


one less than that of the original nucleus. The mass
number remains the same. (When another orbital
electron fills the vacancy in the inner-shell orbital
created by electron capture, an X-ray photon is
emitted).

(v) Gamma Emission (): Emission of a gamma


photon from an excited nucleus, e.g.
99𝑚
43𝑇𝑐  43
99
𝑇𝑐 + 00
A metastable nucleus is a nucleus in an excited state
with a lifetime of at least one nanosecond, 10-9 s. The
product nucleus is simply a lower-energy state of the
original nucleus, so there is no change of atomic
number or mass number.

Rate of radioactive decay


The decay of a radioactive isotope takes place by
disintegration of the atomic nucleus. It is not
influenced by any external conditions. Therefore the
rate of decay is characteristic of an isotope and
depends only on the number of atoms present. If N
be the number of undecayed atoms of an isotope
present in a sample of the isotope, at time t,

where – dN/dt means the rate of decrease in the


number of radioactive atoms in the sample; and λ is
the proportionality factor. This is known as the decay
constant or disintegration constant. Putting dt = 1 in
above equation. We have
– dN/N = 
Thus decay constant may be defined as the
proportion of atoms of an isotope decaying per
second.

Units of radioactivity
The standard unit of radioactivity (i.e. rate of
disintegration) is Curie (c). A curie is a quantity of
radioactive material decaying at the same rate as 1 g
of Radium (3.7 × 1010 dps). Rutherford is a more
recent unit.
1 Curie = 3.7 × 1010 dps
1 Rutherford = 106 dps
The S.I. unit is Becquerel
1 Bq = 1 dps

The activity of a radioactive substance


It is defined as the rate of decay or the number of
disintegrations per unit time. Alternatively, the
amount of substance (in gram) which gives 3.7 × 1010
disintegration per second. The activity of a sample is
denoted by A. It is given by the expression:
A = dN/dt = N
Unit: Curie (Ci) or Becquerel (Bq).
Average life
In a radioactive substance, some atoms decay earlier
and others survive longer. The statistical average of
the lives of all atoms present at any time is called the
Average life. It is denoted by the symbol  and has
been shown to be reciprocal of decay constant, .
 = 1/
The average life of a radioactive element is related to
its half-life by the expression:
Average life = 1.44 × Half-life
or  = 1.44 × t1/2
The average life is often used to express the rate of
disintegration of a radioactive element. The average
life of radium is 2400 years.
SOLVED PROBLEM 1. Calculate the half-life of
radium-226 if 1 g of it emits 3.7 × 1010 alpha particles
per second.
SOLVED PROBLEM 3. The half-life period of
radon is 3.825 days. Calculate the activity of radon.
(atomic weight of radon = 222)
SOLVED PROBLEM 1. Cobalt-60 disintegrates to
give nickel-60. Calculate the fraction and the
percentage of the sample that remains after 15 years.
The disintegration constant of cobalt-60 is 0.13 yr–1.
SOLVED PROBLEM 2. How much time would it
take for a sample of cobalt-60 to disintegrate to the
extent that only 2.0 per cent remains? The
disintegration constant is 0.13 yr–1.

SOLVED PROBLEM 3. A sample of radioactive


133
I gave with a Geiger counter 3150 counts per
minute at a certain time and 3055 counts per unit
exactly after one hour later. Calculate the half life
period of 133I.
27. The activity of a radioactive sample falls to 85 %
of the initial value in four years. What is the half life of
the sample? Calculate the time by which activity will
fall by 85 %. (Tuli-141)
Answer: 17.05 years; 46.735 years
Exercise: 2 (b), 3 (b), 7(b), 10(b), 12, 16, 17(b), 18
(b), 20 - 22, 27, 43, 45, 48(b), 49(b)
MCQ: 27, 30, 31, 34, 35
Reaction Mechanisms: For projectiles with less
than 50 MeV energy, there are two mechanisms that
predominate: compound-nucleus formation and
direct interactions.
(i) Compound-nucleus Formation: In the
compound-nucleus mechanism, the projectile strikes
the target nucleus and is assimilated into it to form a
compound nucleus (lifetime 10-14 to 10-20 s). The
projectile adds its kinetic energy and binding energy
to the target nucleus, so the compound nucleus is in
an excited state. The energy brought in by the
projectile is transferred to the other nucleons through
numerous collisions until a state of statistical
equilibrium is reached. At this time the original
particle is indistinguishable from the other nucleons.
If sufficient energy is present, eventually one nucleon
or a composite of nucleons, will acquire enough
energy to overcome nuclear binding forces and
escape from the nucleus. This process is called
evaporation, because it is analogous to the way in
which a molecule in the liquid phase acquires
sufficient energy to escape into the vapor phase. If
residual energy remains after one nucleon or
composite unit has been evaporated, more may be
emitted, until there is too little energy left to cause
particle emission. The nucleus may then further de-
excite via a  transition.

Thus it is a two step processes: (1) capture of the


projectile to form a compound nucleus, (ii) decay of
the compound nucleus via particle evaporation and/
or -ray emission.

Nuclear capture reactions are good examples of


processes that proceed via compound- nucleus
formation. For thermal neutrons, the compound
nucleus does not usually have enough energy for
particle emission to take place. Therefore, emission
of a -ray is the primary means of de-excitation.
These (n, ) reactions are examples of radiative
capture reactions. Neutrons with higher energies do
excite the compound nucleus sufficiently to cause
particle emission. Reactions such as (n, n'), (n, p),
and (n, ) are common with the 14 Mev neutrons.

Reactions induced by charged particles also may


proceed via compound-nucleus formation. Charged
particle induced reactions will almost always result in
particle evaporation.

(ii) Direct Interaction: In contrast to compound-


nucleus formation, direct interactions do not result in
the assimilation of the projectile into the nucleus.
Instead, the incoming particle interacts directly with
the nucleons at the periphery of the nucleus in a very
fast process (10-22 s) that involves the prompt
emission of a nucleon. Direct interaction happens in
only one-step, with no time for equilibration between
the initial interaction and resulting nuclear change.
Therefore, the decay modes are related to the means
by which the nucleus was originally excited. Direct
interactions may occur by either Lock-on or transfer
processes.

In the knock-on mechanism, a collision between the


projectile and an outer nucleon results in the direct ejection
of one, or more particles. This type of process is most
commonly observed for projectiles with higher energies
(~50 MeV).

Transfer mechanisms can be divided into stripping and


pickup reactions. Transfer reactions are most often
observed for particles with energies less than 50 MeV. The
deuteron-induced reactions

are the examples of transfer processes. In the pickup


reactions, the deuteron gained either a proton (d, 3He) or
another deuteron (d, ) from the 59Co target. The stripping
reactions show the loss of either a proton (d, n) or a neutron
(d, p) from the deuteron to the 59 Co.

Mechanisms of Fission Reactions: N. Bohr and J.


Wheeler proposed the liquid drop model in the late
1930s in explaining the phenomenon of nuclear
fission. Fission reaction may be induced by neutrons,
charged particles like protons, deuterons or 'He, but
the neutron induced fission reaction is more
common. In this process, a neutron is captured by a
heavy nuclide to form a compound nucleus. This
compound nucleus then splits into two smaller
fragments and emits two or three free neutrons, along
with about 200 MeV of energy per nucleus fissioned.
Fig. 4 gives a schematic illustration of how neutron-
induced fission is thought to occur.

In the target nucleus, there are two competing forces


at work: a binding force that holds the nucleons
together and a Coulomb force that tends to push
them apart. For low-mass nuclides, the binding force
is much stronger than the repulsive force, so they are
not subject to fission. For heavier nuclides, the two
forces are more equal in magnitude, and it is possible
to shift the balance in favor of the repulsive forces by
addition of a neutron. The neutron provides some
internal excitation energy that deforms the nucleus
into a transition state. If the deformation is severe
enough, the Coulomb forces take over and split the
nucleus into two fragments. These fission fragments
are repelled away from each other with considerable
kinetic energy, due to the Coulomb repulsive forces.
The fragments are neutron rich, so they will decay by
negatron emission, often followed by y-tration, until a
favorable N/Z ratio is attained and a stable product
results.

Carbon-14 Dating
Radiocarbon dating is a method that provides
objective age estimates for carbon-based materials
that originated from living organisms. An age could
be estimated by measuring the amount of carbon-14
present in the sample and comparing this against an
internationally used reference standard. It has found
applications in geology, hydrology, geophysics,
atmospheric science, oceanography,
paleoclimatology and even biomedicine.
Natural carbon is composed of three isotopes. It is
98.89% carbon-12, 1.11% carbon-13, and
0.00000000010% carbon-14. The last of these,
carbon-14, is most important in radiocarbon dating.
Carbon-14 atoms are constantly being produced in
our upper atmosphere through neutron
bombardment of nitrogen atoms.

Once formed, the carbon-14 is quickly oxidized to


produce carbon dioxide, CO2, which is then
converted into many different substances in plants.
When animals eat the plants, the carbon-14 becomes
part of the animals too. For these reasons, carbon-14
is found in all living things. This radioactive isotope is
a beta emitter with a half-life of 5730 years (±40
years), so as soon as it becomes part of a plant or
animal, it begins to disappear.

As long as a plant or animal remains alive, its intake


of carbon-14 balances the isotope’s continuous
decay, so that the ratio of 14C to 12C in living tissues
remains constant at about 1 in 109
(1,000,000,000,000). When the plant or animal dies,
it stops taking in fresh carbon, but the carbon-14 it
contains continues to decay. Thus the ratio of 14C to
12
C drops steadily. Therefore, to date an artifact, a
portion of it is analyzed to determine the 14C/12C ratio,
which is then used to calculate its age.
For example, if the 14C/12C ratio had dropped to one
half of the ratio found in the air today, the object
would have been described as about 5730 years old.
A 14C/12C ratio of one-fourth of the ratio found in the
air today would date it as 11,460 years old (2 half-
lives).

For example, it is now believed that the large


quantities of carbon periodically released into the
atmosphere by volcanoes have in many cases been
isolated from the air for so long that they have much
lower than average levels of carbon-14. The levels of
cosmic radiation also fluctuate, and lower levels of
neutron bombardment produce lower levels of
carbon-14.

