Sie sind auf Seite 1von 46

0

Catalysis
1

Table of content

Sr.No. Topic Page no.

1 Catalysis 3
1.1 positive catalysis 3
1.2 Negative catalysis 3
1.3 promoters 4
1.4 Inhibitors 4
2 History 6
3 Types of catalysis 8
3.1 What is a phase? 8
3.2 Homogeneous catalysis 9
3.2.1 The destruction of atmospheric ozone 10
3.2.2 Homogenous catalysis in liquid phase 12
3.2.3 Enzymes catalysis 12
3.3 Heterogeneous catalysis 13
4 Autocatalysis 16
5 Characteristic of catalysis 18
6 Theories of catalysis 19
6.1 Intermediate compound formation theory 20
6.1.1 Limitations of intermediate compound formation theory 21
6.2 Adsorption theory 21
7 Applications 23
7.1 Catalysis in Fluid Catalytic Cracking 23
7.1.1 Catalyst Specifications 24
7.1.2 Chemistry involve in catalytic cracking 26
7.2 Hydrogenation 27
7.3 Oxidation 29
7.4 Water-gas shift (WGS) reaction 30
7.4.1 High temperature shift (HTS) catalyst 31
7.4.2 Low temperature shift (LTS) catalyst 32
2

7.5 Production of Carbon Monoxide 33


7.6 Methanol production 34
7.6.1 Steam reforming 36
7.6.2 Methanol decomposition 37
7.6.3 Partial oxidation 37
7.6.4 Combined reforming 37
7.6.5 Ammonia synthesis 37

8 Catalyst characterization techniques 38

8.1 Scanning Electron Microscopy 38

8.2 Physisorption analyses 39

8.3 Chemisorption analyses 40

8.4 Activity measurements 42

8.5 Chemical analyses 45

9 References 46
3

1. Catalysis
An acceleration of the rate of a process or reaction, brought about by a catalyst, usually
present in small managed quantities and unaffected at the end of the reaction. A catalyst
permits reactions or processes to take place more effectively or under milder conditions
than would otherwise be possible. [1]
Unlike other reagents that participate in the chemical reaction, a catalyst is not consumed
by the reaction itself. A catalyst may participate in multiple chemical transformations.
1.1 . Positive catalysis
A catalyst which increase the rate of a reaction is called a positive catalyst and the is
called positive catalysis. There are million of examples of positive catalysis
For example
(1) In the preparation of oxygen MnO 2 at as a positive catalyst in the decomposition of
KCLO3 .

MnO
2KCLO3 2
 2KCL  3O 2

(2) Manufacture of ammonia by Heber’s process using finally divided iron as catalyst

Fe
N 2  3H 2  2NH 3

1.2. Negative catalysis


A catalyst which retards the rate of a reaction is called negative catalysis and the
phenomenon is called negative catalysis. For example
(1) Decomposition of H2O2 is retarded in the presence of traces of acetanilide.

2H 2 O 2 acetanilid
 e 2H 2 O  O 2

(2) Oxidation of chloroform is retarded in the presence of a small quantity of ethyl


alcohol.
4

C H OH
2CHCl3  O 2   2COCl 2  2HCL
2 5

1.3. Promoters (Activators)


Some times the presence of some other substances which are not necessarily catalysts
increases the efficiency of a catalyst. Such substances thus act as catalyst for the catalyst
and are called promoters or activators. Molybdenum act as promoter for the iron catalyst
in the manufacture of ammonia by Haber’s process.

1.4. Inhibitors (Catalytic poisons)


In some cases the presences of small quantities of impurities in the reacting substances
make the catalyst in active. Such substances are called catalytic poisons or inhibitors.
Catalytic poisons are usually also poisonous to catalytic poisons. Arsenious oxide, As 2O3,
and hydrogen oxide poisons platinized asbestos used as catalyst in the manufacture of
H 2 SO4 by contact process.[2]

Catalytic reactions have a lower rate-limiting free energy of activation than the
corresponding uncatalyzed reaction, resulting in higher reaction rate at the same
temperature.
A catalyst can do this by participating in the activated complex for the rate-limiting step,
even though the catalyst itself is neither a reactant nor a product in the overall
stoichiometric equation.
A good example of catalysis is provided by the effect of I2 on the rate of isomerization of
cis-2-butene. The rate law for the catalyzed reaction involves the concentration of I2
raised to the one-half power. This implies that half a molecule of I2 (that is, an I atom) is
involved in the activated complex, which probably has the structure

Since there is no double bond between the two central C atoms, one end of the activated
complex can readily twist around the other.
5

The currently accepted mechanism for this catalyzed reaction involves three steps:

½ I2 I

I + cis-C4H8 → C4H8I‡ → trans-C4H8 + I

I ½ I2

The first and last steps have coefficients of one-half associated with I 2 because each I2
molecule that dissociates produces two I atoms, only one of which is needed to help a
given cis-2-butene molecule to read. Note also that for every I2 molecule which
dissociates in the first step of the mechanism, an I2 molecule is eventually regenerated by
the last step. As a result, the concentration of I2 after the reaction remains exactly the
same as before.
If we consider the energetics of each step in the proposed mechanism, we find that much
less than the 262 kJ mol–1 activation energy of the uncatalyzed reaction is required. A
complete energy profile for the catalyzed process is compared with that of the
uncatalyzed one in Fig.
The bond enthalpy is 151 kJ mol–1 and so 75.5 kJ mol –1 will be required for formation of
the I atom in the first step. An additional increase in energy occurs as the I atom collides
with cis-2-butene and bonds with it. Then about 12 kJ mol –1 is required for twisting
around the C—C single bond in the activated complex. All told 115 kJ mol –1 is required
to go from the initial molecules to the activated complex. When rotation to a trans
structure is complete, the I atom dissociates fromtrans-2-buten and eventually reacts with
another I atom to form I2. These last two processes involve an overall decrease in energy
which is nearly the same as the increase required to achieve the activated complex.

The reduction in activation energy illustrated in Fig


6

Fig. Energy profile for the cis-trans conversion of 2-butene catalyzed by iodine. Iodine
atoms must first be produced from I 2 molecules after which the addition of an I atom to a
butene molecule produces the activated complex. Since the activated complex contains
no double bond, one part of the molecule can now swivel easily around the other,
producing the trans isomer after the loss of the I atom. Since this pathway requires much
lower activation energy, it allows the reaction to occur much more rapidly. [3]

Catalysts may affect the reaction environment favorably, or bind to the reagents to
polarize bonds, e.g. acid catalysts for reactions of carbonyl compounds, or form specific
intermediates that are not produced naturally, such as osmate esters in osmium tetroxide-
catalyzed dihydroxylation of alkenes, or cause lysis of reagents to reactive forms, such as
atomic hydrogen in catalytic hydrogenation.
2. History
This term derived from Greek καταλύειν, meaning "to annul," or "to untie," or "to pick
up." The word "catalysis" (Katalyse) was coined by Berzelius in 1835:
"Catalysts are substance that increase the rate of reaction but remain unchanged after the
reaction."
7

