You are on page 1of 25

Unit I Transition Elements and

Coordination Chemistry
Transition elements
General group trends with special reference
to electronic configuration,
variable valency, colour, magnetic and
catalytic properties, ability to form
complexes and stability of various oxidation
states (Latimer diagram) for
3d-series. 6
Lanthanoides and actinodes : electronic
configurations. Oxidation states,
colour, spectral and magnetic properties,
lanthanide contraction,
separation of lanthanoides (ion exchange
method only). 4
Chemistry of 3d metals:Chemistry
(excluding metallurgy) of chromium,
manganese, iron and cobalt in various
oxidation states and their biological
importance. 11
Coordination chemistry
Valence bond Theory (VBT): Inner and
outer orbital complexes of Cr,
Fe, Co, Ni and Cu (coordination numbers 4
and 6). Structural and
Steroisomerism in complexes with
coordination numbers 4 and 6.
Drawbacks of VBT.
IUPAC system of Nomenclature.
Hard and soft acid-base (HSAB) concept as
applied to complexes. 8
Crystal field theory: Crystal field effect.
Octahedral symmetry. Crystal
field stabilization energy (CFSE). Crystal
field effects for weak and strong
fields. Tetrahedral symmetry. Factors
affecting the magnitude of ?.
Spectrochemical series. Comparison of
CFSE for Oh and Td complexes.
Tetragonal distortion of octahedral
geometry. Jahn-Teller distortion.
Square planar coordination. 10
166
Unit II Organometallic Compounds (6 L)
Organometallic compounds
Definition and classification based on
nature of metal-carbon bond (ionic,
s, p and multicentre bond). Structures of
methyl lithium, Zeiss salt and
ferrocene. EAN rule as applied to
carbonyls. Preparation, structure,
bonding and properties of mononuclear and
polynuclear carbonyls. of 3d
metals. p – acceptor behaviour of carbon
monoxide. Synergic effect (VB
approach). 6
Unit III Solid State & Molecular Symmetry
(11 L)
Solids
Bravais lattices. Identification of lattice
planes. Miller indices. X-ray
diffraction. The Bragg Law. Types of
crystals— molecular, covalent,
metallic and ionic with examples.
Characteristics of these crystals.
Structures of NaCl, CsCl, ZnS & CaCl2.
Point defects in ionic crystals.
Colour centers and dislocations. 6
Molecular Symmetry
Symmetry elements and symmetry
operations: identity (E). rotation about
an axis (Cn), improper rotation (Sn). Plane
of symmetry and center of
symmetry. Assignment of point groups for
simple molecules. Important
information from point groups of
molecules. 5
UNIT IV CHEMICAL KINETICS &
PHOTOCHEMISTRY (14 L)
Chemical Kinetics
Derivation of first and second order rate
equations (both for equal and
unequal concentrations of reactants) Half
life time. . Methods for
determining the order of a reaction
Influence of temperature on reaction
rate. Activation energy and its calculation
from Arrhenius equation.
Lindemann Theory of unimolecular
reactions. Complex reactions such as
consecutive reactions, parallel reactions and
opposing reactions (with
examples) and their differential rate
equations only.
Theories of Reaction Rates: Collision
theory and Activated Complex theory
of bimolecular reactions. Comparison of the
two theories.
167
Introduction to Femtochemistry.
Mechanism and kinetics of enzyme
catalysed reactions— Michaelis-
Menten equation. Effect of temperature on
enzyme catalysis. 10
Photochemistry
Absorption of light. Lambert-Beer law.
Primary and secondary effects of
light absorption, Laws of photochemistry,
Quantum efficiency-Reasons for
low and high quantum yields, Photoelectric
cells. Phosphorescence and
fluorescence. Jablonski diagram.
Chemiluminescence. 4
168
CH 302 ORGANIC AND PHYSICAL
CHEMISTRY
(3 Lectures per week)
Unit I Polynuclear, Hetero-Aromatic
Compounds & Polymers (15 L)
Polynuclear and Hetero-aromatic
compounds
Preparation and properties of the following
compounds and important
derivatives-Naphthalene (including
structure elucidation). Anthracene,
Pyrrole, Furan, Thiophene, Pyridine and
Quinoline. 7
Polymers
Definition and classification. Mechanism of
polymerization (ionic, free
radical and Ziegler-Natta catalyst).
Preparation, properties and uses of
the following polymers-nylons, polyesters,
polyvinyl chloride, teflon,
bakelite, urea and melamine-formaldehyde
resins. Natural rubber
(isolation, structure and vulcanization).