By checking the 14C/12C ratio of the wood in tree rings


(which are formed once a year), scientists discovered
that the ratio has varied by about ±5% over the last
1500 years. Further study of very old trees, such as
the bristlecone pines in California, has allowed
researchers to develop calibration curves for
radiocarbon dating that go back about 10,000 years.
These calibration curves are now used.
Methodology of carbon dating
 Samples of wood, charcoal or peat are first
carefully picked over to remove any modern
rootlets.
 Then the sample is treated with hot, dilute NaOH
for 24 hours to dissolve away the humic acid.
 After washing with distilled water the sample is
treated with in hydrochloric acid for 48 hours to
remove carbonate contaminants.
 The sample is dried, burned in a stream of
oxygen and the CO2 from the combustion is
collected.
 In the final step the CO2 is converted to methane
for counting of its 14C radioactivity.
 Gas proportional counting is a conventional
radiometric dating technique that counts the beta
particles emitted by a given sample. Beta
particles are products of radiocarbon decay.

SOLVED PROBLEM. The amount of carbon-14 in a


piece of wood is found to be one-sixth of its
amount in a fresh piece of wood. Calculate the age
of old piece of wood. (Tuli-119)
In a new method for determining the age of old wood (or
fossil) the measurement of radioactivity is avoided. The
ratio 14C/12C is found with the help of a mass spectrometer
in the old wood and fresh wood from a living plant. It is
assumed that the ratio 14C/12C in the fresh wood today is
the same as it was at the time of death of the plant.
SOLVED PROBLEM. A bone taken from a garbage
pile buried under a hill-side had 14C/12C ratio 0.477
times the ratio in a living plant or animal. What
was the date when the animal was buried? (Tuli-
120)
29. The cloth shroud from around a mummy is found to
have a 14C activity of 8.9 disintegrations per minute per
gram of carbon as compared with living organisms that
undergo 15.2 disintegrations per minute per gram of
carbon. From the half-life for 14C decay, 5.73 × 103 yr,
calculate the age of the shroud. (Tuli-145)
43. Calculate the half life and average life period of a
radioactive element if its decay constant is 7.37 × 10–3
hour–1. (Tuli-142)
Answer. 0.0261 sec; 0.0376 sec
45. Radioisotope 32P has a half life of 15 days. Calculate
the time in which the radioactivity of its 1 mg quantity will
fall to 10 % of the initial value.
Answer: 49.84 days
46. An old wooden article shows 2.0 counts per minute per
gram. A fresh sample of wood shows 15.2 counts per
minute per gram. Calculate the age of the wooden article
(t½ of 14C = 5460 years)
Answer. 16861 years
48. A radioactive isotope has half-life period of 20 days.
What is the amount of the isotope left over after 40 days if
the initial concentration is 5 g?
49. 2 g of radioactive element degraded to 0.5 g in 60
hours. In what time will it be reduced to 10 % of its original
amount?
PP: 23, 25, 36, 34(a), 46
MCQ: 29, 33
Nuclear Fission and Electric Power Plants
The reason of the fission of uranium-235 can generate a lot
of energy in a short period of time is that under the right
circumstances, it can initiate a chain reaction, a process in
which one of the products of a reaction initiates another
identical reaction.
To sustain a chain reaction for the fission of uranium-235,
an average of at least one of the neutrons generated in each
reaction must go on to cause another reaction. If this does
not occur, the series of reactions slows down and
eventually stops. At the natural composition of uranium is
99.27% uranium-238, 0.72% uranium-235, and a trace of
uranium-234. When natural uranium is bombarded by
neutrons, the chain reaction cannot be sustained. Too many
neutrons are absorbed by other entities without leading to
fission. The only one of these isotopes that undergoes
fission is uranium-235. Uranium-238 does absorb
neutrons, but the uranium-239 that then forms reacts by
beta emission rather than nuclear fission.
There is another way to increase the likelihood that
neutrons will be absorbed by uranium-235 instead of
uranium-238. Both uranium-235 and uranium-238 absorb
fast neutrons, but if the neutrons are slowed down, they are
much more likely to be absorbed by uranium-235 atoms.
Therefore, in a nuclear reactor, the fuel rods are
surrounded by a substance called a moderator that slows
the neutrons as they pass through it. Several substances
have been used as moderators, but normal water is most
common.

Another problem associated with the absorption of


neutrons by uranium-238 is that it leads to the creation of
plutonium-239 (look again at the series of equations
presented above). This is a cause for concern because
plutonium-239 undergoes nuclear fission easier than
uranium-235, and is thus a possible nuclear weapon fuel.
Moreover, plutonium has a relatively long half-life, so the
radioactive wastes from nuclear reactors (which contain
many other unstable nuclides besides plutonium) must be
carefully isolated from the environment for a very long
time.

An efficient nuclear reactor needs to sustain the chain


reaction but should not allow the fission reactions to take
place too rapidly. For this reason, nuclear power plants
have control rods containing substances such as cadmium
or boron, which are efficient neutron absorbers. At the first
sign of trouble, these control rods are inserted between the
fuel rods, absorbing the neutrons that would have passed
from one fuel rod to another and preventing them from
causing more fission reactions. The deployment of the
control rods stops the chain reaction.

The control rods serve another purpose in the normal


operation of the power plant. When fresh fuels rods are
introduced, the control rods are partially inserted to absorb
some of the neutrons released. As the uranium-235 reacts
and its percentage of the total mixture in the fuel rods
decreases, the control rods are progressively withdrawn. In
this way, a constant rate of fission can be maintained, even
as the percentage of the fissionable uranium-235
diminishes (Figure 16.6).

The Effect of Radiation on the Body


Alpha particles, beta particles, and gamma photons are
often called ionizing radiation, because as they travel
through a substance, they strip electrons from its atoms,
leaving a trail of ions in their wake.
Picture an alpha particle moving through living tissue at
up to 10 % the speed of light. Remember that alpha
particles are helium nuclei, so they each have a +2 charge.
As such a particle moves past, say, an uncharged water
molecule (a large percentage of our body is water), it
attracts the molecule’s electrons. One of the electrons
might be pulled toward the passing alpha particle enough
to escape from the water molecule, but it might not be able
to catch up to the fast-moving alpha particle. Instead, the
electron is quickly incorporated into another atom or
molecule, forming an anion, while the water molecule that
lost the electron becomes positively charged. The alpha
particle continues on its way, creating many ions before
slowing down enough for electrons to catch up with it and
neutralize its charge. When a neutral water molecule,
which has all of its electrons paired, loses one electron, the
cation that is formed has an unpaired electron. Particles
with unpaired electrons are called free radicals.

The water cations, H2O•+, and free electrons produced by


ionizing radiation react with uncharged water molecules to
form other ions and free radicals. These very reactive ions
and free radicals then react with important substances in
the body, leading to immediate tissue damage and to
delayed problems, such as cancer.

The cells that reproduce most rapidly are the ones most
vulnerable to harm, because they are the sites of greatest
chemical activity. This is why nuclear emissions have a
greater effect on children, who have larger numbers of
rapidly reproducing cells, than on adults. The degree of
damage is, of course, related to the length of exposure, but
it is also dependent on the kind of radiation and whether
the source is inside or outside the body.

If radioactive strontium-90 is ingested, it concentrates in


the bones in substances that would normally contain
calcium. This can lead to bone cancer or leukemia. For
similar reasons, radioactive cesium-137 can enter the cells
of the body in place of its fellow alkali metal potassium,
leading to tissue damage.
Non-radioactive iodine and radioactive iodine-131 are
both absorbed by thyroid glands. Because iodine-131 is
one of the radioactive nuclides produced in nuclear power
plants, the Chernobyl accident released large quantities of
it. To reduce the likelihood of thyroid damage, people were
directed to take large quantities of salt containing non-
radioactive iodine-127. This flooding of the thyroid glands
with the non-damaging form of iodine made it less likely
that the iodine-131 would be absorbed.
Cobalt-60 emits ionizing radiation in the form of beta
particles and gamma photons. Gamma photons, which
penetrate the body and damage the tissues, do more
damage to rapidly reproducing cells than to others. This
characteristic coupled with the fact that cancer cells
reproduce very rapidly underlies the strategy of using
radiation to treat cancer. Typically, a focused beam of
gamma photons from cobalt-60 is directed at a cancerous
tumor. The ions and free radicals that the gamma photons
produce inside the tumor damage its cells and cause the
tumor to shrink.

Like many other radioactive nuclides used in medicine,


cobalt-60 is made by bombarding atoms of another
element with neutrons. The iron contains a small
percentage of iron-58, which forms cobalt-60 in the
following steps:
For example, fluorine-18 containing substances collect in
the bones, so that nuclide is used in bone scanning.
Glucose molecules are used throughout the body, but are
especially concentrated in the brain, so glucose
constructed with carbon-11 can be used to study brain
function.
Positron emission tomography (PET) scans reveal which
parts of the brain use glucose most actively during
different activities. This ability to study dynamic processes
in the body, such as brain activity or blood flow, makes the
PET scan a valued research and diagnostic tool.
Table 16.4 Uses for Radioactive Nuclides

Nuclide Nuclear change Application


Argon-41 beta emission Measure flow of gases from
smokestacks
Barium-131 electron capture Detect bone tumors
Carbon-11 positron emission PET brain scan
Carbon-14 beta emission Archaeological dating
Cesium-133 beta emission Radiation therapy
Cobalt-60 beta emission Cancer therapy
Copper-64 beta emission, Lung and liver disease
positron diagnosis
emission, electron
capture
Chromium- electron capture Determine blood volume and
51 red blood cell lifetime;
diagnose gastrointestinal
disorders
Fluorine-18 beta emission, Bone scanning; study of
positron cerebral sugar metabolism
emission, electron
capture
Gallium-67 electron capture Diagnosis of lymphoma and
Hodgkin disease; whole
body scan for tumors
Gold-198 beta emission Assess kidney activity
Hydrogen-3 beta emission Biochemical tracer;
measurement of the water
content of the body
Indium-111 gamma emission Label blood platelets
Iodine-125 electron capture Determination of blood
hormone levels
Iodine-131 beta emission Measure thyroid uptake of
iodine
Iron-59 beta emission Assessment of blood iron
metabolism and diagnosis of
anemia
Krypton-79 positron emission Assess cardiovascular
and electron function
capture
Nitrogen-13 positron emission Brain, heart, and liver
imaging
Oxygen-15 positron emission Lung function test
Phosphorus- beta emission Leukemia therapy, detection
32 of eye tumors, radiation
therapy, and detection breast
carcinoma
Polonium- alpha emission Radiation therapy
210
Potassium- beta emission Geological dating
40
Radium-226 alpha emission Radiation therapy
Selenium- beta emission and Measure size and shape of
75 electron capture pancreas
Sodium-24 beta emission Blood studies and detection
of blood clots
Technetium- gamma emission Bone scans and detection of
99 blood clots
Xenon-133 beta emission Lung capacity measurement

Nuclear hazards: Nuclear hazards are threat posed by the


invisible and odorless contamination of the environment
by the presence of radioactive materials such as radio-
nuclides in air water or soil. These radio-nuclides emit
high energy particles (alpha and beta rays) and
electromagnetic radiations (gamma rays). Radio nuclides
are elements, such as uranium 235, uranium 238, thorium
232, potassium 40, radium 226, carbon 14 etc., with
unstable atomic nuclei and release ionizing radiations in
the form of alpha, beta and gamma rays.