The Chinese Tsoo Mei is more picturesque; it also means "the marriage broker," and so
implies a theory of catalytic action. The idea of catalysis extends far back in to chemical
history. The quest of the alchemist for the philosopher's stone seems like the search of the
modern petroleum chemist for the magical catalyst that will convert crude petroleum into
high octane fuel. In a fourteenth-century Arabian manuscript, Al Alfani described the
"Xerion, aliksir, noble stone, magisterium, that heals the sick, and turns base metals into
gold, without in itself undergoing the least change"
The earliest consciously used catalysts Were the ferments or enzymes. "In the seeding
and growth of plants, and in the diverse changes of the fluids of the animal body in
sickness and in health, a fermentative action takes place ".
Its action has been likened to that of a coin inserted in a slot machine that yields valuable
products and also returns the coin. In a chemical reaction the catalyst enters at one stage
and leaves at another. The essence of catalysis is not the entering but the falling out.
Other early chemists involved in catalysis were Alexander Mitscherlich who
referred to contact processes and Johann Wolfgang Döbereiner who spoke of contact
action and whose lighter based on hydrogen and a platinum sponge became a huge
commercial success in the 1820s. Humphry Davy discovered the use of platinum in
catalysis. In the 1880s, Wilhelm Ostwald at Leipzig University started a systematic
investigation into reactions that were catalyzed by the presence of acids and bases, and
found that chemical reactions occur at finite rates and that these rates can be used to
determine the strengths of acids and bases. For this work, Ostwald was awarded the 1909
Nobel Prize in Chemistry.
Wilhelm Ostwald was the first to emphasize that the catalyst influences the rate of a
chemical reaction but has no effect on the position of equilibrium. His famous definition
was:
"A catalyst is a substance that changes the velocity of a chemical reaction without itself
appearing in the end products."
Ostwald showed that a catalyst cannot change the equilibrium position, by a simple
argument based on the First Law of Thermodynamics. Consider a gas reaction that
proceeds with a change in volume. The gas is confined in a cylinder fitted with a piston;
8

the catalyst is in a small receptacle with in the cylinder, and can be alternately exposed
and covered. If the equilibrium position were altered by exposing the catalyst, the volume
Would change, the piston would move up and down, and a perpetual-motion
Machine would be available.
Since a catalyst can change the rate but not the equilibrium, it follows that a catalyst
Must accelerate the forward and reverse reactions in the same proportion, since
k  k f kb . Thus catalysts that accelerate the hydrolysis of esters must also accelerate the

esterification of alcohols; dehydrogenation catalysts like nickel and platinum are also
good hydrogenation catalysts; enzymes like pepsin and papain that catalyze the splitting
of peptides must also catalyze their synthesis from the amino acids.
A distinction is generally made between homogeneous catalysis, the entire reaction
occurring in a single phase, and heterogeneous catalysis at phase interfaces. The latter is
[8]
also called contactor surface catalysis.

3. Types of catalysis:
There are two major classes of catalysis
Homogeneous
Heterogeneous
Before discussing homogeneous and heterogeneous catalysis first we discuss phase.

3.1. What is a phase?


If we look at a mixture and can see a boundary between two of the components, those
substances are in different phases. A mixture containing a solid and a liquid consists of
two phases. A mixture of various chemicals in a single solution consists of only one
phase, because we can't see any boundary between them.
9

we might wonder why phase differs from the term physical state (solid, liquid or gas). It
includes solids, liquids and gases, but is actually a bit more general. It can also apply to
two liquids (oil and water, for example) which don't dissolve in each other. we could see
the boundary between the two liquids.

If we want to be fussy about things, the diagrams actually show more phases than are
labelled. Each, for example, also has the glass beaker as a solid phase. All probably have
a gas above the liquid - that's another phase. We don't count these extra phases because
they aren't a part of the reaction. [4]
3.2. Homogeneous catalysis
homo implies the same (as in homosexual). Homogeneous catalysis has the catalyst in the
same phase as the reactants.
Typically everything will be present as a gas or contained in a single liquid phase.[9]
10

Examples
The reaction between ethanol and ethanoic acid is very slow even when heat is applied.
C2H5OH + CH3COOH CH3COOC2H5 + H2O
The addition of a few mL of sulfuric acid greatly increases the rate of reaction.
Since ethanol, ethanoic acid and sulfuric acid are all liquids, sulfuric acid, as used in this
example, is a homogeneous catalyst. [10]
An example of homogeneous catalysis in the gas phase is the effect of iodine vapor
On the decomposition of aldehydes and ethers. The addition of a few percent of iodine
often increases the rate of pyrolysis several hundredfold. The reaction velocity follows
the equation,
 d (ether )  k2 ( I 2 )(ether )
dt
Dependence of the rate on catalyst concentration is characteristic of homogeneous
catalysis. The catalyst acts by providing a path for the decomposition that has
considerably lower activation energy than the uncatalyzed path. In this instance the
Uncatalyzed pyrolysis has an Ea  53 kcal, where as with added iodine the E a drops
to34 kcal.
3.2.1. The destruction of atmospheric ozone
This is a good example of homogeneous catalysis where everything is present as a gas.
Ozone, O3, is constantly being formed and broken up again in the high atmosphere by the
action of ultraviolet light. Ordinary oxygen molecules absorb ultraviolet light and break
into individual oxygen atoms. These have unpaired electrons, and are known as free
radicals. They are very reactive.

The oxygen radicals can then combine with ordinary oxygen molecules to make ozone.
11

Ozone can also be split up again into ordinary oxygen and an oxygen radical by
absorbing ultraviolet light.

This formation and breaking up of ozone is going on all the time. Taken together, these
reactions stop a lot of harmful ultraviolet radiation penetrating the atmosphere to reach
the surface of the Earth.
The catalytic reaction we are interested in destroys the ozone and so stops it absorbing
UV in this way.
Chlorofluorocarbons (CFCs) like CF2Cl2, for example, were used extensively in aerosols
and as refrigerants. Their slow breakdown in the atmosphere produces chlorine atoms -
chlorine free radicals. These catalyse the destruction of the ozone.
This happens in two stages. In the first, the ozone is broken up and a new free radical is
produced.

The chlorine radical catalyst is regenerated by a second reaction. This can happen in two
ways depending on whether the ClO radical hits an ozone molecule or an oxygen radical.
If it hits an oxygen radical (produced from one of the reactions we've looked at
previously):

Or if it hits an ozone molecule:

Because the chlorine radical keeps on being regenerated, each one can destroy thousands
of ozone molecules.[5]
12

3.2.2. Homogenous catalysis in liquid phase


Important examples of homogenous catalysis in liquid phase are acid base catalysysis.
The common acid catalyst in water is the hydronium ion and the most common base
catalyst is the hydroxyl ion. If a acid catalyses a reaction the reaction is said to be the
subject of acid catalysis. Inversion of cane sugar and hydrolysis of esters are some
common examples of acid catalysed reaction. However it was shown that different acids
have different catalytic activity; hydrochloric acid has greater activity than acetic acid.so
it is evident that the actual catalysts are H+ ions. The rate of reaction is found to be
proportional to the concentration of H+ ions and the concentration of the reacting
molecules or ions.

C12 H 22O11  H 2O 
H
C6 H 12O6  C6 H 12O6

H
CH 3COOC 2 H 5  H 2O  CH 3COOH  C2 H 5OH

Such reaction which are catalysed by certain acids are said to be specific acid catalysis.
Similarly there are reactions which are catalysed by OH - ions only and are said to be
specific hydroxyl ion catalysis. Conversion of acetone in to diaacetonyl alcohol is
example of hydroxyl ion catalysis.

CH 3COOCH3  CH 3COOCH 3 OH
 Chs CCH 2C (CH 3) 2 OH

3.2.3. Enzymes catalysis


Enzymes are homogeneous biological catalysts. They are complex protein molecules with
three dimensional structures. These are responsible for catalyzing the chemical reaction
in living organisms. The diameter of the enzyme molecules fall in the range of 10-100
nm. Enzymes are often present in colloidal states and are extremely specific in their
catalytic function. Various enzyme catalyzed reactions are known. Some important
examples are
(1) Urease, an enzymes that catalyses the hydrolysis of urea but has no effect on the
hydrolysis of substituted urea, e.g. methyl urea.
NH 2CONH 2  H 2O urease
 2 NH 3  CO2

(2) peptide, glycyl-L-glutamy1-L-tyrosine, is hydrolysed by an enzyme known as


pepsin.
13

(3) Hydrolysis of starch in to maltose by diastase.