Synthetic elastomers-buna-S, butyl
rubber, polyurethane and foam.
Use of additives in improving
environmental degradability of vinyl
polymers
(polyethylene). Development of bio-
degradable polymers-polylactic acid
and polyhydroxybutyric acid. 8
Unit II Natural Products (20 L)
Amino acids, Peptides and Proteins
Natural and essential amino acids:
Synthesis of simple amino acids by
following methods- Amination of haloacids.
Gabriel’s phthalimide, malonic
ester and Erlenmeyer azlactone synthesis.
Configuration of natural amino
acids and their properties.
Peptides: Primary structure determination
by degradation, N-terminal
(Edman and DNP method), C-terminal
(hydrazinolysis) and hydrolysis of
peptides. Synthesis of simple tripeptides
only. Synthesis of peptides-by
use of N-protecting groups.
(t-butyloxycarbonyl and phthaloyl) C-
activating groups, Merrifield solidphase
synthesis.
Proteins: Importance, primary, secondary,
tertiary and quaternary
structures (definition only). 10
169
Carbohydrates: Definition, classification
and nomenclature. Determination
of configuration of monosaccharides.
Ascending and descending of
monosaccharides series. Interconversion of
aldoses and ketoses,
Structure elucidation of glucose and
fructose (open chain and cyclic),
mutarotation. Structure of sucrose, starch
and cellulose (excluding
structure elucidation).
Alkaloids: Definition and classification.
Structure, synthesis and uses of
Atropine. 10
Unit III Spectroscopy (10 L)
Ultraviolet and Visible: Electromagnetic
radiations, electronic transitions
? max, chromophore, auxochrome, batho
and hypsochromic shifts.
Application of electronic spectroscopy and
Woodward rules for calculating
? max of conjugated dienes and a, ß -
unsaturated carbonyl compounds.
Colour and constitution.
Infra red: IR radiations and types of
molecular vibrations. Functional group
and finger print region. Sampling in IR
spectroscopy, IR spectra of alkanes,
alkenes, alcohols (Inter and Intramolecular
hydrogen bonding), aldehydes,
ketones, carboxylic acids and their
derivatives (effect of substitution
on > C=O stretching absorptions).
Unit IV Quantum Chemistry & Molecular
Spectroscopy (16 L)
Postulates of quantum mechanics.
Discussion of observables and
quantum mechanical operators. Writing the
time independent. Schrödinger
equation for different systems (e.g. particle
in a box, rigid rotator, linear
harmonic oscillator and hydrogen atom).
Interpretation of wave function.
What is spectroscopy? Importance of
spectroscopy. Role of quantum
mechanics in spectroscopy. Difference
between atomic and molecular
spectroscopy. Absorption and emission
spectroscopy. Regions of the
electromagnetic spectrum. Width and
intensity of spectral lines. Born-
Openheimer approximation. Separation of
molecular energies into
translational, rotational, vibrational and
electronic components.
Applications
Translational motion: Schrödinger equation
for a particle in one
dimensional box and its results (solution not
required). Quantisaiton of
170
the translational energy levels; properties of
the solutions; Generalisation
to three dimensions, concept of degeneracy.
Rotational motion: Schrödinger equation of
a rigid diatonic rotator and
its results (solution not required).
Quantisation of rotational energy levels.
Microwave (pure rotational) spectra of
diatomic molecules. Selection rules,
Structural information from rotational
spectra.
Vibrational motion: Schrödinger equation
for a linear harmonic oscillator
and its results (solution not required). Brief
discussion of results.
Quantisation of vibrational energy levels.
Infrared (Vibrational) Spectra of
diatomic molecules. Selection rules.
Structural information from vibrational
spectra.
Raman spectra: A brief introduction.
Rotational Raman Spectra and
Vibrational-Raman spectra. Structural
information from Raman Spectra.
Vibrations of polyatomic molecules. (e.g.
CO2, H2O) (qualitative treatment),
Normal modes of Vibration.
Unit V Surface Chemistry & Polymers (9 L)
Surface Chemistry : Adsorption by solids.
Langmuir theory of adsorption
of a gas on a solid. Langmuir adsorption
isotherm. BET theory of multilayer
adsorption of a gas on a solid. BET
equation (derivation not required).
Calculation of surface area of adsorbent
from the BET equation. Types