Radioactive or Nuclear pollution - It can be defined as


the release of radioactive substances or high energy
particles into the air water, or earth mostly as a result of
human activity, either by accident or by design. Sometimes
natural sources of radioactivity, such as radon gas emitted
from beneath the ground, are considered pollutants when
they become a threat to human health. Radio-nuclides
occur naturally in our environment. They are even found
in human bodies and every day we ingest or inhale these
radio-nuclides through air, water or food. Out of the
known 450 radioisotopes only some are of environmental
concern like strontium 90, tritium, plutonium 239, argon
41, cobalt 60, cesium 137, iodine 131, krypton 85 etc.
These can be both beneficial and harmful, depending on
the way in which they are used.

Control of nuclear hazards


a. Leakages from nuclear reactors, careless handling,
transport and use of radioactive fuels, fission products and
radioactive isotopes have to be totally stopped.
b. Safety measures should be enforced strictly and
strengthened against nuclear accidents.
c. There should be regular monitoring and quantitative
analysis through frequent sampling in the risk areas.
d. Preventive measures should be followed so that
background radiation levels do not exceed the permissible
limits.
e. Appropriate steps should be taken against occupational
exposure.
f. Waste disposal must be careful, efficient and effective.
g. In India, a Waste Immobilization Plant (WIP) was
commissioned in 1985 at Tarapore.

Precautions after the disposal of nuclear waste


1) Monitoring radioactivity around the disposal sites.
2) Prevention of erosion of radioactive waste disposal
sites.
3) Prevention of any drilling activity in and around the
waste disposal site.
4) Periodic and long-term monitoring of such disposal sites
and areas of naturally occurring uranium rich rocks.
Crystal Lattice and Crystal Symmetry
Books
1. G. H. Stout and L. H. Jensen, X-ray Structure
Determination – A Practical Guide, John Wiley
& Sons.
2. A. R. West, Solid State Chemistry and its
Applications, John Wiley & Sons.

Crystal lattice: The regular Lattice points (may


molecules or ions)
be atoms,

array of points, which defines


the position of atoms,
molecules or ions in a crystal,
Crystal lattice
is called crystal lattice. Each
point must have the same
number of neighbors as every
other point and the neighbors
must always be found at the
same distances and
directions. All points are in
the same environment.

Lattice point: The point that describes the position of


an atom, molecule or ion in a crystal is called lattice
point.
Motif: A motif is a unit of pattern. In a crystal, it
is an atom, a molecule, an ion or a group of atoms,
molecules or ions.

Lattice + Motif Crystal

Unit cell: The smallest repeating unit of a crystal


lattice is called unit cell. It contains an integer number
of lattice points. A unit cell has one lattice point at
each corner, and there may be lattice points in the
faces, and in the interior of the cell as well.
Z
c

 b
 Y

a
X Unit cell
Unit cell

The unit cell is characterized by six parameters, three


axial lengths and three interaxial angles. The lengths
of the unit cell edges are designated a, b, c, and the
interaxial angles α, β, γ. The angle α is between b and
c, β is between c and a, while γ is between a and b.

The entire crystal consists of a large number of unit


cells adjacent to one another in all three dimensions.
The unit cell of a crystal possesses all the structural
properties of the given crystal. The unit cell can be
described as a parallelepiped defined by the cell edges
a, b, and c which are vectors. Its volume can be
calculated by the scalar triple product, V = (a · b × c,
where a, b, and c are the lattice vectors) and
corresponds to the square root of the determinant of
the metric tensor.

Crystal system: The crystal system is a grouping of


crystal structures that are categorized according to the
axial system used to describe their lattice. Each
crystal system consists of a set of three axes in a
particular geometrical arrangement.
The seven unique crystal systems are:
Crystal system No. of Examples
independent
parameters
Triclinic 6 K2Cr2O7,
a ≠ b ≠ c, H3BO3
α ≠ β ≠ γ ≠ 90o
Monoclinic 4 NaHCO3,
a ≠ b ≠ c, S
α = γ = 90o, β >
90o
Orthorhombic 3 KNO3,
a ≠ b ≠ c, BaSO4
α = β = γ = 90o
Tetragonal 2 SnO2,
a = b ≠ c, TiO2
α = β = γ = 90o
Cubic 1 NaCl,
a = b = c, KCl
α = β = γ = 90o
Hexagonal 2 SiO2,
a = b ≠ c, ZnO
α = β = 90o, γ
=120o

Rhombohedral 2 CaSO4,
a = b = c, KMnO4
α = β = γ ≠ 90o

Unit cell volume: The dimensions of the unit cell of


each crystal system give its volume and are calculated
as follows:
Crystal Volume
system
Triclinic abc√1 - cos2α – cos2β – cos2γ +
2cosα cosβ cosγ
Monoclinic abc sinβ
Orthorhombic abc
Tetragonal a2 c
Cubic a3
3√3 a2c
---------
Hexagonal
2

Rhombohedral a3√1 – 3cos2α + 2cos3α

where a, b, and c are the unit cell axes dimensions and


α, β, and γ are the inclination angles of the axes in the
unit cell.
Problem: A 1,10-phenanthrolinium(1+)
tetraphenylborate compound crystallizes in a
monoclinic crystal system, with a = 10.09 Å, b =
15.03 Å, c = 7.03 Å and β = 97o. Calculate the volume
of its unit cell.
Solution:
The unit cell of 1,10-phenanthrolinium(1+)
tetraphenylborate compound is monoclinic with a =
10.09 Å, b = 15.03 Å, c = 7.03 Å and β = 97o.
The volume of its unit cell is (10.09 Å)(15.03 Å)(7.03
Å) sin 97o = 1058.17 Å3
Classification of unit cell:
Unit cell

Primitive unit cell Nonprimitive unit cell


Lattice points present only at the Lattice points present positions
corner positions other than corners in addition to
those at corners

Side centered unit cell Face centered unit cell Body centered (Innenzentrierte)
Having additional Having additional unit cell
lattice points on the lattice points at the Having an additional
centres of two centres of its six lattice point at the
opposite faces faces centre

Primitive Side Face Body


(P, H or centered (A, centered centered (I)
R) B or C) (F)
Non primitive
A primitive cell consists of 8 lattice ⅛ lattice
points at the 8 corners of the unit cell. point at 8
corners
Each lattice point is shared equally
between 8 adjacent unit cells, and
therefore the number of lattice points
present in a primitive unit cell = ⅛ × 8
= 1.

A side-centered unit cell consists of 8 ⅛ lattice


lattice points at the 8 corners, each point at 8
corners
shared by 8 unit cells and 2 lattice
points at the center of the two opposite
faces, each shared equally by 2
½ lattice
adjacent unit cells. Thus the total
point at 2
number of lattice points present in side- faces
centered unit cell ⅛×8 + ½×2 = 2.
A face-centered unit cell consists of 8 ⅛ lattice
lattice points, each shared by 8 unit point at 8
corners
cells at 8 corners and 6 face-centered
lattice points, each shared by 2 cells.
Thus the total number of lattice points ½ lattice
point at 6
present in face-centered unit cell ⅛×8 faces
+ ½×6 = 4.

A body-centered unit cell consists of 8 ⅛ lattice


lattice points at the 8 corners, each point at 8
corners
shared by 8 unit cells and one lattice
point at the body center, which wholly
belongs to the unit cell. Therefore, the 1 atom at
center
total number of lattice points present in
the body-centered unit cell = ⅛×8 + 1
= 2.
Bravais lattices: The great French
crystallographer Auguste Bravais
(1850) first demonstrated that there
could be only 14 (7 primitive and 7 Auguste
non-primitive) different types of space Bravais
lattices. This means that there are only
14 different types of three-dimensional
arrays whereby atoms, molecules, or
ions may pack together when forming
a crystal lattice. Because of Bravais
contributions to their study, these space
lattices are called Bravais lattices.

These lattices have seven different unit cell shapes


with different symmetry properties corresponding to
the seven lattice systems.

The Bravais lattices


System No. of Lattice No. of Nature of
lattices symbols lattice unit cell
in this points axes and
system angles

Triclinic 1 P 1 a ≠ b ≠ c, α
≠β≠γ≠
90o
Monoclinic 2 P 1 a ≠ b ≠ c, α
C 2 = γ = 90o,
β > 90o
Orthorhom 4 P 1 a ≠ b ≠ c, α
bic C 2 =β=γ=
I 2 90o
F 4
Tetragonal 2 P 1 a = b ≠ c,
I 2 α=β=γ=
90o
Cubic 3 P 1 a = b = c,
I 2 α=β=γ=
F 4 90o
Hexagonal 1 P 1 a = b ≠ c, α
= β = 90o,
γ =120o
Rhombohe 1 R 1 a = b = c,
dral α=β=γ≠
90o

The symbol ≠ implies non-equality by reason of


symmetry; accidental equality may, of course, occur.

Not all combinations of the crystal systems and lattice


centerings are unique. There are in total 7 × 6 = 42
combinations, but it can be shown that several of
these are in fact equivalent to each other. For
example, the monoclinic I lattice can be described by
a monoclinic C lattice by different choice of crystal
axes. Similarly, all A- or B-centered lattices can be
described either by a C- or P-centering. This reduces
the number of combinations to 14 conventional
Bravais lattices.

