You are on page 1of 17

Catalyst Fundamentals

A catalyst is a substance which can increase the rate of a chemical reaction.


Heterogeneous catalysts supported on high surface area porous oxides are used in emission
control applications. The overall catalytic conversion in a heterogeneous catalyst is composed
of several sub-processes which involve chemical reaction, bulk mass transfer and pore
diffusion. Catalytic processes controlled by reaction kinetics can be modeled using the
Arrhenius equation. Catalytic conversions in the mass transfer controlled region can be
estimated using mass transfer correlations developed for monolithic catalyst supports. Even
though catalysts are not used in the reaction, they undergo gradual deterioration due to
thermal deactivation and poisoning.

INTRODUCTION

Chemical Reaction Kinetics and Equilibrium


A chemical reaction which results from the simultaneous combination of a molecules of A, b
molecules of B, and c molecules of C … can be represented by the following equation:

(1)aA + bB + cC … ↔ rR + sS + …

The reaction rate for the forward reaction may be written as

(2)r = k·CAa·CBb·CCc…

where:
r - rate of reaction, kmole/(m3·s)
k - reaction rate constant, 1/s
Ci - concentration of reactant i, kmole/m3.

Equation (2) defines the reaction rate constant (k) which, under ideal conditions, depends
only on temperature and the nature of the reaction. Another important term is the order of
the reaction, defined as the sum of the exponents a, b, c … of Equation (1). This sum equals
to 1 for a first-order reaction, 2 for a second-order reaction, etc.

Some chemical reactions may not proceed completely to the end, i.e., to the right side of
Equation (1). Rather, a thermodynamic equilibrium may be established where certain
concentrations of reactants A, B, C, … will co-exist in the system with certain concentrations
of products R, S, …. From the chemical kinetics point of view, this is represented as a
dynamic balance, where the original reactants A, B, C, … are being re-created in a reverse
reaction between R, S, …. This bi-directional character of the process is indicated by the
double-head arrow in the reaction equation. The rate of the reverse reaction can be written as

(3)r' = k'·CRr·CSs…

The net rate of the reaction is then equal to

(4)r0 = r - r' = k·CAa·CBb·CCc… - k'·CRr·CSs…

The net reaction rate equals zero at equilibrium, r0 = 0, thus

(5)(CRr·CSs…)/(CAa·CBb·CCc…) = k/k' = K

K is the reaction equilibrium constant, which depends on temperature. It can be calculated


from thermodynamic functions of the system. Once known, the equilibrium constant allows
for calculation of the concentrations of reaction products at equilibrium.

Reaction progress can be limited by reaction kinetics or by thermodynamic equilibrium. If


the reaction rate constant (k) is low, the reaction, even if thermodynamically allowed, may
take a long time to complete. Some reactions are so slow that their rates practically equal
zero, and the processes could take years to complete. Fortunately, reaction rates may be
increased by the use of catalysts.

On the other hand, if the reaction reaches its equilibrium, any further reaction progress is
thermodynamically impossible. It is important to realize that catalysts have no effect on the
reaction equilibrium. From the kinetics point of view, if a catalyst increases the forward
reaction rate, Equation (2), it also increases, to the same degree, the reaction rate of the
reverse reaction, Equation (3).

Catalyst Definition
Catalyst is usually defined as a substance which influences the rate of a chemical reaction but
is not one of the original reactants or final products, i.e. it is not consumed or altered in the
reaction. In several known catalytic reaction mechanisms the catalyst forms intermediate
compounds with reactants. Many other catalytic processes are not explained or understood
in their entirety. Neither are the principles governing the selection and preparation of
catalysts for specific purposes. Many of the developments in this field are achieved through
elaborate exploration programs involving trials of countless materials. Catalysts are widely
used in chemical and petrochemical processing to facilitate reactions which otherwise are too
slow, or which require high temperatures to yield good efficiencies. Catalysts are also used to
convert harmful components of engine exhaust gases, such as hydrocarbons, into harmless
substances, such as carbon dioxide and water vapor.

During a chemical reaction the reactants must pass through an energy barrier, called the
activation energy (E), before the final product is produced. The diagram shown in Figure 1 is
an example of the energy path on the way from reactants to products. In this thermodynamic
model, the catalyst provides a path with lower activation energy in the transformation of
reactants to products, which increases the net rate of reaction. The difference between the
energy states of the reactants and products is the heat of reaction. In this example, the
enthalpy of products is lower than that of reactants, indicating an exothermic reaction with
positive heat effect (i.e., with generation of heat during the reaction). In general, catalysts
also can be used for endothermic reactions with negative heat effects.