of adsorption isotherms. Adsorption
chromatography. 4
Polymers: Different schemes of
classification of polymers. Molar mass
of polymers and its distribution- the
number average and mass average
molar masses. Methods of determining
molar mass-osmotic pressure,
sedimentation, viscosity and light scattering
methods. Introduction to
electrically conducting polymers. 5
171
CH 303 CHEMISTRY LABORATORY - II
Note : Practical examination will include
three exercises — one each out
of the following physical, organic and
inorganic chemistry
experiments.
1. Determination of partition coefficient for
iodine between water and
carbon tetrachloride.
2. Construction of a phase diagram of a
binary system (ureanaphthalene)
by cooling curves method.
3. Study of the kinetics of the hydrolysis of
methyl acetate in presence
of hydrochloric acid using (i) initial rate
method and (ii) integrated
rate method.
4. Analysis of the given organic compounds
containing only one of the
following functional groups:
Carboxylic acids, alcohols, phenols,
aldehydes & ketones,
carbohydrates (monosaccharides), acid
amides, aromatic nitro
compounds, aromatic primary amines.
5. Preparation of
tetraamminecarbanatocobalt (III) nitrate
and
measurement of its conductivity.
6. (a) Preparation of potassium
trioxalatoferrate (III) trihydrate and
measurement of its conductivity.
(b) Estimate the amount of iron present in
the above complex or
in a standard solution as Fe2O3
gravimetrically.
7. Estimation of the amount of nickel
present in a given solution as Bis
(dimethylglyoximato) nickel (II)
gravimetrically.
8. Estimation of (i) Mg2- or (ii) Zn2- by
complexometric titrations using
EDTA.
9. Estimation of total hardness of a given
sample of water by
complexometric titration.
10. Determination of the composition of the
Fe3+ -salicylic acid complex
in solution by Job’s method.
172
Suggested Readings
1. James E. Huheey, Ellen Keiter and
Richard Keiter: Inorganic
Chemistry: Principles of Structure and
Reactivity, Pearson
Publication.
2. G.L. Miessler and Donald A. Tarr:
Inorganic Chemistry, Pearson
Publication.
3. J.D. Lee : A New Concise Inorganic
Chemistry, E.L.B.S.
4. F.A. Cotton & G. Wilkinson: Basic
Inorganic Chemistry, John Wiley
& Sons.
5. I. L. Finar: Organic Chemistry (Vol. I &
II). E.L.B.S.
6. John R. Dyer: Applications of Absorption
Spectroscopy of Organic
Compounds, Prentice Hall.
7. R.M. Silverstein, G.C. Bassler and T.C.
Morrill: Spectroscopic
Identification of Organic Compounds, John
Wiley & Sons.
8. R.T. Morrison & R.N. Boyd: Organic
Chemistry, Prentice Hall.
9. George Odian: Principles of
Polymerization, Wiley-Interscience.
10. Peter Sykes: A Guide Book to
Mechanism in Organic Chemistry,
Orient Longman.
11. P.W. Atkins: Physical Chemistry,
Oxford University Press.
12. G.W. Castellan: Physical Chemistry,
Narosa Publishing House.
13. C.N. Banwell: Fundamentals of
Molecular Spectroscopy. Tata
McGraw Hill.
Additional Reference Books
1. Dougles, McDaniel and Alexader:
Concepts and Models in Inorganic
Chemistry, John Wiley & Sons.
2. A.G. Sharpe: Inorganic Chemistry,
Pearson Education.
3. F.A.Carey and R. J. Sundberg: Advanced
Organic Chemistry,
Plenum Publishers.
4. W.J. Moore, Physical Chemistry.
Prentice-Hall.
5. G.M. Barrow, Physical Chemistry. Tata
McGraw-Hill.
6. Donald A. McQuarrie : Quantum
Chemistry. Oxford University Press.
173
PHYSICS Chemical Kinetics
Integrated Rate Laws
Concepts
The differential rate law describes how the rate of reaction varies with the concentrations of
various species, usually reactants, in the system. The rate of reaction is proportional to the
rates of change in concentrations of the reactants and products; that is, the rate is proportional
to a derivative of a concentration.
To illustrate this point, consider the reaction
A → B
The rate of reaction, r, is given by
d [A]
r=-
dt
Suppose this reaction obeys a first-order rate law:
r = k [A]
This rate law can also be written as
d [A]
r=- = k [A]
dt
This equation is a differential equation that relates the rate of change in a concentration to the
concentration itself. Integration of this equation produces the corresponding integrated rate
law, which relates the concentration to time. When you viewed concentration-time curves in
previous pages, you viewed the integrated rate laws.
d [A]
=-kdt
[A]