When the fourteen Bravais lattices are combined with


the 32 crystallographic point groups, we obtain
the 230 space groups.
Unit cells of the 14 Bravais lattices
Crystal Lattices
system

Triclinic
Primitive
(P)
Monoclinic C-
Primitive
centered
(P)
(C)

Orthorhombi Primitive C- Body- Face-


(P) centered centered centered
c (C) (I) (F)

Tetragonal
Body-
Primitive
centered
(P)
(I)

Cubic Body- Face-


Primitive
centered centered
(P) (F)
(I)

Hexagonal
Primitive
(H)
Rhombohedr
al
Primitive
(R)

Any space lattice corresponds to one or other of the


fourteen shown in figure and no other distinct space
lattices can occur. For example, there are only two
distinct monoclinic lattices, described by P and C
cells, and not amenable one to the other (shown
below).
c c c
c'
c' = c

c' b' = b c'


a b' = b a a b' = b b'
= = = b
a' a' a' a
a'
(d)
(a) (b) (c)

Figure: Monoclinic lattices: (a) reduction of a B-


centered cell to a Primitive cell; (b) reduction of an
Body-centered cell to an A-centered cell; (c)
reduction of an F-centered cell to a C-centered cell;
(d) reduction of a C-centered cell to a Primitive non-
monoclinic cell.

Miller indices: British c

mineralogist W. H. Miller in 1839 l


a
h
developed the very useful system k

of indexing of lattice planes.


The Miller plane is the plane which intercepts the
three axes of a unit cell at the points a/h, b/k, c/l,
where a, b and c are the unit cell vectors. Thus the
Miller indices are proportional to the inverses of
the intercepts of the plane with the unit cell. The
indices h, k and l are integers, and are
conventionally enclosed in round bracket as (hkl).

If one or more of the indices is zero, it simply means


that the planes do not intersect that axis (i.e. the
intercept is at infinity). Negative integers are written
with a bar, as in 3̅ for −3. The precise meaning of this
notation depends upon a choice of lattice vectors for
the crystal.

h k l = 1/1 h k l = h k l = 1/1 h k l = 1/-1


1/∞ 1/∞ 1/1 1/1 1/∞ 1/1 1/1 1/∞ 1/∞
= (100) = (110) = (111) = (͞1 00)

c (200) c (020)

b b
a a

h k l = 1/½ h k l = 1/∞ h k l = 1/∞ h k l = 1/½


1/∞ 1/∞ 1/½ 1/∞ 1/∞ 1/½ 1/½ 1/1
= (200) = (020) = (002) = (221)
Planes with different Miller indices in crystals
The symbol { } is used to indicate sets of planes that
are equivalent; for example, the sets (100), (010) and
(001) are equivalent in cubic crystals and may be
represented collectively as {100}.

The procedure for determining the Miller indices for


a cubic crystal plane is as follows:
(i) Choose a plane that does not pass through the
origin at (0, 0, 0).
(ii) Determine the intercepts of the plane in terms of
the crystallographic x, y, and z axes for a unit cube.
These intercepts may be fractions.
(iii) Form the reciprocals of these intercepts.
(iv) Clear fractions and determine the smallest set of
whole numbers that are in the same ratio as the
intercepts.
These whole numbers are the Miller indices of the
crystallographic plane and are enclosed in
parentheses.

Problem 1: Determine the Miller Indices of a plane


which is parallel to x-axis and cuts intercepts of 2 and
½, respectively along y and z axes.
Solution:
(i) Intercepts  2b ½c
(ii) Division by unit /a =  2b/b = 2 c/2c = ½
translation
(iii) Reciprocals 1/ ½ 2
(iv) After clearing 0 1 4
fraction
(multiply by 2)

Therefore, the required Miller indices of the plane


are (014).

Problem 2: Determine the Miller Indices of a plane


that makes intercepts of 2Å, 3 Å, 4 Å on the co-
ordinate axes of an orthorhombic crystal with a:b:c =
4:3:2.
Solution: Here the unit translations are a = 4, b = 3
and c = 2 following the same procedure:

(i) Intercepts 2 3 4
(ii) Division by unit 2/4 = ½ 3/3 = 1 4/2 = 2
translation
(iii) Reciprocals 2 1 ½
(iv) After clearing 4 2 1
fraction
(multiply by 2)

Therefore, the Miller indices of the plane are (421).

Note that when Miller indices are of double digit,


these are separated by commas e.g. (5, 12, 15).
Miller indices are also sometimes useful to calculate
the perpendicular distance dhkl between parallel
planes. Consider the (hk0) planes of a rectangular
lattice built from an orthorhombic unit cell of sides of
lengths a and b. We can write the following
trigonometric expressions for the angle φ shown in
the illustration.
sinφ = dhkl/(a/h) = hdhkl/a,
cosφ = dhkl/(b/k) = kdhkl/b
Then because sin2φ + cos2φ = 1,
b/k
 d
we obtain hkl 
a/h
b
h2d2hkl/a2 + k2d2hkl /b2 = 1
a

which we can rearrange into


1/d2hkl = h2/a2 + k2/b2

In three dimensions, this

expression becomes

1/d2hkl = h2/a2 + k2/b2 + l2/c2

Thus,
Crystal Perpendicular distance dhkl between parallel

system planes

Cubic 1/d2hkl = (h2 + k2 + l2) /c2


Tetragonal 1/d2hkl = h2/a2 + k2/a2 + l2/c2
Orthorhombic 1/d2hkl = h2/a2 + k2/b2 + l2/c2
Hexagonal 1/d2hkl = 4/3a2 (h2 + hk + k2) + l2/c2
Rhombohedral
1 (h2 + k2 + l2) sin2α + 2(hk + kl + lh) (cos2α – cosα)
1/d2hkl = ---- × --------------------------------------------------------------
a2 1 + 2cos3α – 3cos2α

Monoclinic
h2 l2 2hl cosβ
----- + ---- - -------------
a2 c2 ac k2
1/d2hkl = ----------------------------------- + -----
sin2β b2

Triclinic
h2 k2 l2 2hk
----- sin α + ----- sin β + ----- sin2γ + ------ (cosα cosβ – cosγ)
2 2

a2 b2 c2 ab

2kl 2lh
+ ------ (cosβ cosγ – cosα) + ------ (cosγ cosα – cosβ)
bc ca
1/d2hkl = ----------------------------------------------------------------------
1 - cos2α - cos2β - cos2γ + 2 cosα cosβ cosγ

Problem: Find the ratio of interplanar distances of


planes (100), (110) and (111) in a simple cubic lattice.
Solution: Perpendicular distance dhkl between
parallel planes in a simple cubic lattice is
a
dhkl = ----------------------
√(h2 + k2 + l2)

a a
d100 = ---------------------- = ------------------- = a
√(12 + 02 + 02) √(1 + 0 + 0)

a a
d110 = --------------------- = ------------------- = a/√2
√(12 + 12 + 02) √(1 + 1 + 0)
a a
d111 = ---------------------- = ------------------ = a/√3
√(12 + 12 + 12) √(1 + 1 + 1)

d100 : d110 : d111 = a : a/√2 : a/√3

Problem: In a tetragonal lattice a = b = 2.5 Å, c = 1.8


Å. Calculate the lattice spacing between the (111)
planes.
Solution: Perpendicular distance dhkl between
parallel planes in a tetragonal lattice is
1
dhkl = -------------------------------
√(h2/a2 + k2/b2 + l2/c2)

1
d111 = ---------------------------------------Å = 1.26Å
√(12/2.52 + 12/2.52 + 12/1.82)

Crystal symmetry and their operation: A symmetry


element is a geometrical entity such as a line (or axis), a
plane or a point with respect to which one or more
symmetry operations may be carried out. Among the five
crystallographic symmetry elements, three are called
point-symmetry elements and two are space-symmetry
elements. The rotation, mirror reflection and rotatory
inversion are point-symmetry elements and their
operations through a point are point-symmetry operations,
since each leaves at least one point of the object in a fixed
position.

n-Fold rotation axes: An n-fold rotation axis of symmetry


is defined as a line, rotation about which produces
congruent positions (i.e. positions indistinguishable from
the initial position) after rotation through 2π/n. Rotation
axes are described with reference to the value of n, which
must of course be integral: one-fold, two-fold, three-fold,
four-fold, and six-fold.

a a

c c

Two-fold Three-fold Four-fold


rotation, 2 rotation, 3 rotation, 4

Note that five-fold rotation axis is not usually found in


regular crystal lattice. This can be explained by the
following illustration:
Operation of 5-fold rotation B4 A1
B3 A2
axis through A produces A B

lattice points B, B1, B2, B3 B2 A3


B1 A4
and B4.
Operation of 5-fold rotation axis through B produces
lattice points A, A1, A2, A3 and A4.
If the resultant array of points is to form a lattice, it must
be a regular array, in particular the spacing of points on
lines parallel to AB such as B4A1, must be equal to AB or
some multiple thereof. It is evident from figure that
B4A1 = AB – AB4 cos 72o – BA1 cos 72o = AB (1- 2cos
72o) = 0.38 AB
Therefore, the array of points generated by 5-fold rotation
axis through adjacent lattice points is not regular, and
consequently not itself a regular lattice.

Mirror plane: If a plane exists in a structure such that


every part on one side of the plane is related to a part on
the other side as if reflected, the structure is said to possess
a plane of symmetry or mirror plane. The symbol m is used
for this symmetry element.

a a

c c

Mirror plane Mirror plane Mirror plane

(⊥r to the plane) (∥ to the plane)

n-Fold rotatory-inversion axes: A n-fold rotatory


inversion axis implies that a rotation of 2π/n (where n is 1,
2, 3, 4, or 6) followed by inversion through some point on
the axis produces no apparent change in the object or
structure. These axes are symbolized as ͞n.

a a

c c
One fold Two fold Three-fold
inversion, ͞1 inversion, ͞2 inversion, ͞3
Combination of the point-symmetry operations with
translations gives rise to two space-symmetry operations.

n-Fold screw axes: The combination of a rotation axis and


a translation parallel to the axis produces a screw axis. The
direction of such an axis is usually along a unit cell edge
and the translation must be a subintegral fraction of the
unit translation in that direction. Screw axes are designated
by an integer n and a subscript m, where n = 1, 2, 3, 4, or
6 is the multiplicity of the axis and m is an integer less than
n. Thus 21 designates a 2-fold screw axis with a translation
between successive points of ½(= m/n) of a unit
translation.
(ii)
½+
a
(i)

A screw axis consists of a 21 parallel to b


rotation followed by a translation

Coordinates of the two equivalent positions related to each


other by a screw axis (21) parallel to b are (i) x, y, z; (ii) –
x, y + ½, -z.

