HC&E Introduction To Catalysis

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

An Introduction to Catalysis

Christian Hulteberg
Professor, founder and managing director of
Hulteberg Chemistry & Engineering AB
Copyright © 2017 by Christian Hulteberg

All rights reserved. This compendium or any portion thereof may not be reproduced or
used in any manner whatsoever without the express written permission of the publisher
except for the use of brief quotations.

Hulteberg Chemistry & Engineeing AB


Rödklintsgatan 2b
218 73 Tygelsjö
Sweden

www.Hulteberg.com
Contents
1 Introduction .............................................................................................................. 1
1.1 Definition ........................................................................................................... 1
1.2 Disposition ......................................................................................................... 4
2 Background ............................................................................................................... 5
2.1 Reaction rates, orders and rate-limiting steps .................................................... 6
2.2 Catalysis kinetics ............................................................................................. 10
2.3 Catalyst and reactant/product interactions ....................................................... 11
3 Heterogeneous catalysis ......................................................................................... 12
3.1 Structure and components ................................................................................ 12
3.2 Characterisation ............................................................................................... 15
3.2.1 Surface area .............................................................................................. 15
3.2.2 Temperature programmed methods .......................................................... 16
3.2.3 Spectroscopic and microscopic methods .................................................. 17
3.3 Development and testing ................................................................................. 20
3.3.1 Reactor gradients ...................................................................................... 20
3.3.2 Test reactors .............................................................................................. 21
3.4 Deactivation ..................................................................................................... 24
3.4.1 Poisoning .................................................................................................. 24
3.4.2 Fouling ...................................................................................................... 25
3.4.3 Sintering ................................................................................................... 25
3.4.4 Loss of material ........................................................................................ 26
3.5 Manufacture ..................................................................................................... 26
3.5.1 Precipitation and co-precipitation............................................................. 27
3.5.2 Hydrothermal synthesis ............................................................................ 27
3.5.3 Impregnation of porous supports .............................................................. 28
3.5.4 Summary of synthesis unit operation ....................................................... 28
4 Homogeneous catalysis ............................................................................ 31
4.1 Structure, components and elementary steps ................................................... 31
4.1.1 Elementary steps ....................................................................................... 32
4.1.2 Structure ................................................................................................... 34
4.2 Characterisation ............................................................................................... 34
4.3 Development and testing ................................................................................. 34
4.4 Deactivation ..................................................................................................... 35
4.5 Manufacture ..................................................................................................... 35
5 Conclusions ............................................................................................................ 36
6 References .............................................................................................................. 37
1 Introduction
Catalysis is omnipresent in the chemical industry. It is found in almost all instances of
bulk chemical production, in emission abatement, fine chemical synthesis and production
of transportation fuels, see Figure 1 for examples of catalysis in chemicals production.
But the phenomenon is not limited to the production of chemicals alone - many biological
systems, e.g. the photosynthesising systems, employ catalysts in the form of enzymes.

Figure 1. The importance of catalysis in the chemical industry (catalytic processes are denoted with a small
dot).

1.1 Definition
Catalysis combines the knowledge from many scientific and technical fields such as in-
organic chemistry, organic chemistry, surface chemistry, thermodynamics, chemical ki-
netics, solid state physics and material science. The phenomenon of catalysis is old, but
the term was coined by J.J. Berzelius in 1835 when he coordinated a number of observa-
tions and introduced the concept of catalysis as the ability

”to awaken affinities, which are asleep at a particular temperature, by their mere pres-
ence and not by their own affinity“.

1
In other terms, a catalyst is a substance that in a small amount causes a large change.
There is however still today discussions on the definition of catalysts and scientists and
practitioners may have different views. However, the definition that “a catalyst is a sub-
stance that increases the rate of reaction toward equilibrium without being appreciably
consumed in the process” is basically operational. The fundamental concept involves a
process in which a site on a catalyst forms a complex with reactants and products are
formed. The original site is thus restored with the leaving of the products. This should
however not be interpreted as that there are no changes of the catalyst in the process.
Metal surfaces may undergo recrystallization, oxide catalysts may lose oxygen, but the
losses or changes are much lower than any stoichiometry of reaction. Indeed, catalysts
come in many forms – from protons, Lewis acids, organometallic complexes, organic and
inorganic polymers to enzymes, figure 2.

Figure 2. Different types of catalysts (adapted from Wikimedia commons).

2
By the definition given above, the addition of an initiator, e.g. in a polymerization reac-
tion, is not a catalyst. It does yield a large number of reactions per molecule added, but it
is not reformed at the end of the process; the course of events is not cyclic. It also excludes
the actions of heat, microwaves, photochemical or other radiation since they are not sub-
stances added. Hence, the action of UV-radiation to increase the reaction rate of H2 oxi-
dation by O2 is not catalytic. Finally, the catalyst cannot change the ultimate equilibrium
determined by thermodynamics, it only accelerates the rate of approach to equilibrium,
figure 3.

Figure 3. Different types of catalysts.

There is also a distinction between heterogeneous and homogeneous catalysis. In hetero-


geneous catalysis the catalyst and the reactants/products are in different states of aggre-
gation, e.g. solid/liquid or solid/gas. In homogenous catalysis the catalyst and the reac-
tants/products are in the same state of aggregation. This latter term has however more and
more come to describe the liquid phase catalytic action of organometallic complexes. As
a side-note, there are no a priori rules as to which type of catalytic process (heterogeneous
or homogeneous) is more economic for a given reaction. Each case will have to be as-
sessed individually, including aspects such as: catalyst selectivity, activity, cost of mate-
rials, waste-creation and so on, in the analysis.

If homogeneous catalysts have come to mean organometallic complexes, heterogeneous


catalysts consist of a carrier, a support phase and an active phase. The carrier ensures
structural integrity in the reactor, e.g. pellets or monolithic structures, and the support
provides the surface area needed for reaction, usually well above 100 m2/g of catalyst,
e.g. aluminum oxide and ceria oxide supports. The active phase is where the reaction
occurs, at specific sites on the surface of the active phase. Common active phases are
noble and transition metals as well as their oxides. In some instances the carrier is made
out of the support material and the support and the active phase can be one and the same,
e.g. alumina in dehydration reactions. An active site may be a single surface atom or a
cluster of surface atoms.
3
Finally, there are also negative catalysts, i.e. catalysts that lower the reaction rate for a
given reaction. These are most commonly found when dealing with radical reactions and
they act by interfering with the free-radical process, converting radicals into less active
forms or removing them from reactions. The prime example is the action of lead as anti-
knocking agent in gasoline fuels. Wherein this case decomposition products of the lead
interact with organic peroxide intermediates (formed in the early stage of combustion)
and eliminate them before they can decompose into radicals which give rise to knocking.

1.2 Disposition
In this compendium the concept of catalysis will be introduced and described with the
purpose of giving an introduction to the subject. The focal point of the text will be heter-
ogeneous catalysis as it is the dominant tool in the production of bulk and intermediate
chemicals. However, the homogeneous catalysts, mostly relevant in fine chemical syn-
thesis, will also be described.