At t = 0, the concentration of A is [A]0. The integrated rate law is thus

[A] = [A]0 e- k t
Experimentally one almost always measures how the concentration of a reactant or product
changes as the reaction progresses. In the previous page, you saw how the derivative of the
concentration-time curve can be used to determine the differential rate law for a reaction.
While this approach might work for simple systems with numerically exact data, this method
for determining a rate law does not work well in practice. The experimental data suffers from
random error and frequently one only collects points infrequently. Consequently it is difficult
or impossible to determine the slope accurately.
A much better practical approach is to make characteristic kinetics plots. For each integrated
rate law, there is a characteristic plot that can be created which will produce a straight line.
These characteristic plots are presented in the table shown below; species A is a reactant in
the chemical reaction.
Slope of Units of
Reaction Differential Integrated Rate Characteristic
Kinetic Rate
Order Rate Law Law Kinetic Plot
Plot Constant
d [A]
mole L-1 sec-
Zero - =k [A] = [A]0 - k t [A] vs t -k 1
dt
d [A]
First - = k [A] [A] = [A]0 e- k t ln [A] vs t -k sec-1
dt
d [A] [A]
- = k [A]2 [A] L mole-1 sec-
Second 1/[A] vs t k 1
dt = 1+kt
[A]0

The series of three graphs shown below illustrate the use of the characteristic kinetic plots.
The graph on the left shows [A] vs t plots for a zero-order (red line), first-order (green line),
and second-order (blue line) reaction. The graph in the middle shows ln [A] vs t plots for
each reaction order, and the graph on the right shows 1/[A] vs t plots for each reaction order.

Notice that for each characteristic kinetic plot, a specific rate law shows a straight line. In the
[A] vs t plot, only the zero-order reaction (red line) produces a straight line; the other lines
curve. In the ln [A] vs t plot, only the first-order reaction (green line) produces a straight line,
and in the 1/[A] vs t plot, only the second-order reaction (blue line) produces a straight line.