2(C6 H10O5 ) n  nH 2O diastase
 nC12 H 22O11
Starch Maltose

(4) Conversion of glucose in to ethanol by zymase present in the yeast


C6 H12O6  H 2O zymase
 2C2 H 5OH  2CO2

(5) Conversion of maltose in to glucose by maltase


C12 H 22O11  H 2O maltase
 2C6 H12O6

(6) Oxidation of alcohol to acetic acid by micoderma aceti


Micoderma
C2 H5OH  O2 CH3COOH  H 2O
aceti

Almost all enzymes fall in to two classes, the hydrolytic enzymes and the oxidation
reduction enzymes. The hydrolytic enzymes appear to be complex acid base catalysis
which accelerates the ionic reaction mainly due to the transfer of hydrogen ions. The
oxidation reduction enzymes catalyze electron transfer perhaps through the formation
of an intermediate radical. [6]

3.3. Heterogeneous catalysis:


The catalyst is in a different phase from the reactants. . Usually the catalyst is a solid and
the reactants are gases, and so the rate-limiting step occurs at the solid surface. Thus
heterogeneous catalysis is also referred to as surface catalysis.
A solid catalyst adsorbs reacting species onto its surface where the reaction takes place

2SO2(g) + O2(g) 2SO3(g)

Solid V2O5 is the catalyst for this reaction.


14

Other example of heterogeneous catalysis is hydrogenation of an unsaturated organic


compound such as ethane (C2H4) by metal catalysts such as Pt or Ni:

The currently accepted mechanism for this reaction involves weak bonding of both H2
and C2H4 to atoms on the metal surface. This is called adsorption.
Ethene molecules are adsorbed on the surface of the nickel. The double bond between the
carbon atoms breaks and the electrons are used to bond it to the nickel surface.

Hydrogen molecules are also adsorbed on to the surface of the nickel. When this happens,
the hydrogen molecules are broken into atoms. These can move around on the surface of
the nickel.

If a hydrogen atom diffuses close to one of the bonded carbons, the bond between the
carbon and the nickel is replaced by one between the carbon and hydrogen.
15

That end of the original ethene now breaks free of the surface, and eventually the same
thing will happen at the other end.

As before, one of the hydrogen atoms forms a bond with the carbon, and that end also
breaks free. There is now space on the surface of the nickel for new reactant molecules to
go through the whole process again.

[11]

These adsorbed H atoms can move across the metal surface, and eventually they combine
with a C2H4 molecule, completing the reaction. Because adsorption and dissociation of H2
on a Pt surface is exothermic (ΔHm° >= –160 kJ mol–1), it can provide H atoms for further
reaction without a large activation energy. By contrast, dissociation of gaseous H 2
16

molecules without a metal surface would require the full bond enthalpy (ΔHm° = +436 kJ
mol–1). Clearly the metal surface makes a major contribution in lowering the activation
energy. [5]
4. Autocatalysis
This is the case of the generation of catalyst during the course of reaction. When one of
the products of a reaction itself act as a catalyst for that reaction, the phenomenon is
called autocatalysis. For example, oxidation of oxalic acid by KMnO 4 in the presence of
sulphuric acid is initially slow, but the rate increase with progress of reaction,
since managanous sulphate produce during the reaction acts as a catalyst for the reaction.

2 KMnO4  5 H 2C2O4  3H 2 SO4  2MnSO4  K 2 SO4  8H 2O  10CO2

Another example is the hydrolysis of ester by water where the product of hydrolysis the
acid is a catalyst in the reaction.

CH 3COOC 2 H 5  H 2O  CH 3COOH  C2 H 5OH

Heat, light, electricity or energy may alter the rate of reaction, but they are not catalyst. A
catalyst must be a material particle.
A catalyst takes part in the formation of the activated complex and is regenerated at the
end. Thus a catalyst is both the reactant and a product of a reaction, a catalyst decreases
the activation energy, by taking part in the activated complex formation and thus
increases the rate.

We can measure this effect by plotting the concentration of one of the reactants as time
goes on. We can get a graph quite unlike the normal rate curve for a reaction.[6]

Most reactions give a rate curve which looks like this


17

Concentrations are high at the beginning and so the reaction is fast - shown by a rapid fall
in the reactant concentration. As things get used up, the reaction slows down and
eventually stops as one or more of the reactants are completely used up.

An example of autocatalysis gives a curve like this:

Note
18

Don't assume that a rate curve which looks like this necessarily shows an example of
autocatalysis. There are other effects which might produce a similar graph.

For example, if the reaction involved a solid reacting with a liquid, there might be some
sort of surface coating on the solid which the liquid has to penetrate before the expected
reaction can happen.

A more common possibility is that we have a strongly exothermic reaction and aren't
controlling the temperature properly. The heat evolved during the reaction speeds the
reaction up. [7]

5. Characteristic of catalysis

 A catalyst remains unchanged in mass and chemical composition at the end of the
reaction. However, its physical forms may be altered. Thus granular MnO 2 used as
a catalyst in the thermal decomposition of KClO3 is left as a fine power at the end
of the reaction.
 A small amount of catalyst is sufficient to bring about a considerable extent of
reaction. Hence a small amount may cause large quantities of materials to react.
However this is not an essential criteria and also cannot be approved. The rate of
catalytic reaction always dependent on the concentration of the catalyst when the
amount of catalyst is small.
 The rate of catalytic reaction depends upon the concentration of the catalyst. This
is true when the concentration of catalyst is small, or concentration of reactants
and catalyst are comparable.
 A catalyst is more effective when finely divided. Thus a lump of platinum will
have much less catalytic activity than colloidal platinum. Finally divided Ni is
better catalyst than lumps of solid nickel.
 A catalyst cannot affect the position of equilibrium, although it shortens the time
required to establish equilibrium. However, a homogeneous catalyst can change
19

the equilibrium composition, but the effect is usually very small since a catalyst is
generally present in small amount.
 A catalyst is in its action and thus it eliminate undesirable side product. When a
several product are possible from same reactant, a catalyst can lower the
activation energy of specific path. Thus it is able to select and direct the course of
the reaction in a specific path. This property is specificity or selectivity. This is a
property of changing the direction.
 A catalyst cannot start a reaction, which is not thermodynamically feasible.
 A catalyst is universal in its use. If it catalyses a reaction, once, it will catalyse
that reaction always. What is more, catalysts are general. For example, silver
oxide is generally used as a catalyst for hydration, and alumina and zircon is used
for the dehydration.
 Change of temperature alters the rate of catalytic reaction as it would do for the
same reaction without a catalyst. Some catalysts are, however, physically altered
by rise in temperature and hence their catalytic activity may be decreased. Thus as
obvious from all the criteria for all purposes a catalyst can
(i) increase the velocity of a reaction
(ii) direct the reaction toward a specific product and
(iii) Eliminate undivisable side product.
In short catalysis is the process of giving higher yield of purer product in shorter
reaction time.
6. Theories of catalysis
A stated earlier, the essential requirement for a reaction to occur is that the
reacting molecules must acquire sufficient energy. In case of catalyst is added to the
reaction, the energy required to activate the molecules is less than in the absence of a
catalyst. Dye to the lower activation energy more molecules will take part in the reaction
and hence the of catalyzed reaction would increase.the action of a catalyst can be
explained by two different mechanism
(1) Intermediate compound formation theory
(2) Adsorption theory
6.1. Intermediate compound formation theory
20

In this theory essentially two steps are involved.