4
2 Background
Catalysis is a kinetic phenomenon and it deals with changes on the route to chemical
equilibrium, it is thus about the kinetics of reaction and not thermodynamic equilibrium.
In passing a mountain range, heading from point A to point B, the reaction thermodynam-
ics describe the elevation in point A and B, while reaction kinetics determine the path
taken to get from point A to point B. Kinetics will be dealt with in little more detail later
on, but first there is a need to discuss other important concepts.

The activity of a catalyst refers to the rate at which it causes the reaction at hand to pro-
ceed towards chemical equilibrium. The rate may be expressed in several ways but the
most common way to denote performance of an industrial reactor is as space-time yield,
which is the quantity of product formed per unit time per volume of reactor, e.g.
[kg s-1m-3] or [mol h-1L-1]. This rate naturally depends on operating conditions such as
temperature, pressure, concentration of reactants and products etc. To compare rates they:

• may be determined at a pre-determined standard condition


• may sometime be reported as a single rate constant, which require detailed
knowledge about true kinetics
• may be expressed as the temperature for achieving a certain conversion for a given
feed composition and pressure
• may be given as the amount of catalyst required for converting a given inlet feed
at a given operation temperature and pressure

The selectivity describes the ability of a catalyst to produce the desired product. It is
determined firstly by the functionality of the catalyst and secondly by the thermodynamic
equilibrium. As an example, please consider ethanol. Ethanol will in the presence of a
copper catalyst dehydrogenate according to the formula below

C2H5OH → CH3CHO + H2

While passed over an alumina catalyst the following reactions occur:

C2H5OH → C2H4 + H2O


2 C2H5OH → C2H5O C2H5 + H2O

Selectivity in this case is determined by the fact that copper adsorb hydrogen (acting as a
dehydrogenation catalyst) while alumina adsorb water and act as a dehydration catalyst.
If the formation of the desired product involves several steps, it is important that each
step is catalyzed by the substance chosen. Selectivity for the production of a product B
from a reagent A may thus be defined as:

𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐴 𝑐𝑜𝑛𝑣𝑒𝑟𝑡𝑒𝑑 𝑡𝑜 𝐵


𝑆𝑒𝑙𝑒𝑐𝑡𝑖𝑣𝑖𝑡𝑦𝐵 =
𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐴 𝑐𝑜𝑛𝑠𝑢𝑚𝑒𝑑
5
Another important concept in catalysis (and reaction engineering in general) is the yield,
which can be defined as:

𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐵 𝑓𝑜𝑟𝑚𝑒𝑑


𝑌𝑖𝑒𝑙𝑑 =
𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐴 𝑒𝑛𝑡𝑒𝑟𝑒𝑑 𝑖𝑛𝑡𝑜 𝑡ℎ𝑒 𝑟𝑒𝑎𝑐𝑡𝑜𝑟

In catalysis there is also the notion of turnover frequency (TOF) and turnover (TON). In
homogenous catalysis TON is the number of cycles a catalyst can run before being deac-
tivated, i.e. the number of times a specific catalyst can convert reactant A into product B.
The TOF is the TON divided by a selected unit of time or the number of times a catalyst
can convert reactant A into product B in one second, minute or hour. In heterogeneous
catalysis the TON and TOF are defined per active site or, more commonly, due to diffi-
culties in establishing the active-site concentration, per g of catalyst, per m2 of catalyst or
per m2 of active phase. For clarity always make sure to denote the units of TON and TOF
to avoid confusion.

2.1 Reaction rates, orders and rate-limiting steps


The rate of the reaction A → B is usually denoted r or ν; assuming a constant volume
batch reactor the rate is simply the number of A consumed or indeed B forming per unit
time:

𝑑[𝐴] 𝑑[𝐵]
𝑟=− =
𝑑𝑡 𝑑𝑡

The rate equation can then be written rate = -rA = k*[A] where [] denote concentration
of the species. The k is the reaction rate constant and its dimension is dictated by the
rate equation. The reaction rate constant is only constant for a given temperature and
pressure and shows strong temperature dependence but weak pressure dependence. For
more complex reactions, e.g. A + 2B → 3C + D, the reaction rate is chosen arbitrarily for
a product or a reactant and the stoichiometric constraints give the other rates

𝑑[𝐴] 1 𝑑[𝐵] 1 𝑑 [𝐶] 𝑑[𝐷]


𝑟=− =− = =
𝑑𝑡 2 𝑑𝑡 3 𝑑𝑡 𝑑𝑡

The relation between temperature and reaction rate constant is expressed using the Ar-
𝐸𝑎
rhenius equation (𝑘 = 𝐴 ∗ 𝑒 −𝑅𝑇 ). The equation is empirical and is the result of experi-
mental observations but well describes the behavior of the rate constant over a wide tem-
perature range. The activation energy Ea is a key parameter in catalysis, describing the
energy barrier that a system has to overcome for the reaction to occur. The activation
energy is typically from 8 to 100 kJ mol-1 and when less than 20 kJ mol-1 the reaction is
said to be diffusion controlled, i.e. the reaction is limited by how fast molecules move

6
towards each other. For activation energies above 20 kJ mol-1, reactions are chemically
controlled, figure 4.

Figure 4. The activation energy regimes as a function of temperature.

A good rule-of-thumb says that with every 10°C temperature increase k doubles. The pre-
exponential factor A is without physical description but is attributed to changes in entropy
and substrate orientation or indeed collisions; the factor commonly ranges from 1011 to
1014 per second in the case of first order reactions. Both the pre-exponential factor and
the activation energy can be determined by using an Arrhenius plot in which the natural
logarithm of the reaction rate constant is plotted vs. the inverse absolute temperature,
Figure 5.

7
𝐼𝑛𝑡𝑒𝑟𝑐𝑒𝑝𝑡 = ln (𝐴)

𝐸𝑎
𝑆𝑙𝑜𝑝𝑒 = −
𝑅

Figure 5. Arrhenius plot of the decay of nitrogen oxide (adapted from Wikimedia commons)

The rate of every chemical reaction depends in one way or the other on the concentration
of the chemicals in the system and the overall law is written as the product of the concen-
trations to a power. So for reaction A+B → C with the assumed exponents α and β repre-
senting the partial reaction order in A and B respectively, the overall reaction order is
n = α + β.

−𝑟𝐴 = 𝑘[𝐴]𝛼 [𝐵]𝛽

The reaction rates can be equal to the stoichiometric coefficient of the species (elementary
reaction) but it does not have to (non-elementary reaction). The non-elementary reaction
rates have to be determined experimentally and cannot be determined from the stoichio-
metric equation.

The rate-limiting step is the one determining the overall reaction rate of an equation. A
catalytic cycle consists of several elementary reactions (steps) and the rate-limiting step
is the slowest one of these. It is thus not the average or cumulative rate but the slowest
one(s) in the cycle. The following image is a good representation of the concept, figure
6.