Experiment
Objectives
• Determine the rate law for a chemical reaction.
• Determine the rate constant for a chemical reaction.
Consider the following reaction between the persulfate ion and the iodide ion:
S2O82- (aq) + 3 I- (aq) → 2 SO42- (aq) + I3- (aq)
In this chemical system, the only species that absorbs visible light is the triiodide ion.
Spectrophotometry can be employed to determine the concentration of triiodide ion, which
with knowledge of the initial concentration of iodide ion permits the concentration of iodide
ion to be determined at any point in time.
In this stopped-flow experiment, one syringe contains a 0.100 M solution of iodide ion and
the other syringe contains a 0.100 M solution of persulfate ion. The two solutions are mixed
in a 3:1 ratio, so that there is a stoichiometric amount of iodide and persulfate ions in the
reaction solution. As the reaction is occurring, the concentration of iodide is plotted versus
time in the left graph.
This reaction is expected to follow a generic rate law of the form
r = k [S2O82-]a [I-]b
This expression indicates that the reaction is of order a with respect to persulfate ion and
order b with respect to iodide ion.
The overall order of the reaction is a + b. Because iodide and persulfate ion are present in a
stoichiometric ratio throughout the reaction,
[I-] = 3 [S2O82-]
Substituting this expression into the general rate expression, one obtains the effective rate law
r = ko [I-]n
where n = a + b is the overall order of the reaction and the observed rate constant ko = k / 3a.
Run the stopped-flow experiment. After the reaction is complete, prepare kinetic plots for the
zero-, first-, and second-order reactions. In the kinetics plots, the data is plotted as red points
and the line-of-best-fit is plotted in blue. Use the kinetics plots to determine the rate law for
this reaction (that is, determine the overall order of the reaction, n). The slope of the kinetic
plot can be used to determine the rate constant ko for the reaction.
Top of Form
Bottom of Form

Top of Form

slope = intercept =
Bottom of Form

Differential Rate Laws Half-Life

Activation energy
From Wikipedia, the free encyclopedia
Jump to: navigation, search

The sparks generated by striking steel against a flint provide the activation energy to initiate
combustion in this Bunsen burner. The blue flame will sustain itself after the sparks are
extinguished because the continued combustion of the flame is now energetically favorable.
In chemistry, activation energy is a term introduced in 1889 by the Swedish scientist Svante
Arrhenius, that is defined as the energy that must be overcome in order for a chemical
reaction to occur. Activation energy may also be defined as the minimum energy required to
start a chemical reaction. The activation energy of a reaction is usually denoted by Ea, and
given in units of kilojoules per mole.
Activation energy can be thought of as the height of the potential barrier (sometimes called
the energy barrier) separating two minima of potential energy (of the reactants and products
of a reaction). For a chemical reaction to proceed at a reasonable rate, there should exist an
appreciable number of molecules with energy equal to or greater than the activation energy.
At a more advanced level, the Arrhenius Activation energy term from the Arrhenius equation
is best regarded as an experimentally determined parameter which indicates the sensitivity of
the reaction rate to temperature. There are two objections to associating this activation energy
with the threshold barrier for an elementary reaction. First, it is often unclear whether or not
reaction does proceed in one step; threshold barriers which are averaged out over all
elementary steps have little theoretical value. Second, even if the reaction being studied is
elementary, a spectrum of individual collisions contributes to rate constants obtained from
bulk ('bulb') experiments involving billions of molecules, with many different reactant
collision geometries and angles, different translational and (possibly) vibrational energies -
all of which may lead to different microscopic reaction rates.