(i) Combination of the catalyst with one or more of the reactants forming
intermediate compound; and
(ii) Decomposition of the intermediate compound or its combination with other
reactant yielding the product and the catalyst back. Consider a reaction
between reactant A and B giving the product AB.

A + B AB
This reaction is very slow and is catalysed by the presence of a catalyst X. the
reaction will therefore proceed as

A + X AX ( intermediate compound)

AX + B AB + X
The formation of an intermediate compound AX is an easy reaction and needs low
energy of activation thereby accelerating the rate of chemical reaction.
Examples
 In lead chamber process for the manufacturing of sulphuric acid, the catalyst
NO first forms the intermediate compound with oxygen.

2 NO  1 / 2O2  2 NO2

NO2  SO2  SO3  NO

 In the preparation of diethyl ether from ethanol using concentrated H 2 SO4


C2 H 5 HSO4 is the first formed as an intermediate.

C2 H 5OH  H 2 SO4  C2 H 5 HSO4  H 2O

C2 H 5 HSO4  C2 H 5OH  C2 H 5OC2 H 5  H 2 SO4

 The formation of water by combination of hydrogen and oxygen in the


presence of copper as a catalyst is as follows.
21

2Cu  1 / 2O2  Cu 2O

6.1.1. Limitations of intermediate compound


Cu 2O  H 2  H 2O  2Cu
formation theory
This theory does not explain the cases of heterogeneous catalysis in general and
more specifically the deactivation by a catalytic poison and activation by a
promoter.
6.2. Adsorption theory
A large number of gaseous reactions take place in the presence of solid catalysts.
The surface of the catalyst has certain active centers due to the unsaturation of
valences. Appreciable quantities of a reactant molecules are adsorbed or retained
by solid surface at these active centers and the reaction occur at the surface of a
solid. For this reason this type of catalysis is sometime referred to as the contact
catalysis. The adsorb molecules from some sort of an active complex on the
surface, which then decompose forming the products. The products are ultimately
desorbed from the surface. A catalytic reaction involves the following steps
(i) Diffusion of the reactant from the bulk on to the surface.
(ii) Adsorption of the reactant on the surface of the catalyst.
(iii) Activation of the adsorbed reactants leading to a reaction in the adsorbed
phase.
(iv) Description of the product from the surface of the catalyst.
(v) Diffusion of the product away from the surface of the catalyst.
Any one of these steps may be slowest and consequently the rate determining but
generally step (iii) is the rate controlling step,
Due to the adsorption the concentration of the reactant tend to increase on the surface
of the catalyst and according to the law of mass action the rate of the reaction will
increase. Further more adsorption being an exothermic process the heat evolved
during adsorption is utilized in the activation of the surface reaction. Adsorption may
also lead to proper orientation of the reacting molecules, partial loosening of the
bonds in the finally divided state and thus requiring only small energy to form the
activated complex.
22

Adsorption theory can explain the enhanced catalytic action of a catalyst in the finely
divided state. It is due to the larger surface area available for adsorption and also the
formation of more active centers.

It has observed that chemisorption plays an important role in the surface catalysis.
The stronger forces are operative in the chemisorptions which tend to bring about a
partial loosening of the bonds in the adsorbed reacting molecules and very small
energy is required to form the activated complex for the reaction.
It has been found that there are some active centers on the surface of the catalyst
where the adsorption take place, X-ray studies of catalysts show that there are
unsatisfied forces present on the surface of the catalyst which can be used to attach
molecules. The surface of the nickel catalyst may be represented as follow.

Isolated nickel atoms and discontinuities probably act as active centers. This explains
the enhanced catalytic activity of finally divided catalysts.[2,6]

7. Applications
23

The rate of reaction for some chemical process may be sped up in one of two ways.
Either temperature can be increased or the mechanism can be altered so as to lower the
activation energy of the reaction. The latter method may be can be done with the
implementation of a catalyst. A catalyst is an intermediate in a chemical reaction that
participates in the activated complex for the rate limiting step. There are three main
features of catalysts:
1. The catalyst allows a reaction to proceed via an alternative mechanism.
2. For every step in the mechanism in which the catalyst appears as a product, there is
another step in which the catalyst is a reactant.
3. The catalyst increases the rate of reaction by lowering the activation energy of the
reaction.

7.1. Catalysis in Fluid Catalytic Cracking


One very important example of the use of catalysis is fluid catalytic cracking. This
process has been fundamental in the last few decades in meeting the high demands for
high octane liquid fuels, such as gasoline.
In fact, without FCC, methods of transportation such as cars and buses in their
present forms would not be sustainable due to such high costs in fuel. From this fact
alone, it is obvious that FCC is an essential process that has an important impact on many
areas, including industry, transportation, and multiple aspects of everyday life.
When petroleum is collected from the stores in the earth, its raw form contains many
different hydrocarbon compounds. These hydrocarbons can contain anywhere from five
to forty carbon atoms. Historically, this small fact has caused a bit of a dilemma for the
energy industry. Hydrocarbons with less carbons in the chain are much more valuable
than the long hydrocarbons with carbon chains of twenty or more.
24

These more compact hydrocarbons are especially valuable in a culture highly


dependent on cars buses, jet planes, and other such aspects of life that require a compact
and potent source of fuel.Due to this imbalance in demand and the excess of heavier
hydrocarbons, the process of fluid catalytic cracking was implemented.

Generally speaking, a longer carbon chain will have a significantly higher boiling point.
This basis forms the means by which the various hydrocarbons in petroleum can be
distinguished and separated. As the goal of FCC is to break down the larger hydrocarbons
into smaller, more valuable compounds, the portion of the crude oil that is used in FCC
(the feedstock) is all those hydrocarbons with an initial boiling point of roughly 340
degrees Celsius or higher.
In an equivalent sense, the feedstock contains all hydrocarbons with a molecular
weight ranging from roughly 200 to 600 grams per mole. In fluid catalytic cracking, these
hydrocarbons are vaporized and broken at high temperatures in the presence of a special
catalyst. The purpose of the catalyst is to make the process more efficient and thus
economically viable.
7.1.1. Catalyst Specifications
The most efficient form for the FCC process is highly desired and has been avidly
searched for. One factor that effects the overall efficiency very significantly is the
substance that is used as the catalyst. After many years of trial and error, a set of criteria
25

for an efficient FCC catalyst has been established. The best form is a fine powder with a
density ranging from 0.80 to 0.96 g/ml and an average particle size of 60-100 microns.
The catalyst should be highly reactive, stable at high temperatures, and retain large pore
sizes. All FCC catalysts contain a crystalline zeolite. In essence, the zeolite component
acts as a molecular sieve that only allows a certain size range of hydrocarbons to enter its
lattice.

Fig. A schematic flow diagram of a Fluid Catalytic Cracking unit as used in


petroleum refineries

The catalytic activity sites are provided by the matrix component of the catalyst,
which contains amorphous alumina that are capable of cracking the larger feedstock
molecules. Other components are present in order to ensure that the catalyst maintains
strength and stability. Due to the nature of the catalyst components, it is highly important
that the catalyst is not introduced to feedstock with metal contaminants. Even
26

concentrations in the range of a few ppm of these contaminants can have detrimental
effects on the performance of the FCC catalyst. Once an efficient catalyst is found, it is
used extensively as a main ingredient. For every one kilogram of feedstock in the
processing unit, there are five kilograms of catalyst compound added. Thus, a comparable
processing unit might circulate 55,900 metric tons of catalyst each day.