8
Figure 6. The hourglass is a good representation of the rate-limiting step. Wikimedia commons

There is no problem for the grains to pass through the large openings of the hourglass,
however the middle constraint means that the overall flow is limited by this central ori-
fice. The fast reaction steps are analogous to the large openings and the rate-limiting step
to the orifice. Each of the reaction steps has its own Ea and the rate-limiting step is the
one with the highest activation energy, independent on if it comes early or late in the
cycle, figure 7.

Figure 7. Energy/reaction coordinate diagram for a two-step reaction.

9
The reaction order of catalytic processes is often quite complicated when considering the
overall reactions. Fortunately, zero-order, first-order or second-order rate laws, may sur-
prisingly often describe the individual steps. The rate-laws basically show the dependence
of the given reaction of the concentration of the reactants with:

𝑑[𝐴]
− = 𝑘 (𝑧𝑒𝑟𝑜 𝑜𝑟𝑑𝑒𝑟 𝐴 → 𝐵)
𝑑𝑡
𝑑[𝐴]
− = 𝑘([𝐴0 ] − [𝐴]) (𝑓𝑖𝑟𝑠𝑡 𝑜𝑟𝑑𝑒𝑟, 𝐴 → 𝐵 + 𝐶)
𝑑𝑡
𝑑[𝐴]
− = 𝑘([𝐴0 ] − [𝐴])([𝐵0 ] − [𝐵]) (𝑠𝑒𝑐𝑜𝑛𝑑 𝑜𝑟𝑑𝑒𝑟, 𝐴 → 𝐵 + 𝐶)
𝑑𝑡

2.2 Catalysis kinetics


If we add a catalyst the reaction rate will change, but so will the rate law! It will most
likely be dependent on the catalyst concentration and thus change to:

𝑟𝑎𝑡𝑒 = 𝑘[𝑐𝑎𝑡𝑎𝑙𝑦𝑠𝑡]𝛼 [𝐴]𝛽 [𝐵]𝛾

To describe the full catalyst-cycle kinetics there are several different models, in this text
only the most common one for heterogeneous catalysis (Langmuir-Hinshelwood) and ho-
mogeneous catalysis (Michaelis-Menten) will be described briefly. In the Langmuir-Hin-
shelwood model it is assumed that the reactant(s) first adsorb on the catalyst surface be-
fore any reactions occur. The reaction occurs at an active site and the product(s) thereafter
desorb from the catalyst to the fluid phase. In an extension of the model, the diffusion of
the reactants and products to and from the surface to the fluid bulk is also included, fig-
ure 8.

Figure 8. Langmuir-Hinshelwood kinetics model.

10
The three steps have their own differential rate expressions which lead to a number of
complicated differential equations that are best solved numerically. As most catalytic cy-
cles contain several steps that are interconnected, the overall rate equations are quite com-
plex. The measuring of reaction intermediates is also a challenge and therefore it is often
assumed that the concentration of the catalytic intermediates is constant during reaction,
the steady-state concentration. This holds true even for batch processes because in every
cycle the catalytic intermediates are formed and consumed continuously.

The Michaelis-Menten kinetic model was originally developed for the enzymatic reaction
of sugar inversion, but has become frequently used to describe homogeneously catalysed
reactions. The model describes a two-step cycle, where the catalyst first reacts reversibly
with the reactant forming a catalyst-reactant complex. Subsequently the complex decom-
poses into the catalyst and the product.

2.3 Catalyst and reactant/product interactions


The interaction between the catalyst and the substrate, the intermediates and the products
are of extreme importance. If the affinity is too strong, the products will never leave the
catalyst, whereas if it is too weak the reactants cannot adsorb to initiate the reaction. This
is sometimes referred to the as the Goldilocks principle of catalysis (although it does also
have a more formal name - Sabatier’s principle). The functional parameter is the heat of
adsorption, at first when the heat of adsorption increases the reaction rate increase. More
substrate binds to the catalyst surface and there is a higher probability of reaction. When
the heat of adsorption is too high though, the products are unable to leave the surface and
the reaction rate starts to decrease, figure 9.

Figure 9. A schematic representation of the Sabatier principle (volcano plot).

11
3 Heterogeneous catalysis
Heterogeneous catalysis is the intersection between material science, chemistry and en-
gineering, the industrial relevance of this field cannot be exaggerated. Indeed more than
90% by volume of the chemicals produced worldwide involve solid catalysts in the pro-
duction process. Typically the reactants are contacted with the catalyst in the gas phase
at elevated temperature and sometimes at high pressures. About 800 000 tonnes of solid
catalysts are prepared worldwide and used yearly; this however does not mean that we
truly understand how they function. Catalyst synthesis and treatment are most often based
on empirical studies.

There are several concepts and conventions that should be understood to be able to fully
appreciate the field and to be able to have meaningful discussions with catalyst suppliers.
These concepts, along with additional information will be presented in this section.

3.1 Structure and components


The catalysts in industrial processes come in several shapes and forms and are composed
of different components. First of all there is the carrier, ensuring the structural integrity
of the catalyst. In addition, the carrier is to a large extent influencing the reactant/product
diffusion, heat-transfer properties and pressure drop of a chemical reactor. The most com-
mon types are pellets or monoliths, figure 10.

Figure 10. The most common carrier types, extrudates, pellets or prills and monoliths.

To the carrier a support is added to increase the surface area available for reaction, alt-
hough indeed the carrier may in itself sometimes consist of the support. Common supports
are oxides of aluminium, silicon and cerium, but also mixed oxides or carbons. An im-
portant factor for selecting the support is the surface area of the material (measured as
m2/g). Table 1 gives a set of typical data for supports.

12
Table 1. selected materials commonly used as catalyst supports.
Material Area (m2/g)
γ-alumina 180
Activated charcoal <1,100
Silica-alumina 250
Silica gel 200
Zeolite 900

The large surface area of the catalyst support is used for dispersing the active phase. The
active phase contains the sites where the reaction actually occurs. The active phase can
consist of metals or metal oxides and is where the key surface interactions take place,
which lead to reaction. Common materials used as active phases are metals from the tran-
sition metal groups in the periodic table, e.g. Pt, Pd, Cu, Co, Ni, Rh. Figure 11 illustrates
the roles of the carrier, the support and the active phase.

Figure 11. The different elements to heterogeneous catalysis.

A heterogeneously catalysed reaction occurs on a surface and this is a circumstance that


complicates things. The reason is that a surface is not as uniform and ordered as can be
imagined. A surface is filled with irregularities in crystal growth (such as kinks and steps),
which surprisingly often is where reaction occurs, figure 12.

13
Figure 12. Examples of active sites.

The active sites are most often not the preferred sites for adsorption, however they obey
Sabatiers principle in that they are good enough for adsorption and still allowing desorp-
tion of the products. Ideally the active sites are all alike and not in contact with each other.
In real life this is however not the case, mainly because industrial catalysts are amorphous
or partly amorphous and consist of multiple components and several solids, leading to
many types of active sites (zeolites being an exception).