Glide planes: The combination of a mirror plane and a


translation parallel to the reflecting plane produces a glide
plane. The translation in such a plane is along an edge or
face diagonal of the unit cell and in most cases, of
magnitude half the axial or diagonal length. A glide plane
is designated by a, b or c, if the translation is a/2, b/2, or
c/2 and by n if (a+b)/2, (a+c)/2 or (b+c)/2 i.e. half way
along one of the face diagonal. There is only one additional
type of glide plane - the diamond glide d, characterized by
a translation (a+b)/4, (a+c)/4 or (b+c)/4.
(ii)

a
(i)

A glide plane consists of a Glide plane parallel


reflection followed by a to a, i.e. a glide
translation

Coordinates of the two equivalent positions related to each


other by a glide, (i) x, y, z; (ii) x + ½, y, -z.

Chapter 4: X-ray Diffraction by Crystals


Generation of X-ray: X-rays are a
form of electromagnetic radiation of
high penetrating power. X-rays have a W. C.
wavelength in the range of 0.1 Å to 10 Röntgen
Å and energies in the range of 120 eV
to 120 keV. They are shorter in
wavelength than UV rays and longer
than gamma rays. X-rays are also
called Röntgen radiation, after
Wilhelm Conrad Röntgen, who is
generally credited as its discoverer, and
who had named it X-radiation to
signify an unknown type of radiation.

X-rays are produced in a device called an X-ray tube.


It consists of an evacuated chamber with a tungsten
filament at one end of the tube, called the cathode, and
a metal target at the other end, called an anode.
Electrical current is run through the tungsten filament,
causing it to glow and emit electrons. A large voltage
difference is placed between the cathode and the
anode, causing the electrons to move at high velocity
from the filament to the anode target. Upon striking
the atoms in the target, the electrons dislodge inner
shell electrons resulting in outer shell electrons
having to jump to a lower energy shell to replace the
dislodged electrons. These electronic transitions
results in the generation of X-rays. The X-rays then
move through a window in the X-ray tube. These
emitted X-ray photons have energies that are equal to
the difference between the upper and lower energy
levels of the electron that filled the core hole.
Schematic cross section of an X-ray tube

Thus X-rays are emitted from a target when it is


bombarded with high-energy electrons. These
electrons knocked out the tightly bound electrons in
the K or L electronic shell of the target. The low-
energy empty levels are filled by a falling back of the
electrons from higher energy levels to the inner levels.
The energy liberated during this process is in the form
of X-rays. The wavelength of the emitted X-rays is
dependent upon the energies of the two levels
involved and hence characteristics of the element.
M(3d) M(3p) M(3s) The selection rules for a single electron moving from
L(2p)
L(2s) one orbital to another are
l = ± 1 and
K(1s) j = 0, ± 1
K(1s) L(2s) Forbidden
Nucleus
K = K(1s) L(2p) Allowed
K(1s) M(3s) Forbidden

K = K(1s) M(3p) Allowed

X-rays due to transition from K

the L to the K shell are called Kα Relative


intensity
K
X-rays. Kα1 and Kα2
corresponding to electrons
Wavelength, A

originating in their different


spin states (1/2, 3/2) at the L-
shell. X-rays due to transitions
from the M to K-shell are called
Kβ and so on. The energy
difference between the M and K
levels is larger than that
between the L and K levels,
therefore, the Kβ lines appear at
shorter wavelength. Note that
the Kα line is about 5 - 10 times
as intense as the Kβ line.

The X-ray radiation used for most crystallographic


studies should be as nearly monochromatic as
possible and should have appropriate wavelength.
The Kα lines fulfill this requirement, but the presence
of the accompanying Kβ is a nuisance. Fortunately,
selective filters may be found that will remove the Kβ
to any desired extent, with a relatively much smaller
loss of Kα.
The ideal choice of material K
K-edge

for an X-ray filter is a metal Relative


Ni
intensity
K
whose atomic number, Z, is
one less than that of the
Wavelength, A

anode target metal for first


row transition metals (or two less for second row
transition metals). Because their absorption edges
fall between the Kα and Kβ lines of the target
material. For example, the absorption edge of
nickel metal at 1.488 Å lies between the Kα
(λ = 1.542 Å) and Kβ (λ = 1.392 Å) X-ray spectral
lines of copper. Thus the effect of passing X-rays
from a copper target through a thin nickel foil is
that the Kβ radiation is selectively almost
completely absorbed.
The thickness of the filtering material is usually
chosen to reduce the intensity of the Kβ line by a
factor of 100 while reducing the intensity of the Kα
line by a factor of 10 or less. The optimum
thickness, x of the filter can be determined from
the mass-absorption law:

I/Io = exp{−(μ/ρ)ρx
where (μ/ρ) is the mass absorption coefficient of the
material, ρ is the density of the material, and I and Io
are the transmitted and incident X-ray intensities,
respectively.

The filter CuKα line (λ = n, l, j


L3 2, 1, 3/2
L2 2, 1, 1/2
1.5418 Å) is a close doublet L1 2, 0, 1/2

(Kα1 = 1.5405 Å and Kα2 = K2 K1

1.5443 Å) because K 1, 0, 1/2

transitions can occur from Atomic levels


involved in copper
two possible electronic K emission.
α
configurations (1s1/2 ←
2p1/2, 2p3/2), which differ
slightly in energy. Kα1 is
twice as intense as Kα2. Thus
the value of 1.5418 Å is a
weighted mean (2Kα1 +
Kα2)/3, the weights being
derived from the relative
intensities (2:1) of the Kα1
and Kα2 lines.

X-ray wavelengths of commonly used target


materials
Target Kα1, Å Kα2, Å K ͞α, Å Kβ, Å Filter
Cr 2.2896 2.2935 2.2909 2.0849 V
Fe 1.9360 1.9399 1.9373 1.7567 Mn
Co 1.7889 1.7928 1.7902 1.6208 Fe
Cu 1.5405 1.5443 1.5418 1.3923 Ni
Mo 0.7093 0.7135 0.7107 0.6323 Nb
Ag 0.5594 0.5638 0.5608 0.4971 Pd

K ͞α is the intensity-weighted mean of Kα1 and Kα2.

Properties of X-rays:

(i) X-rays are highly penetrating, invisible rays,


(ii) They have a very short wavelength (about the
same size as the diameter of an atom)
(iii) X-rays are electrically neutral,
(iv) X-rays cannot be deflected by electric field or
magnetic field,
(v) Ionization of a gas results when an X-ray beam is
passed through it,
(vi) Photographic film is blackened by X-rays,
(vii) They are absorbed (stopped) by metal and bone,
(viii) X-rays interact with matter produce
photoelectric and Compton effect.

Diffraction of X-ray by crystals: Diffraction is the


bending of a wave around objects or the spreading
after passing through a gap. Diffraction processes are
most noticeable when the obstruction or gap is about
the same size as the wavelength of the impinging
wave.

Diffracted Diffracted
waves waves
Diffracted waves

Diffraction effects get less


obvious as the gap gets
larger
When a beam of light passes through
two adjacent pinholes, it forms a
pattern of alternating bright and dark
regions - the diffraction pattern, A

occurs whenever distance diffraction


the
between the pinholes is comparable pattern
to the wavelength of the light. The
bright regions appear where light
waves reinforce each other by
arriving in phase, an effect called
constructive interference. The
intervening dark areas occur where
light waves arrive out of phase and
cancel each other (destructive
interference).
Crystals are ordered, three-dimensional arrangements
of atoms with characteristic periodicities. As the
spacing between atoms is on the same order as X-ray
wavelengths (1-3 Å), crystals can diffract the
radiation when the diffracted beams are in-phase.
Note that the X-ray scattering units are the electron
clouds associated with the atoms in the crystal
structure.

Bragg’s equation: In 1913,


British Physicist Sir William
Henry Bragg and his son Sir
Sir W. Sir W.
William Lawrence Bragg
H. Brag L. Bragg
noted the similarity of g
diffraction to ordinary They were
reflection and deduced a awarded the Nobel
simple equation treating
diffraction as reflection from Prize in Physics in
planes in the crystal lattice. 1915.
Figure shows a beam of X-
rays falling on the crystal
surface. Two successive
planes of the crystal are
shown separated by a distance
d. Let the X-rays of
wavelength λ strike the first
plane at an angle θ. Some of
the rays will be reflected at the
same angle. Some of the rays
will penetrate and get
reflected from the second
plane.
Incident rays Reflected rays

Atom or ion
A
 
  d Crystal planes
C D

B

Diffraction of X-rays from crystal planes.

These rays will reinforced those reflected from the


first plane if the extra distance traveled by them (CB
+ BD) is equal to integral number, n, of wavelengths.
That is,
nλ = CB + BD …………………………….(i)
Geometry shows that
CB = BD = ABsinθ……………………....(ii)
From (i) and (ii) it follows that
nλ = 2ABsinθ
nλ = 2dsinθ
This is known as the Bragg’s equation. When Bragg’s
equation is satisfied, the reflected beams are in phase
and interference constructively. At angles of
incidence other than the Bragg angle, reflected beams
are out of phase and destructive interference or
cancellation occurs.

Since the maximum value of the sinθ is unity, then for


a given wavelength of X-rays, there is a lower limit to
the spacing, d, that can give observable diffraction
lines, viz. dmin = λ/2.