Fluid catalytic cracking is a form of applied catalysis that has become essential to the
energy industry. Its ability to recycle large hydrocarbons present in petroleum into more
valuable and smaller hydrocarbons has allowed for demand imbalances to be reconciled.
During the year 2007, in the United States alone, over 5,300,000 barrels of crude oil were
processed using FCC. This astounding data proves the importance of fluid catalytic
cracking to industry, and the entire process could not be implemented without a well
devised incorporation of catalysis. [12]
7.1.2. Chemistry involve in catalytic cracking
Olefins or alkenes, which are unsaturated straight-chain or branched hydrocarbons, do
not occur naturally in crude oil.

Fig: Diagrammatic example of the catalytic cracking of petroleum hydrocarbons


27

In plain language, the fluid catalytic cracking process breaks large hydrocarbon
molecules into smaller molecules by contacting them with powdered catalyst at a high
temperature and moderate pressure which first vaporizes the hydrocarbons and then
breaks them. The cracking reactions occur in the vapor phase and start immediately when
the feedstock is vaporized in the catalyst riser.
Figure is a very simplified schematic diagram that exemplifies how the process
breaks high boiling, straight-chain alkane (paraffin) hydrocarbons into smaller straight-
chain alkanes as well as branched-chain alkanes, branched alkenes (olefins) and
cycloalkanes (naphthenes). The breaking of the large hydrocarbon molecules into smaller
molecules is more technically referred to by organic chemists as scission of the carbon-
to-carbon bonds.
As above in Figure, some of the smaller alkanes are then broken and converted into
even smaller alkenes and branched alkenes such as the gases ethylene, propylene,
butylenes, and isobutylenes. Those olefinic gases are valuable for use as petrochemical
feedstocks. The propylene, butylene and isobutylene are also valuable feedstocks for
certain petroleum refining processes that convert them into high-octane gasoline blending
components.
As also depicted in Figure, the cycloalkanes (naphthenes) formed by the initial
breakup of the large molecules are further converted to aromatics such as benzene,
toluene, and xylenes, which boil in the gasoline boiling range and have much higher
octane ratings than alkanes.
By no means does Figure include all the chemistry of the primary and secondary
reactions taking place in the fluid catalytic process. There are a great many other
reactions involved.
7.2. Hydrogenation
An example of catalytic action is found in the hydrogenation of alkenes. The alkene
(2) adsorbs by forming two bonds with the surface ( 3), and on the same surface there
may be adsorbed H atoms. When an encounter occurs, one of the alkene–surface bonds
is broken (forming 4 or 5) and later an encounter with a second H atom releases
the fully hydrogenated hydrocarbon, which is the thermodynamically more stable
species. The evidence for a two-stage reaction is the appearance of different isomeric
28

alkenes in the mixture. The formation of isomers comes about because, while the
hydrocarbon chain is waving about over the surface of the metal, an atom in the chain
might chemisorb again to form (6) and then desorb to ( 7), an isomer of the original
molecule. The new alkene would not be formed if the two hydrogen atoms attached
simultaneously.

A major industrial application of


catalytic hydrogenation is to the
formation of
edible fats from vegetable and animal
oils. Raw oils obtained from sources
such as the
Soya bean have the structure CH 2
(OOCR)CH(OOCR′) CH 2 (OOCR″ ),
where R, R ′, and R ″ are long-chain
hydrocarbons with several double bonds.
One disadvantage of the presence of
many double bonds is that the oils are
susceptible to atmospheric oxidation,
and therefore are liable to become
rancid. The geometrical configuration of
29

the chains is responsible for the liquid nature of the oil, and in many applications a solid
fat is at least much better and often necessary. Controlled partial hydrogenation of an oil
with a catalyst carefully selected so that hydrogenation is incomplete and so that the
chains do not isomerize (finely divided nickel, in fact), is used on a wide scale to produce
edible fats. The process, and the industry, is not made any easier by the seasonal variation
of the number of double bonds in the oils. [13]
7.3. Oxidation
Catalytic oxidation is also widely used in industry and increasingly in pollution control.
Although in some cases it is desirable to achieve complete oxidation, as in the
elimination of nitrogen oxide from engine emission nitric acid from ammonia, in others
partial oxidation is the aim. For example, the complete oxidation of propane to carbon
dioxide and water is wasteful, but its partial oxidation to propenal (acrolein,
CH2=CHCHO) is the start of important industrial processes. Likewise, the controlled
oxidations of ethane to ethanol, ethanal (acetaldehyde), and (in the presence of acetic acid
or chlorine) to chloroethene (vinyl chloride, for the manufacture of PVC), are the initial
stages of very important chemical industries.Some of these oxidation reactions are
catalysed by d-metal oxides of various kinds. [6]

7.4. Water-gas shift (WGS) reaction


The water gas shift reaction was first used industrially at the beginning of the 20th
century. Today the WGS reaction is used primarily to produce hydrogen that can be used
for further production of methanol and ammonia.
WGS reaction:
CO + H2O ↔ H2 + CO2
The reaction refers to carbon monoxide (CO) that reacts with water (H2O) to from carbon
dioxide (CO2) and hydrogen (H2). The reaction is exothermic with ΔH= -41.1 kJ/mol and
have an adiabatic temperature rise of 8–10 °C per percent CO converted to CO2 and H2.
The most common catalysts used in the water-gas shift reaction are the high temperature
shift (HTS) catalyst and the low temperature shift (LTS) catalyst. The HTS catalyst
consists of iron oxide stabilized by chromium oxide, while the LTS catalyst is based on
copper. The main purpose of the LTS catalyst is to reduce CO content in the reformate
which is especially important in the ammonia production for high yield of H 2. Both
30

catalysts are necessary for thermal stability, since using the LTS reactor alone increases
exit-stream temperatures to unacceptable levels.
The equilibrium constant for the reaction is given as:
Kp= (pH2 x pCO2)/ (pCO x pH2O)

Kp= e((4577.8K/T-4.22))
Low temperatures will therefore shift the reaction to the right, and more products will be
produced. The equilibrium constant is extremely dependent on the reaction temperature,
for example is the Kp equal to 228 at 200 °C, but only 11.8 at 400 °C. The WGS reaction
can be performed both homogenously and heterogeneously, but only the heterogeneously
way is used commercially.
7.4.1. High temperature shift (HTS) catalyst
The first step in the WGS reaction is the high temperature shift which is carried out at
temperatures between 320 °C and 450 °C. As mentioned before, the catalyst is a
composition of iron-oxide, Fe2O3(90-95%), and chromium oxides Cr2O3 (5-10%) which
have an ideally activity and selectivity at these temperatures. When preparing this
catalyst, one of the most important step is washing to remove sulfate that can turn into
hydrogen sulfide and poison the LTS catalyst later in the process. Chromium is added to
the catalyst to stabilize the catalyst activity over time and to delay sintering of iron oxide.
Sintering will decrease the active catalyst area, so by decreasing the sintering rate the life
time of the catalyst will be extended. The catalyst is usually used in pellets form, and the
size play an important role. Large pellets will be strong, but the reaction rate will be
limited.
In the end, the dominate phase in the catalyst consist of Cr 3+ in α-Fe2O3 but the catalyst is
still not active. To be active α-Fe2O3 must be reduced to Fe3O4 and CrO3 must be reduced
to Cr2O3. This usually happens in the reactor start-up phase and because the reduction
reactions are exothermic the reduction should happen under controlled circumstances.
The lifetime of the iron-chrome catalyst is approximately 3–5 years, depending on how
the catalyst is handled.
Even though the mechanism for the HTS catalyst has been done a lot of research on,
there is no final agreement on the kinetics/mechanism. Research has narrowed it down to
31

two possible mechanisms: a regenerative redox mechanism and an


adsorptive(associative) mechanism.
The redox mechanism is given below:
First a CO molecule reduces an O molecule, yielding CO2 and a vacant surface center:
CO+(O) →CO2 + (*)
The vacant side is then reoxidized by water, and the oxide center is regenerated:
H2O+(*)→H2+ (O)
The adsorptive mechanism assumes that format species is produced when an adsorbed
CO molecule reacts with a surface hydroxyl group:
H2O →OH(ads)+ H(ads)