Aside from the support and the active phase, promoters or modifiers are added to the
catalyst. These can be of several kinds and with different promoters and modifiers added
with different purposes. They can be added to enhance the activity or selectivity of a
catalyst. Another purpose for adding promoters or modifiers is to suppress side-reactions,
e.g. alkali is added to steam reforming catalysts to lower carbon formation. Promoters are
often electropositive substances, while catalyst poisons (treated later on in the text) are
electronegative atoms. The promoters may also be of textural nature, added to enhance
the stability of the catalyst support or the active phase. The effects of promoters are based
on empirical observations and may have no initial effect, but increase long-term stability.
This of course requires empirical observation over long periods of time and since the
effects are rarely fundamentally understood, promoters are often left out of academic re-
search.

Since catalysts consist of a number of different components, there is a need for a gener-
alized syntax in naming them. Even though there is more to a catalyst than the elements
of which it is composed, these are used for naming it. Since there is often an uncertainty
as to which phase (metallic, oxide, sulphide etc.) a specie is in under relevant operating
conditions these are usually omitted. For instance the hydrodesulphurisation catalyst con-
sisting of cobalt and molybdenum supported on aluminium oxide is usually denoted
CoMo/Al2O3. The solidus separates the active components (CoMo) from the support
(Al2O3). The catalyst is supplied in an oxide form and converted to a complex sulphide
before use. Sometimes a catalyst may be referred to as a single compound, e.g. “zinc

14
chromite” methanol catalysts. These methanol catalysts were materials consisting of zinc
oxide with chromium oxide incorporated in the structure. However, the zinc chromite
mineral, Zn(Cr2O4), is indeed inactive for the reaction.

3.2 Characterisation
There are many ways of characterising catalysts, with respect to both physical and chem-
ical properties. The methods can be separated into bulk-sensitive and surface-sensitive
methods. The first set of methods gives insight into the overall catalyst composition or
average properties. The second set of methods gives insight into the first 1-5 layers of
atoms on the surface. Both are of interest and complement each other in the type and
quality of information they result in. Since there are quite a few methods available for
characterisation of catalysts, combinations of them are used to make certain that the re-
sults are credible. The methods outlined in this section (fine particle analysis i.e. surface
area methods and temperature programmed methods as well as spectroscopy/microscopy
methods) are the main methods used. The level of detail of description is not very high,
on the other hand there is much information available elsewhere on these methods.
3.2.1 Surface area
Since the first step in any gas/solid reactions system is the adsorption of reactants on the
catalyst surface, the investigation of this surface is of interest. The keyword here is ac-
cessibility of reactants to the surface. The accessibility is largely determined by the size
and appearance of the pores available, as illustrated in figure 13. The pores are defined
by their diameter with micropores having a diameter below 2 nm, mesopores with a
diameter from 2 nm to 50 nm and macropores with a diameter above 50 nm.

Figure 13. Some important pore-types in catalyst a) open pore, b) closed pore, c) ink bottle, d) cylindrical
and open-ended.

To determine the character of a catalyst surface two types of adsorption are employed:
physisorption and chemisorption. In physisorption molecules are adsorbed to a surface
by van der Waals interactions, characterised by low heats of adsorption (10-40 kJ mol-1).
Chemical adsorption on the other hand involves the breaking and forming of chemical
bonds between the surface and the gas, resulting in much higher heats of adsorp-
tion (80-400 kJ mol-1).
15
Physisorption is used for total surface area and pore volume determination. Nitrogen is
the most common probe molecule used, at temperatures close to the boiling point of the
nitrogen (-196°C). Practically, testing is performed by cooling the sample with liquid ni-
trogen and repeatedly injecting a small, known amount of gas into the sample chamber.
The increase in the pressure with each injection is used to calculate the surface area based
on the Braunauer-Emmet-Teller equation, from which the method also has its name: BET
surface measurement. The adsorption can also be used to determine the pore volume and
pore-size distribution. Using nitrogen, pores from about 0.35 nm to 400 nm can be deter-
mined, by changing probe molecule this range can be varied slightly.

Chemisorption is used to determine more specific characteristics of a catalyst, e.g.


Brønstedt acid sites or the area of a metallic active phase. In general, a probe molecule is
introduced in a controlled volume at a pre-determined temperature while measuring how
much is adsorbed by the surface. Chemisorption is often combined with temperature pro-
grammed methods.

As there are limits in the size of pores that can be analysed using adsorption methods, an
alternative method has been developed for measuring the macropores, mercury intru-
sion porosimetry. The method uses liquid mercury which is contacted with the sample
at different pressures, from 0 to 414 MPa. With increasing pressure the surface tension of
the mercury is overcome for smaller and smaller pores, figure 14. The intrusion volume
at every specific pressure is measured and can be used for calculating the pore volume
and pore size distribution. The method measure pores ranging from 0.003-360 μm
(3-360,000 nm).

Figure 14. Mercury intrusion porosimetry in action.

3.2.2 Temperature programmed methods


Under the heading temperature programmed methods or techniques, several different
types of analysis are hidden. The methods have in common that they all measure the
interaction of a species with a surface. The use of oxygen as a reactive agent (temperature

16
programmed oxidation, TPO) reveals how the catalyst oxidises as a function of tempera-
ture. Similarly, the reductive properties of a catalyst as a function of temperature can be
measured using hydrogen (temperature programmed reduction, TPR). By feeding more
than one reactant, the properties of catalyst to perform certain reactions can be investi-
gated (TPRx). The inverse is also used, in determining the acid strength of a catalyst the
sample is first saturated with ammonia at near-ambient temperature. By applying a tem-
perature ramp and measuring the desorbing ammonia, not only the overall number of
acidic sites can be determined but also the strength and frequency of these sites; similar
protocols reveal the basic sites of the catalyst.
3.2.3 Spectroscopic and microscopic methods
This is the area seeing the most development in catalysis characterisation as of today. By
bombarding a surface with photons, electrons or ions, much information regarding the
surface composition, the interactions between species present at the surface and the dis-
tribution of the active phase can be discerned. Some of the methods are bulk methods and
some are surface sensitive, with electron microscopy being the prime example in this
category. Many of these methods have until lately required to be performed under high
vacuum and low temperatures, quite far from actual operating conditions of catalytic re-
actions. However much development has been done and in-situ or operando spectro-
scopic methods are becoming available at higher pressures and temperatures, closing the
gap between analysis and reaction conditions.

3.2.3.1 X-ray diffraction (XRD)


By beaming a surface with x-ray photons from different angles and measuring the elastic
scattering of the photons one can establish an “x-ray finger print” or diffractogram of the
material, figure 15. The fingerprint will contain peaks at certain angles where the crystal
structure of the material result in signal amplification and these peaks can be matched
against existing databases for identification.

Figure 15. Diffraction pattern or diffractogram of silicon carbide. Wikimedia commons.

The method analyse single crystals, but is most commonly performed on powders (pow-
der x-ray diffraction) in catalysis applications. The method can accurately detect down to

17
5-10% of a material in a catalyst, but is only sensitive to crystalline material. X-ray dif-
fraction can be used to determine the particle size of the material with the Debye-Scherrer
equation.