Contents
[hide]
• 1 Negative activation energy
• 2 Temperature independence and the relation to the Arrhenius equation
• 3 Catalysis
• 4 See also
• 5 External links

[edit] Negative activation energy


In some cases rates of reaction decrease with increasing temperature. When following an
approximately exponential relationship so the rate constant can still be fit to an Arrhenius
expression, this results in a negative value of Ea. Elementary reactions exhibiting these
negative activation energies are typically barrierless reactions, in which the reaction
proceeding relies on the capture of the molecules in a potential well. Increasing the
temperature leads to a reduced probability of the colliding molecules capturing one another
(with more glancing collisions not leading to reaction as the higher momentum carries the
colliding particles out of the potential well), expressed as a reaction cross section that
decreases with increasing temperature. Such a situation no longer leads itself to direct
interpretations as the height of a potential barrier.
[edit] Temperature independence and the relation to the
Arrhenius equation
The Arrhenius equation gives the quantitative basis of the relationship between the activation
energy and the rate at which a reaction proceeds. From the Arrhenius equation, the activation
energy can be expressed as

where A is the frequency factor for the reaction, R is the universal gas constant, T is the
temperature (in kelvins), and k is the reaction rate coefficient. While this equation suggests
that the activation energy is dependent on temperature, in regimes in which the Arrhenius
equation is valid this is cancelled by the temperature dependence of k. Thus Ea can be
evaluated from the reaction rate coefficient at any temperature (within the validity of the
Arrhenius equation).
[edit] Catalysis
Main article: Catalysis

The relationship between activation energy (Ea) and enthalpy of formation (ΔH) with and
without a catalyst. The highest energy position (peak position) represents the transition state.
With the catalyst, the energy required to enter transition state decreases, thereby decreasing
the energy required to initiate the reaction.
A substance that modifies the transition state to lower the activation energy is termed a
catalyst; a biological catalyst is termed an enzyme. It is important to note that a catalyst
increases the rate of reaction without being consumed by it. In addition, while the catalyst
lowers the activation energy, it does not change the energies of the original reactants nor
products. Rather, the reactant energy and the product energy remain the same and only the
activation energy is altered (lowered).
[edit] See also
• Arrhenius equation
• Chemical kinetics
• Quantum tunnelling
[edit] External links
• "Activation energy" (from the IUPAC "Gold Book")
• Chapter 14: Activation energy
• The Activation Energy of Chemical Reactions
The Arrhenius equation is a simple, but remarkably accurate, formula for the temperature
dependence of the reaction rate constant, and therefore, rate of a chemical reaction.[1] The
equation was first proposed by the Dutch chemist J. H. van 't Hoff in 1884; five years later in
1889, the Swedish chemist Svante Arrhenius provided a physical justification and
interpretation for it. Nowadays it is best seen as an empirical relationship.[2] It can be used to
model the temperature-variance of diffusion coefficients, population of crystal vacancies,
creep rates, and many other thermally-induced processes/reactions.
A historically useful generalization supported by the Arrhenius equation is that, for many
common chemical reactions at room temperature, the reaction rate doubles for every 10
degree Celsius increase in temperature.

Contents
[hide]
• 1 Overview
• 2 Kinetic theory's interpretation of Arrhenius equation
○ 2.1 Collision theory
○ 2.2 Transition state theory
○ 2.3 Limitations of the idea of Arrhenius Activation Energy
• 3 See also
• 4 Notes and references
• 5 External links

[edit] Overview
In short, the Arrhenius equation gives "the dependence of the rate constant k of chemical
reactions on the temperature T (in absolute temperature, such as kelvins or degrees Rankine)
and activation energy[3] Ea", as shown below:[1]

where A is the pre-exponential factor or simply the prefactor and R is the gas constant. The
units of the pre-exponential factor are identical to those of the rate constant and will vary
depending on the order of the reaction. If the reaction is first order it has the units s−1, and for
that reason it is often called the frequency factor or attempt frequency of the reaction. Most
simply, k is the number of collisions that result in a reaction per second, A is the total number