CO(ads)+ OH(ads)→COOH (ads)


The format decomposes then in the presence of steam:
COOH(ads)→CO2+H(ads)

2H(ads)→H2

7.4.2. Low temperature shift (LTS) catalyst


The low temperature process is the second stage in the process, and is designed to take
advantage of higher hydrogen equilibrium at low temperatures. The reaction is carried out
between 200 °C and 250 °C, and the most commonly used catalyst is based on copper.
While the HTS reactor used an iron-chrome based catalyst, the copper-catalyst is more
active at lower temperatures thereby yielding a lower equilibrium concentration of CO
and a higher equilibrium concentration of H2. The disadvantage with a copper catalysts is
that it is very sensitive when it comes to sulfide poisoning, a future use of for example a
cobalt- molybdenum catalyst could solve this problem. The catalyst mainly used in the
industry today is a copper-zinc-alumina (Cu/ZnO/Al2O3) based catalyst.
Also the LTS catalyst has to be activated by reduction before it can be used. The
reduction reaction CuO + H2 →Cu + H2O is highly exothermic and should be conducted
in dry gas for an optimal result.
32

As for the HTS catalyst mechanism, two similar reaction mechanisms are suggested. The
first mechanism that was proposed for the LTS reaction was a redox mechanism, but later
evidence showed that the reaction can proceed via associated intermediates. The different
intermediates that is suggested are: HOCO, HCO and HCOO. In 2009 there are in total
three mechanisms that are proposed for the water-gas shift reaction over Cu(111), given
below.
Intermediate mechanism (usually called associative mechanism): An intermediate is first
formed and then decomposes into the final products:
CO + (species derived from H2O) →Intermediate→CO2

Associative mechanism:
CO2 produced from the reaction of CO with OH without the formation of an
intermediate:
CO + OH →H + CO2
Redox mechanism:
Water dissociation that yields surface oxygen atoms which react with CO to produce
CO2:
H2O→O (surface)

O (surface) + CO → CO2
It is not said that just one of these mechanisms is controlling the reaction, it is possible
that several of them are active. Q.-L. Tang et al. has suggested that the most favorable
mechanism is the intermediate mechanism (with HOCO as intermediate) followed by the
redox mechanism with the rate determining step being the water dissociation.
For both HTS catalyst and LTS catalyst the redox mechanism is the oldest theory and
most published articles support this theory, but as technology has developed the
adsorptive mechanism has become more of interest. One of the reasons to the fact that the
literature is not agreeing on one mechanism can be because of experiments are carried out
under different assumptions.

7.5. Production of Carbon Monoxide


33

CO must be produced for the WGS reaction to take place.


This can be done in different ways from a variety of carbon sources such as:

-passing steam over coal:


C + H2O = CO +H2
-steam reforming methane, over a nickel catalyst:
CH4 + H2O = CO +3H2
-or by using biomass. Both the reactions shown above are highly endothermic and can be
coupled to an exothermic partial oxidation. The products of CO and H2 are known as
syngas.
When dealing with a catalyst and CO, it is common to assume that the intermediate CO-
Metal is formated before the intermediate reacts further into the products. When
designing a catalyst this is important to remember. The strength of interaction between
the CO molecule and the metal should be strong enough to provide a sufficient
concentration of the intermediate, but not so strong that the reaction will not continue.
CO is a common molecule to use in a catalytic reaction, and when it interacts with a
metal surface it is actually the molecular orbitals of CO that interacts with the d-band of
the metal surface. When considering a molecular orbital(MO)-diagram CO can act as an
σ-donor via the lone pair of the electrons on C, and a π-acceptor ligand in transition metal
complexes. When a CO molecule is adsorbed on a metal surface, the d-band of the metal
will interact with the molecular orbitals of CO. It is possible to look at a simplified
picture, and only consider the LUMO (2π*) and HOMO (5σ) to CO. The overall effect of
the σ-donation and the π- back donation is that a strong bond between C and the metal is
being formed and in addition the bond between C and O will be weakened. The latter
effect is due to charge depletion of the CO 5σ bonding and charge increase of the CO 2π*
antibonding orbital.
When looking at chemical surfaces, many researchers seems to agree on that the surface
of the Cu/Al2O3/ZnO is most similar to the Cu(111) surface. Since copper is the main
catalyst and the active phase in the LTS catalyst, many experiments has been done with
copper. In the figure given here experiments has been done on Cu(110) and Cu(111). The
figure shows Arrhenius plot derived from reaction rates. It can be seen from the figure
34

that Cu(110) shows a faster reaction rate and a lower activation energy. This can be due to
the fact that Cu(111) is more closely packed than Cu(110).

7.6. Methanol production

Production of methanol is an important industry today and methanol is one of the largest
volume carbonylation products. The process uses syngas as feedstock and for that reason
the water gas shift reaction is important for this synthesis. The most important reaction
based on methanol is the decomposition of methanol to yield carbon monoxide and
hydrogen. Methanol is therefore an important raw material for production of CO and H 2
that can be used in generation of fuel.

BASF was the first company (in 1923) to produce methanol on large-scale, then using a
sulfur-resistant ZnO/Cr2O3 catalyst. The feed gas was produced by gasification over coal.
Today the synthesis gas is usually manufactured via steam reforming of natural gas. The
most effective catalysts for methanol synthesis are Cu, Ni, Pd and Pt, while the most
common metals used for support are Al and Si. In 1966 ICI (Imperial Chemical
Industries) developed a process that is still in use today. The process is a low-pressure
process that uses a Cu/ZnO/Al2O3 catalyst where copper is the active material. This
catalyst is actually the same that the low-temperature shift catalyst in the WGS reaction is
using. The reaction described below is carried out at 250 °C and 5-10 MPa:
CO+2H2→CH3OH (l)

CO2+3H2→CH3OH (l) +H2O (l)


Both of these reactions are exothermic and proceeds with volume contraction. Maximum
yield of methanol is therefore obtained at low temperatures and high pressure and with
use of a catalyst that has a high activity at these conditions. A catalyst with sufficiently
high activity at the low temperature does still not exist, and this is one of the main
reasons that companies keep doing research and catalyst development.
A reaction mechanism for methanol synthesis has been suggested by Chinchen et al.:
35

CO2→CO2*

H2→2H*

CO2*+ H*→HCOO*

HCOO*+3H*→CH3OH+ O*

CO+ O*→CO2

H2 + O*→H2O
Today there are four different ways to catalytically obtain hydrogen production from
methanol, and all reactions can be carried out by using a transition metal catalyst (Cu,
Pd):
7.6.1. Steam reforming
The reaction is given as:
CH3OH(l)+ H2O (l) →CO2+ 3H2 ΔH= +131 KJ/mol
Steam reforming is a good source for production of hydrogen, but the reaction is
endothermic. The reaction can be carried out over a copper-based catalyst, but the
reaction mechanism is dependent on the catalyst. For a copper-based catalyst two
different reaction mechanisms have been proposed, a decomposition-water-gas shift
sequence and a mechanism that proceeds via methanol dehydrogenation to methyl
formate. The first mechanism aims at methanol decomposition followed by the WGS
reaction and has been roposed for the Cu/ZnO/Al2O3:
CH3OH+ H2O →CO2+ 3H2

CH3OH→CO+ 2H2

CO+ H2O →CO2+H2


The mechanism for the methyl format reaction can be dependent of the composition of
the catalyst. The following mechanism has been proposed over Cu/ZnO/Al2O3:
2CH3OH→CH3OCHO+ 2H2
36

CH3OCHO+H2O→HCOOH+CH3OH

HCOOC→CO2+H2
When methanol is almost completely converted CO is being produced as a secondary
product via the reverse water-gas shift reaction.