Small-angle x-ray scattering (SAXS) works similarly but while XRD covers a range of
10°-180° scattering range, SAXS handles smaller angles (2°). This enables the determin-
ing of catalyst particle sizes in the range of 50-500 nm and SAXS may be used for calcu-
lating the specific surface area of the particles.

3.2.3.2 Raman spectroscopy


Raman spectroscopy uses concentrated monochromatic light in the visible range, near
infrared or near ultraviolet frequency range to induce inelastic or Raman scattering of the
light. The original light is filtered out and the scattered light is analyzed. The method
produces a fingerprint of the vibrational energy specific to the bonds of the material. It
can then be used for identifying the material in use. In figure 16 the Raman spectrogram
of two WO3/ZrO2 catalysts for glycerol dehydration are shown. The fingerprints reveal a
shift from monoclinic to triclinic tungsten at catalyst pretreatment at 900°C instead of
800°C.

0,14

0,12

0,1

0,08
A.U.

0,06 800 C

0,04 900 C

0,02

0
0 200 400 600 800 1000 1200
-0,02
Frequency (cm-1)

Figure 16. Raman spectrogram of WO3/ZrO2catalysts, image: own work.

3.2.3.3 Electron microscopy


Two types of electron microscopes are mainly used in catalyst research, scanning electron
microscopes (SEM) and transmission electron microscopes (TEM). The SEM uses a
highly concentrated electron beam to sweep a small rectangle of the surface. This gener-
ates low-energy secondary electrons, out of which some escape from the surface and can
be used to “see” the shape of the surface as well as information on the average atom
number in the scanned area, figure 17. The method is specific to the investigated area and
has a significant influence on the surface investigated.

18
Figure 17. SEM image of hematite which is a promising catalyst for water splitting, image from Wikimedia
commons.

In TEM an electron beam is shot at the sample and the electrons that are transmitted
through the sample are magnified by electromagnetic lenses. The optics allows for very
high magnifications and traditional TEMs commonly give 300,000 times magnification
with a 0.5 nm resolution. High-resolution TEMs can give 1,000,000 times in magnifica-
tion and atomic resolution. The method can be combined with other analysis methods,
either using the information contained in the electrons or by simultaneously shooting pho-
tons onto the surface, creating images with additional information, figure 18.

Figure 18. TEM image combined with a line trace from energy dispersive x-ray spectroscopy (XEDS) The
image shows the cobalt intensity along the marked line, image: own work.

3.2.3.4 Infrared spectroscopy


The use of infrared (IR) light is the most widely spread technique for catalyst characteri-
sation. It can give information about the surface chemistry as well as on species adsorbed
on the surface. By using probe molecules, e.g. CO, NO and NH3, information about the
nature and environment of the atoms and ions adsorbed on the surface can be discerned.
The method is based on the absorption, transmission or reflection of IR light by a catalyst
surface. The IR is used for exciting molecular vibration and the analysis can be performed
by transmission of light through a thin slice of a catalyst or indeed by reflection of IR

19
from the surface. The energy of the measured vibrations is compared to “fingerprints” in
literature and can reveal information on which way molecules are adsorbed on surfaces.
The method is surface sensitive, non-destructive and non-invasive.

3.2.3.5 X-ray photoelectron spectroscopy


A method using low-energetic x-ray photons, such as those generated by Al Kα or Mg Kα,
was developed by Swedish Noble laureate Kai Siegbahn in the mid-1960s. The x-ray
photoelectron spectroscopy (XPS) method is straightforward and useful for identifying
surface atoms in catalysis. It is based on the photoelectric effect, i.e. the phenomenon that
electrons are emitted from solids when they absorb energy from light photons. The radi-
ation only penetrates the first few atomic layers of the surface. The kinetic energy of the
emitted electron can be used to measure the bonding energy of the electron in the material.
The method is useful for determining the surface concentration of species as well as for
analysing elemental oxidation states. The method measures 2-20 atom layers deep and is
sensitive for all species except for H and He.

3.3 Development and testing


Laboratory testing of catalysts differs significantly from industrial operation of catalytic
reactors. The reactors are for instance designed for operation at a wide range of tempera-
tures, and sometimes also pressures, compared to industrial reactors that are only de-
signed for one operating condition. The reactor is also flexible in that it should be easy to
replace the catalyst and is equipped with much more instrumentation (sensors for temper-
ature, pressure and flow rate) than an industrial unit. The laboratory reactors can also be
limited with respect to conversion to measure kinetic parameters or indeed operating in
integral mode to mimic actual operations of industrial reactors. The reactors are com-
monly electrically heated and isothermal operation is desired, heat integration is not per-
formed as in industrial type systems.
3.3.1 Reactor gradients
To ensure that what is measured is correct, especially when working with highly exother-
mic or endothermic reactions it is of the essence to determine if there are any gradients in
the reactor. The main gradients to consider is the intra-particle ones (inside the catalyst
particle), interphase (between fluid and particle) and reactor gradients in both radial and
axial dimension. This text is not aimed at the level of detail in which these criteria are
dealt with in detail; however an example of the effects of ignoring the intra-particular
effects will be given to illustrate the effects of ignoring these gradients. The Thiele mod-
ulus describes the relation between the diffusion time and the reaction time and can be
used to express an effectiveness factor (η) relating the diffusive reaction rate in a pore to
the reaction in the bulk stream. As can be seen in figure 19, by plotting a version of the
Thiele modulus (the Weisz modulus, ø) vs. the effectiveness factor, the reaction rate can
be largely overestimated. In the figure, due to insufficient heat removal and a temperature
increase in the catalyst grain or particle, the measured reaction rate surpasses the actual
reaction rate by a factor of 102 in some instances while it is a factor of 10-2 lower for some

20
given conditions. A good way to show that the measured reaction rate is the correct one
is to decrease the catalyst particle size until the reaction rate is constant.

Figure 19. The difference between measured and actual reaction rate of reaction.

3.3.2 Test reactors


There are many different reactor-types available for testing and the choices are dependent
the purpose of the investigation as well as the reaction characteristics. The main catego-
ries are stationary and non-stationary reactors. The non-stationary are divided into batch,
fed-batch and transient reactors while the stationary are continuous reactors of the plug
flow, fluidizing bed or continuously stirred tank reactor types. The plug-flow reactors are
further divided into integral or differential type reactors. In the following section the types
of reactors used for gas-phase catalysis research and development will be discussed.

A common reactor type is the scouting laboratory reactor with the aim of providing a
means of testing of catalyst activity without interference of effects caused by concentra-
tion and/or temperature. The reactor is typically 6 mm in diameter and held at nearly
constant temperature by a fluidized sand bath or a tube furnace. The narrow reactor di-
ameter maximises the degree of approach to isothermality. Thermocouples are embedded
in the reactor wall and the tube is oriented vertically to avoid bypassing and downward
flow is preferred to avoid fluidisation of the catalyst. The catalyst quantity is small (a few
ml), which allows for rapid testing of multiple catalyst preparations. After approximately

21
50-100 h on-stream it is unusual to witness performance drops that are not due to poison-
ing or fouling effects.