of collisions (leading to a reaction or not) per second and is the probability that any
given collision will result in a reaction. When the activation energy is given in molecular
units instead of molar units, e.g., joules per molecule instead of joules per mole, the
Boltzmann constant is used instead of the gas constant. It can be seen that either increasing
the temperature or decreasing the activation energy (for example through the use of catalysts)
will result in an increase in rate of reaction.
Given the small temperature range kinetic studies occur in, it is reasonable to approximate the
activation energy as being independent of the temperature. Similarly, under a wide range of
practical conditions, the weak temperature dependence of the pre-exponential factor is
negligible compared to the temperature dependence of the factor; except in the
case of "barrierless" diffusion-limited reactions, in which case the pre-exponential factor is
dominant and is directly observable.
Some authors define a modified Arrhenius equation,[4] that makes explicit the temperature
dependence of the pre-exponential factor. If one allows arbitrary temperature dependence of
the prefactor, the Arrhenius description becomes overcomplete, and the inverse problem (i.e.,
determining the prefactor and activation energy from experimental data) becomes singular.
The modified equation is usually of the form
where T0 is a reference temperature and allows n to be a unitless power. Clearly the original
Arrhenius expression above corresponds to n = 0. Fitted rate constants typically lie in the
range -1<n<1. Theoretical analyses yield various predictions for n. It has been pointed out
that "it is not feasible to establish, on the basis of temperature studies of the rate constant,
whether the predicted T½ dependence of the pre-exponential factor is observed
experimentally."[2] However, if additional evidence is available, from theory and/or from
experiment (such as density dependence), there is no obstacle to incisive tests of the
Arrhenius law.
Another common modification is the stretched exponential form

where β is a unitless number of order 1. This is typically regarded as a fudge factor to make
the model fit the data, but can have theoretical meaning, for example showing the presence of
a range of activation energies or in special cases like the Mott variable range hopping.
Taking the natural logarithm of the Arrhenius equation yields:

So, when a reaction has a rate constant that obeys the Arrhenius equation, a plot of ln(k)
versus T −1 gives a straight line, whose slope and intercept can be used to determine Ea and A.
This procedure has become so common in experimental chemical kinetics that practitioners
have taken to using it to define the activation energy for a reaction. That is the activation
energy is defined to be (-R) times the slope of a plot of ln(k) vs. (1/T):

[edit] Kinetic theory's interpretation of Arrhenius


equation
Arrhenius argued that for reactants to transform into products, they must first acquire a
minimum amount of energy, called the activation energy Ea. At an absolute temperature T,
the fraction of molecules that have a kinetic energy greater than Ea can be calculated from the
Maxwell-Boltzmann distribution of statistical mechanics, and turns out to be proportional to

. The concept of activation energy explains the exponential nature of the relationship,
and in one way or another, it is present in all kinetic theories.
[edit] Collision theory
Main article: Collision theory
One example comes from the "collision theory" of chemical reactions, developed by Max
Trautz and William Lewis in the years 1916-18. In this theory, molecules are supposed to
react if they collide with a relative kinetic energy along their lines-of-center that exceeds Ea.
This leads to an expression very similar to the Arrhenius equation.
[edit] Transition state theory
Another Arrhenius-like expression appears in the "transition state theory" of chemical
reactions, formulated by Wigner, Eyring, Polanyi and Evans in the 1930s. This takes various
forms, but one of the most common is

where ΔG‡ is the Gibbs free energy of activation, kB is Boltzmann's constant, and h is Planck's
constant.
At first sight this looks like an exponential multiplied by a factor that is linear in temperature.
However, one must remember that free energy is itself a temperature dependent quantity. The
free energy of activation is the difference of an enthalpy term and an entropy term multiplied
by the absolute temperature. When all of the details are worked out one ends up with an
expression that again takes the form of an Arrhenius exponential multiplied by a slowly
varying function of T. The precise form of the temperature dependence depends upon the
reaction, and can be calculated using formulas from statistical mechanics involving the
partition functions of the reactants and of the activated complex.