Problem: X-rays of wavelength 0.154 nm strike an


aluminium crystal; the rays are reflected at an angle
of 19.3o. Assuming that n = 1, calculate the spacing,
d, between the planes of aluminium atoms in pm that
is responsible for this angle of reflection.
Solution:
Λ
d = --------- [ ∵ n = 1]
2sinθ

1000 pm
0.154 nm × ------------
1 nm
= --------------------------- = 233 pm
2 sin19.3o

Problem: Determine the expected diffraction angle


for the first-order reflection from the (310) set of
planes for BCC chromium when monochromatic
radiation of wavelength 0.0711 nm is used. Given that
atomic radius of Cr is 0.1249 nm.
Solution: For a BCC unit cell,
a = 4R/√3 [R = Radius of a
given atom)
= (4 × 0.1249 nm)/√3
= 0.2884 nm
3a

a
a
2a
a

The distance between the 310 BCC Lattice


planes
d310 = a/√(32 + 12 + 02)
= 0.2884 nm/√10
= 0.0912 nm
From Bragg’s equation,
sinθ = nλ/2d310 = (1 × 0.0711
nm)/(2 × 0.0912 nm) = 0.390

∴ θ = sin-1(0.390) = 22.94o

∴ 2θ = 45.88o
2.1 Perfect and Imperfect Crystals
In a perfect crystal, all the atoms are at rest on their
correct lattice positions. Such a perfect crystal can be
obtained, hypothetically, only at absolute zero; at all
real temperatures, crystals are imperfect. Atoms
vibrate, which may be regarded as a form of defect,
but also a number of atoms are inevitably misplaced.
In some crystals, the number of defects may be very
small, <<1%, as in, e.g., high-purity diamond or
quartz. In others, high defect concentrations may be
present.

Crystals are invariably imperfect


The presence of defects up to a certain concentration
leads to a reduction of free energy, Fig. 2.1. Let us
consider the effect on the free energy of a perfect
crystal of creating a single defect, say a vacant cation
site. This requires a certain amount of energy, H, but
causes a considerable increase in entropy, S,
because of the large number of positions which this
defect can occupy. Thus, if the crystal contains 1 mol
of cations, there are ∼1023 possible positions for the
vacancy. The entropy gained is called configurational
entropy and is given by the Boltzmann equation:
S = k lnW …..(2.1)
where the probability, W, is proportional to 1023;
other, smaller, entropy changes are also present due
to the disturbance of the crystal structure in the
neighborhood of the defect. As a result of this
increase in entropy, the enthalpy required to form the
defect initially is more than offset by the gain in
entropy. Consequently, the free energy, given by
G = H − TS ……. (2.2)
decreases.
If we go now to the other extreme where, say, 10% of
the cation sites are vacant, the change in entropy on
introducing yet more defects is small because the
crystal is already very disordered in terms of occupied

Figure 2.1 Energy changes on introducing defects


into a perfect crystal.
and vacant cation sites. The energy required to create
more defects may be larger than any subsequent gain
in entropy and hence such a high defect concentration
may not be stable. In between these two extremes lie
most real materials. A minimum in free energy exists
which represents the number of defects present under
conditions of thermodynamic equilibrium, Fig. 2.1.

Although this is a simplified explanation, it does


illustrate why crystals are imperfect. It also follows
that the equilibrium number of defects increases with
temperature; assuming that H and S are
independent of temperature, the –TS term becomes
larger and the free energy minimum is displaced to
higher defect concentrations with increase in
temperature.

For a given crystal, curves such as shown in Fig. 2.1


can be drawn for every possible type of defect and the
main difference between them is the position of their
free energy minimum. The defect that predominates
is the one which is easiest to form, i.e. with the
smallest H and for which the free energy minimum
occurs at the highest defect concentration. Thus, for
NaCl it is easiest to form vacancies (the Schottky
defect), whereas in AgCl the reverse is true and
interstitial (Frenkel) defects predominate. In Table
2.1, the defects which predominate in a variety of
inorganic solids are summarised.

2.2 Types of Defect: Point Defects


Various classification schemes have been proposed
for defects, none of which is entirely satisfactory.
Thus defects can be divided into two groups:
stoichiometric defects, in which the crystal
composition is unchanged on introducing the defects,
and non-stoichiometric defects, which are a
consequence of a change in composition.
They are also referred to as intrinsic defects and
extrinsic defects, respectively.
Table 2.1 Predominant point defects in
various ionic crystals
Figure 2.2 2D representation of a Schottky
defect with cation and anion vacancies.
Alternatively, the size and shape of the defect can be
used for classification:
Point defects involve only one atom or site, e.g.
vacancies or interstitials, although the atoms
immediately surrounding the defect are also
somewhat perturbed;
Line defects, i.e. dislocations, are effectively point
defects in two dimensions but in the third dimension
the defect is very extensive or infinite;
Plane defects, whole layers in a crystal structure can
be defective. Sometimes the name extended defects is
used to include all those which are not point defects.
2.2.1 Schottky defect
In ionic solids such as halides or oxides, the Schottky
defect, a stoichiometric defect, is a pair of vacant
sites, an anion vacancy and a cation vacancy. To
compensate for the vacancies, there should be two
extra atoms at the surface of the crystal for each
Schottky defect. The Schottky defect is the principal
point defect in the alkali halides and is shown for
NaCl in Fig. 2.2. There must be equal numbers of
anion and cation vacancies in order to preserve local
electroneutrality.

The vacancies may be distributed at random in the


crystal or may associate into pairs or larger clusters.
They tend to associate because they carry an effective
charge, and oppositely charged vacancies attract each
other. Thus an anion vacancy in NaCl has a net
positive charge of +1 because the vacancy is
surrounded by six Na+ ions, each with partially
unsatisfied positive charge. Put another way, the
anion vacancy has charge +1 because, on placing an
anion of charge –1 in the vacancy, local
electroneutrality is restored. Similarly, the cation
vacancy has a net charge of –1. In order to dissociate
vacancy pairs, energy equivalent to the enthalpy of
association, 1.30 eV for NaCl (∼120 kJmol−1) must be
provided.

The number of Schottky defects in a crystal of NaCl


is either very small or very large, depending on one’s
point of view. At room temperature, typically one in
1015 of the possible anion and cation sites is vacant,
an insignificant number in terms of the average
crystal structure of NaCl as determined by X-ray
diffraction. On the other hand, a grain of salt weighing
1 mg (and containing ∼1019 atoms) contains ∼104
Schottky defects, hardly an insignificant number! The
presence of defects, even in small concentrations,
often influences properties. Thus, Schottky defects
are responsible for the optical and electrical
properties of NaCl.

2.2.2 Frenkel defect


This stoichiometric defect involves an atom displaced
off its lattice site into an interstitial site that is
normally empty. AgCl (which also has the NaCl
crystal structure) has predominantly this defect, with
Ag as the interstitial atom, Fig. 2.3.
Figure 2.3 (a) 2D representation of a Frenkel
defect in AgCl; (b) interstitial site showing
tetrahedral coordination by both Ag and Cl.
The nature of the interstitial site is shown in Fig.
2.3(b). It is surrounded tetrahedrally by four Cl− ions
but also, and at the same distance, by four Ag+ ions.
The interstitial Ag+ ion is in an eight-coordinate site,
and therefore with four Ag+ and four Cl− nearest
neighbors. There is probably some covalent
interaction between the interstitial Ag+ ion and its
four Cl− neighbors which acts to stabilize the defect
and give Frenkel defects, in preference to Schottky
defects, in AgCl. On the other hand, Na+, with its
‘harder’, more cationic character, would not find
much comfort in a site which was tetrahedrally
surrounded by four other Na+ ions. Frenkel defects
therefore do not occur to any significant extent in
NaCl.

Calcium fluoride, CaF2, has predominantly anion


Frenkel defects in which F− occupies interstitial sites.
These interstitial sites (empty cubes) can be seen in
Fig. 1.34. Other materials with fluorite and
antifluorite structures have similar defects, e.g. Na2O
(Na+ interstitial).

As with Schottky defects, the vacancy and interstitial


are oppositely charged and may attract each other to
form a pair. These pairs are electrically neutral but are
dipolar and attract other dipoles to form larger
aggregates or clusters. Clusters similar to these may
act as nuclei for the precipitation of phases of
different composition in non-stoichiometric crystals.

2.2.3 Color centres


The best known example of a colour centre is the F-
centre (from the German Farbenzentre), shown in
Fig. 2.5; it is an electron trapped on an anion vacancy.
F-centres can be prepared by heating an alkali halide
in vapor of an alkali metal. NaCl heated in Na vapor
becomes slightly non-stoichiometric due to the uptake
of Na to give NaCl: δ<< 1, which has a greenish
yellow color. The process must involve the absorption
of Na atoms, which ionize on the crystal surface. The
resulting Na+ ions stay at the surface but the ionized
electrons

Figure 2.5 The F-centre, an electron trapped


on an anion vacancy.

Figure 2.6 (a) H-centre and (b) V-centre in


NaCl.
diffuse into the crystal where they encounter and
occupy vacant anion sites. To preserve charge
balance, an equal number of Cl− ions must find their
way out to the surface. The trapped electron provides
a classic example of an ‘electron in a box’. A series
of energy levels are available for the electron within
this box and the energy required to transfer from one
level to another falls in the visible part of the
electromagnetic spectrum; hence the color of the F-
centre. The magnitude of the energy levels and the
color observed depend on the host crystal and not on
the source of the electron. Thus, NaCl heated in K
vapor has the same yellowish color as NaCl heated in
Na vapor, whereas KCl heated in K vapor is violet.

Another means of producing F-centres is by


irradiation. Using one of the normal methods of
recording an X-ray diffraction (XRD) pattern (see
Chapter 5), powdered NaCl turns a greenish yellow
colour after bombardment with X-rays. The cause of
the colour is again trapped electrons, but in this case
they cannot arise from a non-stoichiometric excess of
Na. They probably arise from ionization of some Cl−
ions.
The F-centre is a single trapped electron which has an
unpaired spin and, therefore, an electron
paramagnetic moment. A powerful method for
studying such color centres is electron spin resonance
(ESR) spectroscopy (see Chapter 6), which detects
unpaired electrons.
Many other color centres have been characterized in
alkali halide crystals; two of these, the H-centre and
V-centre, are shown in Fig. 2.6. Both contain the
chloride molecule ion, Cl2−, but this occupies one site
in the H-centre and two sites in the V-centre; in both
cases the axis of the Cl2− ion is parallel to a <101>
direction.
The V-centre occurs on irradiation of NaCl with X-
rays. The mechanism of formation presumably
involves ionization of a Cl− ion to give a neutral Cl
atom, which then covalently bonds with a
neighboring Cl− ion.