The successful catalyst candidates obtained from the afore-mentioned scouting laboratory
reactor will now have to be prepared and tested in a form that would be satisfactory com-
mercially, e.g. as pellets or extrudates. The reactor for testing these preparations, the
catalyst-optimization reactor, is significantly larger in diameter (25 mm) and longer
(300 to 1,000 mm) than the scouting laboratory reactor. The heating is still performed
externally by electrical heating and in this case the bed is operated closer to industrial-
like conditions. This means that instead of avoiding for instance temperature gradients
they have to be dealt with and incorporated as a part of the final solution; operation at
elevated pressures will have a significant impact on reaction rate and heat evolution in
the reactor but almost no influence on heat transfer, exacerbating gradient issues.

The next step in development is the prototype reactor consisting of a large enough re-
actor section to represent the envisioned commercial reactor. If the reaction appears to be
suitable to perform in a multi-tubular reactor, a single tube is a suitable element for test-
ing. The reasons for performing experiments in this size are several. First of all the high
linear flow rates and other parameters (e.g. tube diameter) used in the catalyst optimiza-
tion reactor will not give the same temperature gradient along the reactor on performance.
Secondly, the pressure drop of the reactor may significantly affect kinetics and/or eco-
nomics, which needs to be determined. Thirdly, the major heat-transfer resistance is
within the bed and the actual gradient should be determined, about half of it is typically
over the catalyst bed and the other half from the bed to the inside tube wall. Finally, the
reproducibility of the pressure drop through the tube is of interest and can be verified by
repacking the tubes multiple times.

A special type of reactor is the gradientless reactors where careful kinetic studies may
be performed at (almost) constant and uniform composition, pressure and temperature.
This can be done in a scouting reactor but due to limitations with respect to heat and mass-
transfer gradients this is rarely accurate enough. These reactors are termed differential
reactors meaning a reactor in which the conversion is limited to not more than a few per
cent. To avoid gradients two main ways are considered, recirculation of the gas through
a stationary catalyst bed or revolving the catalyst in the gas atmosphere, figure 20.

22
Figure 20. An example of a spinning basket gradientless reactor.

Lately the practice of using high-throughput experimentation in the development of cat-


alysts has been implemented in rather large scale. The practice involves preparing a large
number of catalysts (50,000), testing them in parallel using an indicator reaction to re-
move about 90% of the catalyst formulations. The remaining 5,000 catalysts are then
tested again in parallel under more real-like conditions and then traditional research
means are used for developing the final catalyst preparation from the generated active
phases. Figure 21 gives an overview of the work method. The method does however seem
to have some issues as it is best suited for activity and selectivity measurement and ig-
nores important parts of catalysis such as longevity, activation and deactivation.

Figure 21. The inverted pyramid of high-throughput experimentation.

23
3.4 Deactivation
The deactivation of catalysts is an important topic as it is very strongly correlated to per-
formance and economics of a certain process. There is always an initial change in activity
of a catalyst when it is brought on-stream but after 50-100 h this change has occurred and
operating conditions and process hygiene determine the lifetime of the catalyst beyond
that. There are four main ways that a catalyst can be deactivated: poisoning, fouling, sin-
tering and loss of material, figure 22.

Figure 22. Illustration of the various deactivation phenomena, a) fouling, b) poisoning, c) and d) sintering.

In addition to the deactivation pathways stated above, physical deterioration of the sup-
port or carrier may also cause deactivation, e.g. the slow crumbling of the support by
chemical attack or physical grinding of the catalysts in agitated vessels.
3.4.1 Poisoning
The interaction of atoms and molecules with the surface of the catalyst is the main reason
for the function of the catalyst. However as previously discussed the Sabatier principle
states that the interaction (bonding) of the species to the surface has an optimum, i.e. not
too strong or too weak. Catalyst poisons are species that are introduced in the reactor in
minute amounts, usually as feedstock contaminants. These species bind irreversibly to
the active sites and block them for useful, desired catalytic actions. The catalytic cycle
comes to a halt when enough poison has been introduced to block all existing active sites.
The poisons are usually electronegative species such as S, Cl or As and the poisoning
may be reversible or irreversible. If the poisoning is reversible the catalyst may be regen-
erated, e.g. by heat treatment causing the release of the poison from the surface.
24
An example of both reversible and irreversible poisoning is the Ni-based steam reforming
catalysts which are very sensitive to sulphur. The initial formation of surface NiS is re-
versible, while the subsequent formation of bulk NiS is irreversible and after which the
catalyst has to be replaced. Inhibition is a milder form of poisoning and can be used to
avoid certain reaction pathways by blocking parts of the active sites. Figure 23 gives a
recount of common catalyst materials and poisons.

Figure 23. Common active phases and catalyst poisons. Image from Wikimedia commons, modified.

3.4.2 Fouling
The covering of the active phase or plugging of pores by some kind of material, stopping
the reactants from reaching the active phase is termed fouling. This phenomenon may
arise from either deposition of material contaminating the feed, e.g. ash, or it may be
formed during reaction conditions in the reactor, e.g. coke formation on steam reforming
catalysts. Depending on the nature of the fouling, it may or may not be possible to regen-
erate the catalyst. If the fouling is caused by carbon deposition the catalyst may be regen-
erated by using heat and steam, something that is routinely performed with for example
the catalytic reforming catalysts.
3.4.3 Sintering
Sintering is caused by the thermal degradation of the support or the active phase. At high
temperatures the pore system of the support collapses and the size and shape of the active
phase will change. Parts of the active material may also be blocked by the collapsing pore
system; both of these factors will decrease the number of active sites available for reac-
tion.

25
3.4.3.1 Support
As already stated, the influence of temperature on the support causes the support material
to change and at higher temperatures it will start to collapse. The temperature at which
this occurs is naturally dependent on the starting material. The collapse of the pore system
will lead both to a rearranging of the active phase as well as a blocking of the pore system,
making parts of the active phase inaccessible for the reactants. These effects in combina-
tion will certainly lead to a lower activity of the catalyst, due to the decrease in the amount
of active sites available for the reaction. The sintering of a catalyst is best studied by
surface area measurement, e.g. by BET or SAXS.

3.4.3.2 Active phase


The sintering of the active phase is more complex, especially when the active phase is
supported on a metal. The sintering may occur by either atom migration or by crystallite
migration. The criteria for the first one can be described by the Hüttig temperature as
0.26*Tmelting and the second one by the Tammann temperature as 0.52*Tmelting (the abso-
lute temperature is to be used). The migration of the active phase is however dependent
on more than the temperature. The atmosphere (oxidizing, inert or reducing) will influ-
ence the change, as will the presence of certain species in the mixture. The change in
surface area of the active phase may be studied by chemisorption of probe molecules such
as H2 or CO if the active phase is metallic.
3.4.4 Loss of material
The most comprehensible method of catalyst deactivation is the loss of material. The
mechanism here is the volatilization of the active phase and subsequent transport of the
material out of the reactor, or at least away from the catalyst surface. The material may
be deposited in lower temperature areas and in best case scenarios it can be recovered
from these surfaces and reused. This is most common in high temperature applications
such as nitric acid production, but it may also be induced by the formation of volatile
species of the active phase, e.g. the formation of volatile carbonyls species by reaction
with CO. Another way that the loss of material can occur is by alloying of the active phase
with a contaminant in the feed, or indeed by reaction with the carrier. An associated phe-
nomenon is the recrystallization of the active phase from one crystal structure to another
less active one.