[edit] Limitations of the idea of Arrhenius Activation Energy


Both the Arrhenius activation energy and the rate constant k are experimentally determined,
and represent macroscopic reaction-specific parameters that are not simply related to
threshold energies and the success of individual collisions at the molecular level. Consider a
particular collision (an elementary reaction) between molecules A and B. The collision angle,
the relative translational energy, the internal (particularly vibrational) energy will all
determine the chance that the collision will produce a product molecule AB. Macroscopic
measurements of E and k are the result of many individual collisions with differing collision
parameters. To probe reaction rates at molecular level, experiments have to be conducted
under near-collisional conditions and this subject is often called molecular reaction dynamics
(see Levine).'
[edit] See also
First order reactions
A reaction is said to be first order if its rate is determined by the change of one concentration
term only.
Consider the reaction
A → products
Let α be the concentration of A at the start and after time t, the concentration becomes (a-x),
i.e, x has been changed into products. The rate of reaction after time 't' is given by the
expression
dx/dt = k(a-x)
or dx/((a-x)) = k dt

Upon integration of above equation,


∫ dx/(a-x) = k∫dt
or -loge (a - x) = kt + c
where c is integration constant.
When t = 0 , x = 0,
.·. c = -loge a
Putting the value of 'c',
-loge (a - x) = kt - loge a
or loge a - loge (a - x) = kt
or loge a/(a-x) = kt
or k =2.303/t log10 a/(a-x)

This is known as the kinetic equation for a reaction of the first order. The following two
important conclusions are drawn from this equation:
(i) A change in concentration unit will not change the numerical value of k. let the new unit
So k = 2.303/t log10 na/n(a-x)
or k = 2.303/t log10 a/(a-x)
Thus for first order reaction, any quantity which is proportional to concentration can be used
in place of concentration for evaluation of 'k'.

Second order reactions


A reaction is said to be of second order if its reaction rate is determined by the variation of
two concentration terms.
The kinetics of second order reactions are given as follows:
(i) When concentration of both reactants are equal or two molecules of the same reactant are
involved in the change, i.e.,
A + B → products
or 2A → products
dx/dt = k(a-x)3
On solving this equation,
k = 1/t.x/a(a-x)
where a = initial concentration of the reactant or reactants and
x = concentration of the reactant changed in time t.
(ii) When the initial concentrations of the two reactants are different, i.e.,
A + B → products
Initial conc. a b
dx/dt = k(a-x)(b-x)
k = 2.303/t(a-b) log10 b(a-x)/a(b-x)
(a - x) and (b - x) are the concentrations of A and B after time interval, t.

Characteristics of the second order reactions


(i) The value of k(velocity constant) depends on the unit of concentration. The unit of k is
expressed as (mol/litre)-1 time-1 or litre mol-1 time-1.
(ii) Half life period (t1/2) = 1/k.0.5a/(a×0.5a) = 1/ka
Thus, half life is inversely proportional to initial concentration.
(iii) Second order reaction conforms to the first order when one of the reactants is present in
large excess.
Taking k = 2.303/t(a-b) log10 b(a-x)/a(b-x); if a>>> b then
(a-x) = a and (a-b) = a
Hence,, k = 2.303/ta log10 ba/a(b-x)
or ka = k' = 2.303/t log10 b/((b-x))
(since 'a' being very large, may be treated as constant after the change). Thus the reaction
follows first order kinetics with respect to the reactant taken relatively in small amount.

Examples of second order reactions


1. Hydrolysis of ester by an alkali (saponification).
CH3COOC2H5 + NaOH → CH3COONa + C2H5OH
2. The decomposition of NO2 into NO and O2.
3. Conversion of ozone into oxygen at 100oC
2NO2 → 2NO + O2
4. Thermal decomposition of chlorine monoxide.
2Cl2O → 2Cl2 + O2

You might also like