Other defect centres which have been identified in the


alkali halides include:
(a) the F-centre, which is two electrons trapped on an
anion vacancy;
(b) the FA-centre, which is an F-centre, one of whose
six cationic neighbors is a foreign monovalent cation,
e.g. K+ in NaCl;
(c) the M-centre, which is a pair of nearest neighbor
F-centres;
(d) the R-centre, which is three nearest neighbor F-
centres located on a (111) plane;
(e) ionized or charged cluster centres, such as M+, R+
and R−.
Radius ratio rule

Radius ratio is the ratio of cation to the ratio of an


anion. Here, Ratio of cation= r, Ratio of anion = R.
Thus, Radius ratio = (r/R). Limiting radius ratio helps
in expressing the range of radius ratio.

Below Table demonstrates the relationship between


radius ratio (limiting ratio) and coordination number.
Close packing: a close packing is a way of
arranging equidimensional objects in space so
that the available space is filled very efficiently.
Such an arrangement is achieved when each
object is in actual contact with the maximum
number of like objects.

 Tetrahedral voids- In a tetrahedral vacuum, an


atom interacts with four atoms positioned at the
tetrahedron’s four corners. When a sphere of the
second layer is placed over the void then of the
first layer, this void is created. The void’s volume
is significantly less than that of the spherical
particle.
 Octahedral voids- each octahedral void is
generated by the conjunction of triangular voids
from the first and second layers, resulting in an
octahedral void. Octahedral voids are empty
spaces found in substances with an octahedral
crystal structure. It is present in compounds with
a tetrahedral configuration in their crystal
structure.
Octahedral voids are voids enclosed by six
spheres in octahedral configurations.

 It is encompassed by six spheres


 Can be noticed in the unit cells’ centres
 There are six coordination points.
 The number of octahedral voids is equal to n
or half that of tetrahedral voids.
Close-packed structures
The individual atoms in a metallic crystal lattice can
be thought of as hard spheres. The spherical atoms are
packed together in the lattice very efficiently in
geometrical arrangements so as to leave minimum
interspaces having least waste of space. A layer of
uniform spheres can be arranged either as in Fig.
12.26(a) or (b). Clearly the second of the patterns uses
space more efficiently. Here the spheres fit into the
hollows between the adjacent spheres. Thus the
vacant spaces (voids) between the spheres are smaller
than in the first pattern. The metallic crystals are of
the second type i.e., close packing.

As clear from Fig. 12.26(b), each sphere in a closely


packed layer is in contact with four others.
Thus each ball touches six other at the corners of a
hexagon. Three dimensional metallic crystals consist
of closely packed layers stacked one over the other.
The spheres forming the second layer fill the sites or
voids in the first layer and the spheres of the third
layer fill the voids in the second layer.
Depending upon the geometrical arrangements of
spheres in the three layers, the close-packed metallic
crystals are of two types:
(a) Hexagonal close-packed (hcp)
(b) Cubic close-packed (ccp)

Hexagonal Close-Packed Structure


The hexagonal close-packed structure of metallic
crystals is shown in Fig. 12.27. It consists of three
layers of spherical atoms packed one over the another.
The bottom layer (A) and the top layer (A) have three
spheres in similar orientation each having three
atoms. The middle layer (B) consists of six spherical
atoms. The three spheres in the top and the bottom
layer fit into the same voids on either side of the
middle layer.
It is noteworthy that each sphere in the structure is in
contact with 12 neighboring spheres, six in its own
layer, three in the layer above and three in the layer
below. Thus the coordination number of the close-
packed structure is 12. In the overall close-packed
structure, the layers repeat in the manner ABABAB.
The examples of metals having hexagonal close-
packed structures are Ba, Co, Mg and Zn.

Cubic Close-Packed Structure


The cubic close-packed (ccp) pattern of a metallic
crystal is illustrated in Fig. 12.27. Its coordination
number is also 12. Like the hcp structure, it consists
of three layers of spherical atoms lying over one
another. There are three spheres in the top layer (C),
six in the middle layer (B), and three in the bottom
layer (A). However, the overall ccp differs in
structure from the hcp structure in respect of the
orientation of the three spheres in the top layer. In hcp
structure both the top and the bottom layers have the
same orientation. But in ccp structure, they are
oriented in opposite directions.
Therefore, the three spheres in the top layer do not lie
exactly on the spheres in the bottom layers. In ccp
structure, the layers are repeated in the order
ABCABCABC. By turning the whole crystal you can
see that the ccp structure is just the face-centred cubic
structure.
Many metals including Ag, Au, Ca, Co, Cu, Ni,
crystallize in ccp structures.
Body-Centred Cubic Structure
About one-third of the metals pack in a body-centred
cubic structure in which the coordination number is
only 8. Each atom touches four atoms in the layer
above and four atoms in the layer below.
When a square-packed layer (non-close packed) is
packed on another layer (Fig. 12.28), a simple cubic
pattern of spherical atoms results. The large sites
remaining in the middle of each cube on slight
expansion can accommodate another sphere to form a
body-centred cube.
Li, Na and K crystallise in body-centred cubic
structures.

Figure 12.28 (a) Layers of non-close spheres


stacked one on the other. (b) Fifth atom when
inserted in a simple cubic structure forms a
body-centred cubic pattern.

NaCl crystal structure

The sodium chloride structure is composed of Na+ and


Cl- ions.
The number of sodium ions is equal to that of Cl- ions. The
radii of Na+ and Cl- ions 95 pm and 181 pm giving the
radius ratio of 0.524. The radius ratio of 0.524 for NaCl
suggest an octahedral void. Chloride ions (In a typical unit
cell) are arranged in cubic close packing (ccp). In this
arrangement, Cl- ions are present at the corners and at the
centre of each face of the cube. This arrangement is also
regarded as face centred cubic arrangement (fcc).
The sodium ions are present in all the octahedral sites.
Since, the number of octahedral sites in ccp structure is
equal to the number of anions, every octahedral site is
occupied by Na+ ions. So that the formula of sodium
chloride is NaCl i.e. stoichiometry of NaCl is 1:1.
Since there are six octahedral sites around each chloride
ions, each Cl- ion is surrounded by 6 Na+ ions at the corner
of the regular tetrahedron. Similarly each Na+ ion is
surrounded by 6 Cl- ions. Therefore, the coordination
number of Cl- as well as of Na+ ions is six.

This is called 6:6 coordination. It should be noted that


Na+ ions to exactly fit the octahedral sites, the radius ratio
rNa+ / rCl- should be equal to 0.414. However, the actual
radius ratio (rNa+ /rCl- = 0.524) exceeds this value.
Therefore to accommodate large Na+ ions, the Cl- ions
move apart slightly i.e. they do not touch each other and
form an expanded face centred lattice. The number of
NaCl units per unit cell is 4. A number of inter-atomic
distances may be calculated for any material with a NaCl
structure using the lattice parameter, a.
NaCl = a/2
NaNa = ClCl = a/2 = 0.707 a

CsCl crystal structure


The cesium chloride crystal is composed of equal number
of cesium (Cs+) and Chloride Cl- ions. The radii of two ions
(Cs+ = 169 pm and Cl– = 181 pm) led to radius ratio of rCs+
to rCl– as 0.93 which suggest a body centred cubic structure
having a cubic site.
RCs+ to RCl– = 169 / 181 = 0.93
The chloride ion form the simple cubic arrangement and
the cesium ions occupy the cubic interstitial sites. In other
words Cl– ions are at the corners of a cube whereas Cs+ ion
is at the centre of the cube or vice versa.
Each Cs+ ion is surrounded by 8 Cl– ions and each Cl– ion
in surrounded by 8 Cs+ ions. Thus the Coordination
number of each ion is eight.
For exact fitting of Cs+ ions in the cubic voids the ratio rCs+
/ rCl– should be equal to 0.732. However, actually the ratio
is slightly larger (0.93). Therefore packing of Cl– ions
slightly open up to accommodate Cs+ ions.
Number of CsCl units per unit cell is 1
A number of inter-atomic distances may be calculated for
any material with a CsCl structure using the lattice
parameter, a.
CsCl = 3a/2
CsCs = ClCl = a
Other common examples of this type of structure are CsBr,
CsI, TlCl, TlBr
ZnS
The zinc sulphide crystals are composed of equal number
of Zn2+ and S2– ions.
The radii of the two ions (Zn2+ = 74 pm and S2–= 184 pm)
led to the radius (r+/r–) as 0.40 which suggests a tetrahedral
arrangement.
rZn2+/rS2–= 0.40

The Zinc ions are arranged in ccp arrangement, i.e.


sulphide ions are present at the corners and the centres of
each face of the cube.
Zinc ions occupy tetrahedral site. Only half of the
tetrahedral sites are occupied by Zn2+ so that the formula
of the zinc sulphide is ZnS i.e. the stoichiometry of the
compound is 1:1 (Only alternate tetrahedral sites are
occupied by Zn2+)
Since the void is tetrahedral, each zinc ion is surrounded
by four sulphide ions and each sulphide ion is surrounded
tetrahedrally by four zinc ions. Thus zinc sulphide has 4:4
Coordination.
For exact fitting of Zn2+in the tetrahedral sites, formed by
close packing of S2– ions, the ratio Zn2+/ S2– should be
0.225. Actually this ratio is slightly large (0.40). There are
four Zn2+ ions and four S2– ions per unit cell as calculated
below:

No. of S2–ions = 8(at corners) × 1/8 + 6(at face centres) ×


1/2 = 4
No. of Zn2+ ions = 4(within the body) × 1 = 4
Thus, the number of ZnS units per unit cell is equal to 4.
Some more examples of ionic solids having Zinc blende
structures are CuC, CuBr, CuI, AgI, beryllium sulphide.
A number of inter-atomic distances may be calculated for
any material with a ZnS structure using the lattice
parameter, a.
ZnS = 3a/4
ZnZn = SS = a

Electrochemistry
Book: Principle of physical chemistry by Yousuf Ali
Mollah
The capacity of conducting electricity is not the same
for all electrolytes. Strong electrolytes are those
which are good conductors of electricity whereas
those electrolytes which are poor conductors are
called weak electrolytes. The distinction between the
two types is rather vague. Salts, mineral acids and
hydroxides of alkali and alkaline earth metals are
grouped as strong electrolytes. Organic acids,
ammonium hydroxide etc. are weak electrolytes.