3.5 Manufacture
The manufacturing of catalysts has an air of black magic and alchemy to it, due to the
complex and diverse protocols of synthesis. Indeed, the final performance of the catalyst
is dependent on every step in the preparation, as well as the quality of the raw materials.
The synthesis of a catalyst involves many different steps, but on a high level five main
activities can be identified:

• Support preparation
• Active phase preparation

26
• Post-treatment
• Forming
• Activation

These in turn contain several different sub-categories. The main classes of catalysts that
are prepared are bulk catalysts and impregnated catalysts. In the first category, the mixed
metals and metal oxides end up. These are catalysts that are used for ammonia synthesis,
hydrocracking and the zeolites. Impregnated catalysts are mainly used in the case of pre-
cious metals or when instable compounds are used, figure 24. The most important syn-
thesis methods will be described briefly in this section.

Figure 24. The two main classifications of catalysts and some commons synthesis methods for their prep-
aration.

3.5.1 Precipitation and co-precipitation


The most common way of synthesising silica and alumina catalysts is via precipitation or
co-precipitation. This method is used for making several industrially relevant catalysts
such as Cu/ZnO/Al2O3. Precipitation gives high-purity materials and with co-precipita-
tion stoichiometric mixtures may be obtained with well-defined mixed crystallites. The
precipitation is performed in a solvent and typically water-soluble salts are mixed and by
changing the pH, the desired salt is precipitated as a gel. The process has three stages,
super-saturation, nucleation and growth. The resulting gel is aged, filtered, washed, dried
and finally calcined (or heat-treated) to give the final catalyst material.
3.5.2 Hydrothermal synthesis
In this process precipitates, gels or flocculates are heated in the presence of water in an
autoclave at temperatures of 100-300°C. This treatment results in textural and/or struc-

27
tural changes such as crystal and/or particle growth, changes in crystal structure and trans-
formation of amorphous solids to crystalline ones. Zeolites are among the most common
materials synthesised in this way, figure 25.

Figure 25. The synthesis of ZSM-5 by hydrothermal synthesis followed by ion-exchange.

3.5.3 Impregnation of porous supports


This is a common method for preparing supported catalysts such as the reforming catalyst
Pt/Al2O3. In wet impregnation, the porous support is immersed in a solution of the active
phase precursor. By spontaneous adsorption or by precipitation (again triggered by a
change in pH etc.) the active phase is deposited on the catalyst. The resulting catalyst is
filtered, dried and calcined. The drawback of this method is the large volumes of water
required, something that may be overcome by the use of the incipient wetness method.
In this method the solution containing the active phase is added to the dry powder such
that only the internal pores of the catalyst fill up.
3.5.4 Summary of synthesis unit operation
The support phase is usually prepared by precipitation or gelation, while the active phase
is added by impregnation or ion exchange. Post treatment of the mixture includes filtra-
tion, drying and calcination. The finished catalyst powder then has to be formed into a
structure that can be used in a reactor. This is the forming step and it may include pelleting
or extrusion. Finally the catalyst must be activated before being ready to use for a desired
reaction. Please keep in mind that the performance of the final catalyst depends on the
specific operating conditions and process hygiene in each step. Figure 26 summarises the
steps involved in catalyst preparation.

28
Figure 26. The unit operations involved in catalysis synthesis.

Drying is straight forward with respect to crystalline solids, but more complicated for
gels, which may very well contain more than 90% water. The solution is stepwise drying,
a method in which the water is removed without destroying the porous system of the gel.
The gel may then be formed directly into pellets or extrudates or it can be calcined before
forming. Calcination is a drying process at high temperatures (300-800°C) in air or in a
selected atmosphere, which may be inert or reducing. Depending on the starting material,
several things occur during calcination, e.g. decomposition of nitrate or carbamate, crystal
formation, recrystallization and at higher temperature, surface hydroxyl groups are re-
moved.

The next step is forming and in most cases this means pelleting or forming into solid
spheres. The strength and textural properties of these pellets are of uttermost importance
as the catalyst pellets will have to support their own weight. It is also important to avoid
dusting and other phenomena that may occur in fixed or fluidised catalytic bed reactors.
In pelleting or tableting the catalyst powder is compacted with a binder under high pres-
sure. In extrusion, the catalyst is mixed with a binder into a paste which is forced through
a former, figure 27.

29
Figure 27. Examples of extrusion of either extrudates or monoliths.

30
4 Homogeneous catalysis
Per definition the term homogeneous catalysis covers all systems where the catalyst and
the reactants are in the same phase. However, the most common type is the liquid phase
homogeneous catalysis. One example of gas phase catalysis was the classic lead chamber
process which was catalyzed by gaseous NO. The definition of “liquid-phase” in this
context does include one or several of the reactants being in a gas/liquid or vapor/liquid
equilibrium. To exemplify, the hydrogenation of 1-hexene to 1-hexane involves H2 in the
reaction equation. Although the hydrogen is gaseous, part of it dissolves in the liquid and
can thus partake in the catalytic reaction. Although 90% by volume of industrial catalysis
is made up by heterogeneous catalysis, homogeneous catalysis is catching up and indeed
adding much value to the synthesis of fine chemicals. The main problems are in catalyst
separation and recovery, often hindering scientific successes from becoming commercial
ones.

4.1 Structure, components and elementary steps


The term homogeneous catalysis means all reaction systems where the catalyst and reac-
tants are in the same phase. It has, at least to chemists, become synonymous with liquid-
phase reactions catalysed by organometallic complexes. In figure 28 examples of homo-
geneous phase reactions are given. With this said there are also examples of homogeneous
catalysis involving acids or bases and indeed based on organic compounds acting as cat-
alysts (mainly composed of (C, H, O, N, S and P atoms).

Figure 28. Examples of homogeneous phase reactions and catalysts, hydrogenation of 1-hexne and hy-
drofomylation of 1-octene to linear nonanal and branched 2-methyloctanal.

31
Many homogeneous phase catalysts are based on a (transition) metal atom that is stabi-
lised by a ligand. The ligand is usually an organic molecule that attaches to the metal
atom. By changing the ligand, the properties of the catalyst can be altered. Selecting the
right metal and ligand can improve the catalyst activity, selectivity and stability. Common
ligands include triphenylphosphine, phenol and pyridine.
4.1.1 Elementary steps
The elementary steps are the building blocks that can be used for discerning the reaction
mechanism of the catalytic reactions. Understanding the exact nature of the steps involved
in the homogeneous catalysis cycle is often easier than in heterogeneous systems since
the molecular nature of the catalyst simplifies things. The elementary steps are:

• dissociation and coordination


• oxidative addition
• reductive elimination
• insertion and migration
• de-insertion and ß-elimination
• and nucleophilic attack on coordinated substrate

A schematic representation of the main elementary steps of homogeneous catalysis is


given in figure 29.