7.6.2. Methanol decomposition


The second way to produce hydrogen from methanol is by methanol decomposition:
CH3OH(l)→ CO + 2H2 ΔH= +128 KJ/mol
As the enthalpy shows, the reaction is endothermic and this can be taken further
advantage of in the industry. This reaction is the opposite of the methanol synthesis from
syngas, and the most effective catalysts seems to be Cu, Ni, Pd and Pt as mentioned
before. Often, a Cu/ZnO-based catalyst is used at temperatures between 200 and 300 °C
but a production of by-product as dimethyl ether, methyl format, methane and water is
common. The reaction mechanism is not fully understood and there are two possible
mechanism proposed (2002) : one producing CO2 and H2 by decomposition of formate
intermediates and the other producing CO and H2 via a methyl formate intermediate.

7.6.3. Partial oxidation


Partial oxidation is a third way for producing hydrogen from methanol. The reaction is
given below, and is often carried out with air or oxygen as oxidant :
CH3OH(l) + 1/2 O2 → CO2 + 2H2 ΔH=-155 KJ/mol
The reaction is exothermic and has, under favorable conditions, a higher reaction rate
than steam reforming. The catalyst used is often Cu (Cu/ZnO) or Pd and they differ in
qualities such as by-product formation, product distribution and the effect of oxygen
partial pressure.
7.6.4. Combined reforming
Combined reforming is a combination of partial oxidation and steam reforming and is the
last reaction that is used for hydrogen production. The general equation is given below:
37

(s+p)CH3OH(l) +sH2O(l) + 1/2pO2→ (s+p)CO2 +(3s+2p)H2


s and p are the stoichiometric coefficients for steam reforming and partial oxidation,
respectively. The reaction can be both endothermic and exothermic determined by the
conditions, and combine both the advantages of steam reforming and partial oxidation.
7.6.5. Ammonia synthesis
Ammonia synthesis was discovered by Fritz Haber, by using an iron catalysts. The
ammonia synthesis advanced between 1909 and 1913, and two important concepts were
developed; the benefits of a promoter and the poisoning effect (see catalysis for more
details).

Ammonia production was one of the first commercial processes that required the
production of hydrogen, and the cheapest and best way to obtain hydrogen was via the
water-gas shift reaction. The, Haber-Bosch process is the most common process used in
the ammonia industry.
A lot of research has been done on the catalyst used in the ammonia process, but the main
catalyst that is used today is not that dissimilar to the one that was first developed. The
catalyst the industry use is an promoted iron catalyst, where the promoters can be K 2O
(Potassium oxide), Al2O3 ( Aluminium oxide) and CaO (Calcium oxide ) and the basic
catalytic material is Fe. The most common is to use fixed bed reactors for the synthesis
catalyst.
The main ammonia reaction is given below:
N2+ 3H2↔ 2NH3
The produced ammonia can be used further in production of nitric acid via the Ostwald
process.[14]

8. Catalyst characterization techniques

Experimental surface characterization methods were used to study ageing-induced


changes in the active metals and washcoat oxides. The characterization techniques
included both microscopic and spectroscopic methods, such as Scanning Electron
Microscopy (SEM) and X-ray Diffraction (XRD). Several characterization techniques are
available to study solid surfaces and the properties of catalysts, and no single
38

characterization method can be used to explain the basis for the catalyst deactivation
phenomena of three-way catalysts .

8.1. Scanning Electron Microscopy

Scanning Electron Microscopy (SEM) was used for high magnification imaging and
elemental analysis. A Jeol JSM-6400 scanning electron microscope equipped with an
energy dispersive spectrometer (EDS) was used for the analysis. In the pretreatment
stage, flat pieces of fresh and aged catalysts were cut, and either potted in epoxy or
fastened with a carbon tape in order to obtain side or top views of the catalyst
respectively. Catalysts were polished down to 1 µm using diamond paste and coaled prior
the analysis to avoid the accumulation of charge. The accelerating voltage and current in
the measurements were 15 kV and 12 nA, respectively, and the resolution of the
instrument was 3.5 nm (35 kV).

8.2. Physisorption analyses

Measurements of gas adsorption isotherms are widely used for determining the surface
area and pore size distribution of solids. The use of nitrogen as the adsorptive gas is
recommended if the surface areas are higher than 5 m2/g (Serwicka 2000). The first step
in the interpretation of a physisorption isotherm is to identify the isotherm type. This in
turn allows for the possibility to choose an appropriate procedure for evaluation of the
textural properties. Non-specific Brunauer-Emmett-Teller (BET) method is the most
commonly used standard procedure to measure surface areas, in spite of the
oversimplification of the model on which the theory is based. The BET equation is
applicable at low p/p0 range and it is written in the linear form (Wachs 1992):

Equation

where
39

p is the sample pressure,

p0 is the saturation vapour pressure,

na is the amount of gas adsorbed at the relative pressure p/p0,

nam is the monolayer capacity, and

C the so-called BET constant.

Equation can be applied for determining the surface areas and pore volumes from
adsorption isotherms. The pore size distributions can be calculated from desorption
isotherms. The pores are usually classified according to their widths as micropores
(diameter less than 2 nm), mesopores (diameter between 2 and 50 nm) and macropores
(diameter exceeding 50 nm) (Hayes & Kolaczkowski 1997). Several approaches have
been developed to assess the micro- and mesoporosity, and to compute pore size
distribution from the adsorption-desorption data. All of these involve a number of
assumptions, e.g. relating to pore shape and mechanism of pore filling. (Serwicka 2000)

In this physisorption measurements were carried out to characterize catalysts before and
after the ageings. Specific surface areas (m2/g) and pore volumes were measured
according to the standard BET method, as described above, by using a Coulter Omnisorp
360CX. The specific surface areas and pore volumes were obtained from N2 adsorption
isotherms at -196°C by assuming the cylindrical shape of pores. Catalysts were outgassed
in a vacuum at 140°C overnight before the measurements. All the BET values in this
study were measured within a precision of ± 5%. Pore size distributions for micropores as
well as meso- and macropores were calculated from N2 -desorption isotherms by
differential HK (Horvath-Kawazoe) and BJH (Barrett-Joyner-Hallender) methods
respectively (see Anon 1992). Since the monoliths showed systematically lower BET
values than the crushed samples after similar ageing procedures, all the BET values
presented in this thesis have been determined for the metallic monoliths with a standard
shape and mass.
40

8.3. Chemisorption analyses

Chemisorption measurements were carried out in order to determine the dispersions of Pd


and Rh metal particles, monolayer capacities and the amount of active metal in the
catalysts. Hydrogen and carbon monoxide were used as the adsorbate gases. H 2-
chemisorption and CO-chemisorption experiments were carried out close to room
temperature (30°C) by volumetric adsorption method by using a Coulter Omnisorp
360CX and a Sorptomatic 1900), respectively. The accuracy of the measurements was
estimated to be better than ± 5%. The experimental procedure for the H 2-chemisorption
measurements is presented in Table . In the chemisorption procedure, the temperature
ramping rate of the furnace was 10°C/min. As shown in Table, the adsorption of H 2 was
measured twice. The difference between these two measurements was assumed to be the
amount of irreversibly adsorbed H2, which is further used to calculate the dispersion
values.