The power of conducting electricity by any conductor


is described in terms of its conductivity or
conductance.
The ease with which electrons can pass in a conductor
is measured by conductance.
The tendency of an aqueous solution to convey an
electric current is expressed numerically as
conductivity.
Unit for conductance is Siemens or ohm-1.
Factors affecting Conductance

The attraction between the ions


Solubility
Temperature
Viscosity
Conductance is reciprocal of resistance. If the
electrical resistance of a conductor is measured the
conductance may be calculated as
Conductance = Resistance
Resistance of an electrical conductor may be
measured by using Ohm's law written as:
E (14.4)
R= -
1
where R is the resistance measured in ohms(), I is the
current in ampere and E is the potential difference
(volts) between the two ends.

The reciprocal of specific resistance is known as the


specific conductance, K (kappa). By definition
specific conductance (K)= -- =- .- x-- (14.7)
pR
The unit of specific conductance can be derived as
follows:
specific conductance (K)= --- =_.! = - .X-- ' =
)Jflcin (14.7a) p R a ohm ctn
The specific conductance of a solution is dependent
on its concentration. In order to compare the
conductance of different electrolytic solutions, molar
conductance (A,,1) is used. The molar conductance is
defined as,
"The conductance of all the ions produced when 1
mole of an electrolyte is dissociated into its ions in a
volume V mL."
This is obtained by multiplying specific conductance
(K) by the volume V in niL that contains I mole of the
electrolyte. In other words,
A,= K V (14.8)
where V is the volume of the solution in V L
containing 1 mole of the electrolyte.
Electrical Conductance and Electrolysis 387
The unit of Jim may be derived as follows:
'1m KX V
1!
=—x---x V
R
1 cm cm3
=
ohm cm 2 mol
The unit of urn is then ohm -1

the molar conductivity of a solution at a given


concentration is the conductance of the volume
of solution containing one mole of electrolyte
kept between two electrodes with the unit area of
cross-section and at a distance of unit length. In
general terms, it is defined as the ratio of specific
conductivity and the concentration of the
electrolyte.

Example 14.3: A metal rod of length 3.2 cm and area


of cross-section 0.45 sq. cm inch offers a resistance of
1.8 ohms. An electrolyte solution of concentration 10
% is placed in the cell. Calculate its conductance
specific conductance, molar conductance.

Kohlrausch law
The molar conductance at infinite dilution is different
for different electrolytes and is equal to the sum of the
conductance of the constituent ions of the
electrolytes.
Expression: m = λ+ + λ-
where, λ+ + λ- are called the ion conductance of the
cation and anion, respectively.
For example, the molar conductance of KC1 at
infinite dilution is 149.86 -1cm2mol-1. The ion
conductances of K and Cl are 73.50 and 76.30 Q cm2
moU'. According to Kohl rausch 's law of independent
ionic migration,
A° _O O
,u(KCI) 'UK' + ' cr
= 73.50 + 76.30 Q' cm2 nIol'
= 149.80 Q' cm n1ol1

Example 14.6: The molar conductances at 25°C of


HO, NaCl and CH3COONa are 426.1, 126.4 and
91.00 -1cm2mol-1' respectively at infinite dilution.
Calculate the molar conductance of CH3COOH at
infinite dilution.
Example 14.7: Calculate the molar conductance at
infinite dilution of NH40H, given that the molar
conductance of NH40, NaCI and NaOH at infinite
dilution at 25°C are 149.7, 126.4 and 248.1 -1 ohm
cm2 mol-1respectively.

Ionic mobility
Ionic mobility (μ) is the ability of charged particles to
move through a medium in response to an electric
field that is pulling them at a specified temperature
and pressure. It is the drift speed acquired by the
ions per unit applied electric field or potential
gradient. It is denoted by (μ).
Unit: m2s−1 volt−1.
The ionic mobility, u, is related to ion conductance A
by the expressions
k OC 1 a or2a =kua (14.17)
and A oc ucor, =ku (14.18)
In equations (14.17) and (14.18), 2a and Ac are the
ion conductance of anion and cation respectively and
k is a constant of proportionality. The value of the
constant k has been shown to be equal to one Faraday.
If we consider the ion conductance at infinite
dilutions it can be shown that
+ A = k(u +u) (14.19)
These relations enable one to calculate the ionic
mobility from the values of ion conductance.
Example 14.8: The ion conductance of IF and Cl-
ions, in ohm-1 cm inor', are respectively 350 and 76.3.
Calculate their ionic mobility. (1F= 96500C)

Ionic mobility increases down the group I


 More no. of water molecules are surrounded to
small cation due to which the hydrated radii of
small size cation is large hence the degree of
hydration is greater
 Smaller the size cation, higher the degree of
hydration, lower the ionic mobility
 Small ions give rise to stronger electric fields
than large ones
Ions u/(10−8 Ions u/(10−8
m2 s−1 V−1) m2 s−1 V−1)
OH 20.64 H+ 36.23
F 5.70 Li+ 4.01
Cl 7.91 Na+ 5.19
Br 8.09 K+ 7.62
I 7.96 Rb+ 7.92
Grotthuss mechanism
 Effective motion of a proton that involves the
rearrangement of bonds in a group of water
molecules
 H9O4+ unit, in which the nearly trigonal planar
H3O+ ion is linked to three strongly solvating H2O
molecules

 The hydrogen bonds in the secondary sphere are


weaker than in the primary sphere
 Rapid adjustment of bond lengths and angles in
the remaining cluster, to form an H5O2+ cation of
structure H2O ···H+ ···OH2
 Reorganization has occurred, a new H9O4+
cluster forms as other molecules rotate into a
position
 Position of positive charge is changed

Transport number
The fraction of the current curried by each ionic
species
𝑢+
𝑡+ =
𝑢+ + 𝑢−
𝑢−
𝑡− =
𝑢+ + 𝑢−

 u+ represents the ionic velocity of the cation and


u- that of the anion.
 t+ and t- are the transport number of cation and
anion, respectively

Case study
i) There is no electrolysis
Same number of anions and cations in each
compartment
ii) The anions and cations move with the same speed
 Equal number of the two ions migrate in opposite
directions
 The number of anions and cations ions in the
middle compartment will remain the same
 For electrical neutrality four anions from the
anode and four cations from the cathode must be
discharge.

iii) The anions move at twice the speed of the cations


 u- = 2u+
 Two cations move to the cathode chamber while
four anions move to the anode chamber
 The composition in the middle compartment will
remain constant
Amount lost in anode chamber
𝑡+ =
Total amount loss in anode and cathode chamber

Amount lost in cathode chamber


𝑡− =
Total amount loss in anode and cathode chamber

Methods
 Hittorfs method
 Moving boundary method

 Hittorfs method
 AgNO3 solution of conc. (0.1-0.05 molL-1)
 Connected to a d-c supply source B, through a
variable resistance R, a milliammeter mA and a
copper or silver coulometer
 Current (10-20 mA), time: 2 h
 Ag is deposited on the cathode and dissolved in
the anode
 The conc. of anode and cathode chamber has
changed but in the middle chamber remains
unchanged
 Stop-cocks SS are closed
 Ag content is measured by titrating with
NH4SCN solution with ferric alum as indicator.
Example 14.10: A dilute solution of CuSO4 was
electrolysed using two Pt electrodes. The amount of Cu per
unit mass of the anodic solution was found to be 0.6350
and 0.6236 g after and before electrolysis respectively. The
amount of Ag deposited in silver coulometer in the series
was 0.1351 g. Calculate the transport numbers of Cu2+ and
SO42- ions. (RAM of Cu = 63.5 and Ag = 107.88) (pp406)

 Moving boundary method


 M'A has the anion A common with MA; MA' and
MA have the common cation M
 Densities increase downwards
 P = The initial sharp boundary between the
solutions of M'A and MA
 P' = The final sharp boundary between the
solutions of M'A and MA
 R = The initial sharp boundary between the
solutions of MA' and MA
 R' = The final sharp boundary between the
solutions of MA‘ and MA
𝑢+ 𝑃𝑃′
𝑡+ = =
𝑢+ + 𝑢− 𝑃𝑃′+𝑅𝑅′
𝑢− 𝑅𝑅′
𝑡− = =
𝑢+ + 𝑢− 𝑃𝑃′+𝑅𝑅′
Factor affecting transport number
 Varies with change in concentration
 t+ increases with temperature, t decreases with
temperature. the transport number of cation and
anion comes closer to 0.5
 Nature of electrolyte

Abnormal Transport Numbers


 Transport numbers decrease rapidly with
increase in concentration
 Silver in ammoniacal solution, cadmium in
cadmium iodide
 Attributed to complex formation
2CdI2 = Cd 2+ + [CdI4]2

Spontaneity of cell reaction


𝐺  𝑛𝐹𝐸
G = Change in free energy
n = number of electrons involved in the reaction
E = cell potential
E = Ecathode(reduction)  Eanode(reduction)
Applications of conductance measurements
 The end points of acid-base titrations
 The end points of precipitation titrations
 The solubility of sparingly soluble salts
 The kinetics of reactions
Advantages of conductance titrations
 No indicators are required, colored solution can
be conveniently titrated.
 For weak acid and weak base no suitable
indicators are available but conductometric
titrations can be conveniently carried out.
 More accurate results are obtained by
conductometric titrations because the end-point is
determined graphically from a number of
measurements and not from a single measurement
as in volumetric titration.
 Accurately carried out for very dilute solutions,
where indicator color change is not sharp.
 Volumetric titration does not give correct end
point for polybasic acids, while conductometric
titration can be conveniently used in such cases

The Quinhydrone electrode


𝐶6 𝐻4 𝑂2 (𝑎𝑞) + 2𝑒 − → 𝐶6 𝐻4 𝑂22− (𝑎𝑞)
Quinone Hydroquinone ion
Q H2Q
𝐶6 𝐻4 𝑂22− (𝑎𝑞) + 2𝐻 + (𝑎𝑞) → 𝐶6 𝐻4 𝑂2 (𝑎𝑞)
𝐶6 𝐻4 𝑂2 (𝑎𝑞) → 𝐶6 𝐻4 𝑂2 (𝑎𝑞 ) + 2𝐻 + (𝑎𝑞 ) + 2𝑒 −

You might also like