Figure 29. Elementary steps in homogeneous catalysis.

32
Dissociation and coordination occur in every instance of catalysis. For a reaction to
occur, the substrate has to interact with an active site. This site has to be vacant, which
means that it is ready for reaction. This doesn’t mean that it isn’t occupied by something
currently as the reaction is not performed in vacuum. The solvent molecules, ligands and
substrate in homogenous catalysis compete for the metal coordination sites, similarly to
how molecules are constantly adsorbed and desorbed on a heterogeneous catalyst surface.
By using an excess of solvent and sometime ligands, the probability of having the correct
configuration for reactions increase. In homogeneous catalysis one species can also coor-
dinate to the metal before the other has left. The two situations (first dissociation then
coordination; or first coordination then dissociation) are similar to SN1 and SN2 nucleo-
philic substitution reactions. Examples of the two mechanisms are given in figure 30.

Figure 30. Dissociative and associative ligand exchange from Ni(CO)4.

In oxidative addition, a metal M is inserted into a covalent bond of compound X-Y. The
X-Y bond is broken and two new bonds are formed: M-X and M-Y. The metal loses two
valence electrons and gain two new ligands. This is a key step in many catalytic cycles
and tends to be the rate-limiting step of the reaction. Oxidative reaction can lead to both
trans and cis configuration of the products depending on the compounds added.

In reductive elimination, the opposite too oxidative addition occurs. Starting from a
metal M bonded to X and Y (X-M-Y) the M-X and M-Y bonds break free. This result in
the free species X-Y leaves the complex, the metal loses two ligands and gain two valence
electrons. This reaction is observed mostly with transition metal elements and especially
for noble metals.

33
Insert and migration steps introduce one unsaturated ligand in to another metal-ligand
bond on the same complex. Insertion and reduction elimination are common bond-form-
ing steps, just as oxidative addition is a common bond-breaking step. De-insertion is
simply an insertion in reverse.

Finally the electronic properties of a molecule changes when it coordinates to a metal


centre and this change can activate the substrate toward nucleophilic or electrophilic
attack by another molecule. The first one is more common as it in most cases is the
coordinating molecule that donates electrons to the metal centre.

These reactions aside, there are another handful of important reactions that can be used
for building up a catalytic cycle. However, further description of these is beyond the scope
of this text
4.1.2 Structure
The metal and its immediate environment are the crucial factors that control the catalytic
activity. Important aspects of this environment are controlled much by the ligand size,
flexibility and symmetry. In general the space available for a reactant to interact with the
metal is limited by the size of the ligands and the resulting space is termed the reaction
pocket. Another important component to the reaction cycle is the electronic effects that
ligands, substrates and solvents have on the metal.

4.2 Characterisation
The characterisation of homogenous catalysts is similar or identical to the methods used
in organic chemistry to reveal the structure of a molecule. Such methods may include
nuclear magnetic resonance (NMR) in which nuclei in a magnetic field absorb and re-
emit magnetic radiation. From this information the quantum mechanical magnetically
properties of the nucleus can be discerned giving much information, e.g. the type se-
quence of atoms in species. Other useful methods include other spectroscopy methods
using infrared or ultraviolet/visible light.

4.3 Development and testing


Much of the testing of these systems is performed in constantly stirred tank reactors. The
use of high pressures may be a necessity working with gaseous reactants with low solu-
bility in the reaction medium, e.g. hydrogen. There are many commercial systems avail-
able that satisfy the requirements of this type of investigations. Indeed, once a suitable
catalyst is identified, a major effort goes into designing the catalyst recovery and recy-
cling needed for economically feasible full-scale operation.

The most common way of separating products in the chemical industry is via distillation,
but the thermal sensitivity of the organometallic complexes oftentimes makes this impos-
sible. Most complexes decompose at temperature above 150°C and therefore distillation

34
can only be used in exceptional cases with low boiling-point products. The simplest so-
lution to the temperature problem is a decrease in pressure during distillation. This is
however not problem free either as catalysts optimised to high pressure processes may
undergo undesired transformations at low pressures.

Two main reasons motivate extensive research an investment into catalyst recovery, with
the primary one being the cost of the catalyst. Bulk chemicals usually operate at a margin
of 200-300% (2-3 time mark-up between feedstock and sales price) and since the catalyst
may very well by 1,000 times more expensive than the products any losses of catalyst
will soon render the process uneconomical. Secondly, the presence of the catalyst may
pose a problem in downstream processing, e.g. by changing the selectivity of a subsequent
step, poison another catalyst or cause corrosion as well as contaminate the product.

A commonly used technique to solving this issue is selective product crystallisation. In


this method, the product is crystallised and the catalyst and reactants stay in the liquid
phase. Another common method is catalyst precipitation in which a salt is formed of the
catalyst, forcing it to precipitate. The salt is then filtered off and the catalyst can be re-
generated in a second step. Other techniques include flash distillation under high vacuum
and liquid/liquid extraction. Supercritical solvents and membrane separations are other
separation technologies earning much interest and research these days.

4.4 Deactivation
The deactivation mechanisms of homogeneous phase catalysts differ somewhat from
those of heterogeneous catalysts. The main deactivating effect is that the catalyst reacts
and forms other, less or non-catalytically active, compounds. This may be due to dimeri-
zation, oligomerization or other side-reactions such as loss of ligands, and can be induced
by factors such pH changes, oxidative degradation, increasing temperature or the for-
mation of additional phases. The reaction rate equation will then deviate from ideal rate
laws and the effects will become difficult to foresee. Indeed incorporating the deactiva-
tion into a kinetic scheme is difficult, if not impossible, and the best way to handle the
problem is the most trivial solution: switch to more stable catalysts.

4.5 Manufacture
The manufacture of the catalysts in homogeneous catalysis is based on chemical reactions
in the liquid phase for the most part. The different reactants are added into a solvent and
the temperature and pressure is controlled to yield the desired end-product. The produc-
tion of homogeneous catalysts is more similar to fine chemical synthesis than manufac-
turing of heterogeneous catalysts.

35
5 Conclusions
It can be concluded that catalysis is a central enabling factor in everyday life. A vast
majority of the products and chemical we use have been produced using heterogeneous
or homogeneous catalysts. Catalysts are also monumental in protecting the environment
from unwanted emissions, from industry as well as from point emission sources of con-
sumer goods, e.g. cars. Given the importance of catalysis it is interesting and inspiring to
note that there is still much that is unknown, especially when using heterogeneous catal-
ysis, and that the research in the area never seems to stop. A solution to one riddle, result
in many more that needs dwelling on.

36
6 References
The information in this text has been compiled using the following sources.

Campbell, I.M. (1988) Catalysis at Surfaces. Chapman and Hall, London.

Rothenberg, G. (2008) Catalysis: Concepts and Green Applications. Wiley-VCH Verlag,


Weinheim.

Satterfield, C.N. (1991) Heterogeneous catalysis in industrial practice, 2nd ed. Krieger
Publishing, Malabar.

37

You might also like