The experimental procedure for the H2-chemisorption measurements

1. Flow of He at 150°C for 5 minutes followed by 10 minutes at 375°C


2. Evacuation at 375°C for 10 minutes
3. Reduction in flowing H2 at 375°C for 10 minutes followed by 5 minutes at 400°C
4. Evacuation at 400°C for 20 minutes followed by 10 minutes at 30 °C
5. Leak test at 30 °C
6. First analysis with H2 at 30 °C
7. Evacuation at 30 °C for 30 minutes
8. Second analysis with H2 at 30 °C

Chemisorption has long been employed as a valuable technique for rapid evaluation of
the active metal dispersions and hence the particle sizes of supported metals (Gasser
1985). This method has, however, undergone severe criticism, since the underlying
assumptions of the stoichiometry between adsorbate gas and precious metal and the
particle geometry may not be true, in particular in the case of small particles (Di Monte et
al. 2000). Furthermore, in the case of metal oxides (such as CeO 2 and Ce-Zr-mixed
oxides) in contact with active metals, adsorbed H 2-molecules can also diffuse from the
41

active metal particles to the washcoat. This spillover effect can be reduced by lowering
the adsorption temperature, as has been reported by Bernal et al. (1993) and Fornasiero et
al. (1995).

Active metal dispersions and particle sizes are calculated by assuming the stoichiometry
factor between chemisorbed gas molecules and surface metal atoms. In this thesis,
chemisorption measurements are based on the assumption of the stoichiometry of 2:1 for
H2 and the stoichiometry 1:1 for CO adsorption, respectively, and regardless of the
particle size. The stoichiometric ratio may depend on the precious metal particle size, a
reason why caution should be exercised when comparing the dispersion values of
different catalysts. However, it is assumed that all the aged catalysts as well as the fresh
catalyst exhibit rather large metal particle sizes due the low dispersion values (below
30%). Therefore, the changes in dispersion values presented in this thesis reflect the
structural changes induced by ageings, such as the sintering of the precious metals. As
well, the chemical correctness behind the stoichiometry assumptions is not relevant
because, in this case, the relative dispersion values are more interesting than the absolute
ones.

8.4. Activity measurements

Catalytic activities were determined by laboratory scale light-off experiments to compare


the catalysts after the ageings. Catalyst light-off is determined as the temperature of 50%
conversion, which is used to indicate the efficiency of an automotive exhaust gas catalyst
(the lower the light-off temperature, the more active the catalyst is). In addition to light-
off temperatures, the conversions of CO and NO at 400°C were also determined.

The experimental set-up for the activity measurements is presented in Figures. Catalytic
activities were determined by using a simple model reaction: the reduction of NO by CO
in lean and rich conditions. The composition of test gas mixture is presented in Table.
Before the measurements, the catalysts were reduced in a hydrogen flow at 500°C for 10
minutes, followed by 15 minutes at 550°C. Activity measurements were carried out at
atmospheric pressure by using a cylindrical catalyst with a volume of 1.4 cm 3 (length 28
42

mm and diameter 8 mm). In the measurements, the gas-solid reactor system equipped
with mass flow controllers, magnetic valves for flow selection, tubular furnace with a
quartz reactor and analysis instruments were used. The total gas flow during the
experiments was 1 dm3/min corresponding to the feed gas hourly space velocity (GHSV)
of 43 000 h-1. The temperature of a catalyst was increased from room temperature up to
400°C, with a linear heating rate of 20°C/min. The concentrations of CO, NO, CO 2, N2O
and NO2 as a function of temperature were measured every 5 seconds by an FTIR gas
analyser and the gas flow was controlled by mass flow controllers. Furthermore, the
effect of poisoning on the catalytic activity was evaluated by changing the flow direction
in the catalyst. Blank tests were carried out with the uncoated metal foil to ensure the
inactivity of metal foil in the thermal treatments. In the following discussion, differences
larger than ± 5°C in the light-off temperatures and ± 1% in the conversions of CO and
NO can be regarded as statistically significant.

Fig. Experimental set-up for the activity measurements


43

Composition of the test gas mixture for the activity measurements

Component Lean Rich


CO 800 ppm 1200 ppm
NO 1000 ppm 1000 ppm
N2 Balance Balance

Fig. Activity measurement system equipped with the GASMETTM gas analyzer

The activity of some aged pre-catalysts and main catalysts was also tested at Kemira
Metalkat Oy, Finland. In these measurements, the conversions of CO, HC and NO X were
44

measured as a function of catalyst’s temperature using a test gas mixture simulating the
real exhaust gas composition. Catalyst light-off temperatures (T50 values) as well as
conversions at 400°C were determined. Furthermore, OSC was measured for fresh and
aged monoliths by CO-O2 exchange experiments at constant adsorption temperatures of
450°C, 600°C and 750°C. The consumption and the adsorption of O 2 and CO were
determined by mass spectrometer. (Härkönen et al. 2001)

8.5. Chemical analyses

Chemical analysis provides the information on elemental composition of the catalyst. The
‘wet’ and ‘dry’ analyses were performed due to low quantities of poisons present in the
catalysts. The dry analysis was carried out beyond the SEM-EDS sensitivity, as described
in section 8.1. In a wet analysis, the fresh and aged catalysts were dissolved in an acidic
solution in order to determine the quantities of the most important catalytic poisons (Ca,
P, S, Pb, Mg and Zn) and the amounts of active metals (Pd and Rh). These elements were
typically present in small quantities, and unevenly distributed in the catalyst. Therefore,
separated samples were prepared from the inlet and outlet parts of the engine-aged and
vehicle-aged catalysts. For the chemical analysis, 0.010 to 0.050 g of the washcoat
(scraped from the monolith) was dissolved in an acidic solution (HNO 3, HCl, HF and
H3BO3) and subjected to the digestion of the sample in the microwave oven (Milestone
MLS 1200). This resulted partly in an incomplete dissolution of the analysed solids. The
decomposed sample was analysed quantitatively by plasma atomic emission spectrometry
(Pye Unicam 7000 ICP-AES). [15]
45

References

1. http://www.chemguide.co.uk/physical/catalysis/introduction.html

2. Chaudhry. G. Rasool, A Text Book of Physical Chemistry, (2009). Lahore: Azeem


Publishers.
3. http://chemed.chem.wisc.edu/chempaths/GenChem-Textbook/Heterogeneous-

Catalysis/chemprime/CoreChem3ACatalysis-1005.html

4. http://www.chemguide.co.uk/physical/catalysis/introduction.html

5. http://chemed.chem.wisc.edu/chempaths/GenChem-Textbook/Heterogeneous-

Catalysis-1005.html

6. Bhatti. H. Nawaz, Principles of Physical Chemistry, (2008). Lahore: The Caravan


Book House.
7. http://www.chemguide.co.uk/physical/catalysis/introduction.html

8. Moora. J. Walter, Physical Chemistry, (1991).

9. http://www.chemguide.co.uk/physical/catalysis/introduction.html

10. http://www.chem.canterbury.ac.nz/letstalkchemistry/electronicversion/

electronicversionnew/chapter16/catalyst.shtml

11. http://www.chemguide.co.uk/physical/catalysis/introduction.html

12. http://chemed.chem.wisc.edu/chempaths/GenChem-Textbook/Heterogeneous-

Catalysis/chemprime/Catalysis_in_Everyday_Life-1005.html

13. Atkin. P. Physical Chemistry, (2006). New York: W. H. Freeman & Company.

14. http://en.wikipedia.org/wiki/Industrial_catalysts

15. http://herkules.oulu.fi/isbn9514269543/html/x900.html

Das könnte Ihnen auch gefallen