Figure 1. Reaction Energy Paths

It is important to recognize that a catalyst does not change the net enthalpy and net free
energy of reaction. A catalyst can affect only the rate of approach to the final product state,
i.e., the kinetics of reaching equilibrium. It does not change the energies of the initial and
final states, nor the equilibrium itself. Thus, the reaction enthalpy (ΔH), reaction free energy
(ΔG) and the equilibrium constant (Ke) are not altered by the presence of a catalyst.
Activity and Selectivity
Catalysts can be characterized by two fundamental properties:

 activity, and

 selectivity.

Activity relates to the catalyst’s ability to enhance reaction rates. If two catalysts influence
the rate of the same chemical reaction, the more active catalyst will produce higher reaction
rates.

Selectivity refers to the catalyst’s ability to accelerate certain reactions in preference to


others. This behavior is illustrated by the influence of different catalysts upon the
decomposition of ethyl alcohol [Hougen 1959]. In the presence of an activated alumina (Al2O3)
catalyst, ethylene and water are formed, whereas, with a copper (Cu) catalyst, acetaldehyde
and hydrogen predominate:

(6a)C2H5OH = C2H4 + H2O (Al2O3, 354°C)


(7)C2H5OH = CH3CHO + H2 (Cu)

The selectivity of a catalyst often depends on temperature. A catalyst which is very selective
at a given temperature may lose its selectivity when the temperature is changed. For
example, the decomposition of ethyl alcohol in the presence of alumina follows Equation (6a)
at 354°C, but at a lower temperature of 269°C ether is the principal product according to
reaction Equation (6b), as follows:

(6b)2C2H5OH = (C2H5)2 + H2O (Al2O3, 269°C)

In many processes, the catalyst selectivity is its most important property. A selective catalyst
can accelerate only the desired reactions and minimize the formation of undesired
byproducts which would result if all reactions were equally increased in rate. However, there
is a price for increasing the selectivity of a catalyst. An empirical rule formulated by Sabatier
states that a catalyst selectivity improvement is always accompanied by a deterioration in
its activity [Sabatier 1920]. In other words, modifying a catalyst to be less active towards the
formation of some undesired byproducts (i.e., increasing its selectivity) will always result in
some decline of its activity towards the desired reaction product. The Sabatier rule, although
never proven or universally recognized by the science of catalysis, has been widely confirmed
by the experience with development of selective catalysts for diesel emission control.
Classification of Catalyst Systems
Catalyst systems are divided into homogeneous and heterogeneous catalysts. The
heterogeneous catalysts are further divided into supported and unsupported catalysts. The
following are the definitions of these terms:

 A homogeneous catalyst is a catalyst that is soluble in the reaction medium, e.g., a


water-soluble substance catalyzing water phase reactions.

 In a heterogeneous catalyst system the catalyst is in different phase than the


reactants, e.g., a solid catalyst catalyzing gas phase reactions.

 A supported catalyst consists of a carrier substance having a suitable form and a


large surface upon which the proper catalyst is deposited.

 In an unsupported catalyst, the entire catalyst structure (particle, bead, various


extruded shapes, etc.) consists of the catalytic material.

In most catalytic emission control processes, the gaseous reactants pass through solid
heterogeneous supported catalysts. In most systems, monolithic support structures are
coated with a refractory oxide carrier (e.g. alumina), on to which a noble metal catalyst (e.g.
platinum) is impregnated. The use of unsupported heterogeneous catalysts, usually in the
form of extruded honeycomb monoliths, is limited to a few base metal catalyst applications
(e.g. vanadia/titania SCR catalysts).

Heterogeneous Catalyst
It is assumed that reactions in heterogeneous catalysis occur on the surfaces of the solids at
points of high chemical activity which are known as active centers. The activity of a catalyst
surface is then proportional to the number of active centers per unit area. In many cases the
concentration of active centers is relatively low, as indicated by the extremely small
quantities of "poisons" which are sufficient to destroy the activity of a catalyst. Since the
catalytic reaction rate depends on the number of active sites, much attention is given to
increase the catalyst surface area and to maximize the number of active sites through a fine
dispersion of catalytic components into this surface. To accomplish this, catalytic
components are usually dispersed on high surface area, porous oxide supports (carriers)
including Al2O3, SiO2, TiO2, and SiO2-Al2O3 combinations. The carriers themselves may or
may not be catalytically active, but in either case they do play a major role in maintaining the
overall stability and durability of the finished catalyst.

A model of heterogeneous catalyst is shown using, as an example, alumina (Al 2O3) which is
the most common carrier used for emission control catalysts. Figure 2 presents a few select
pores of a high surface area Al2O3. The diameter of the pores ranges from 20 to 100 Å.
Platinum (Pt) has been deposited into the pores by solution impregnation. The Pt particles or
crystallites are shown as dots.

Figure 2. Catalytic Sites Dispersed on a Carrier

The alumina layer which is bonded to a monolithic honeycomb support is called a washcoat.
Depending on the type of Al2O3 and its thermal history, the internal surface of the washcoat is
rich in surface OH¯ groups (not shown). These OH¯ groups, which cover the entire surface
including the walls of each pore, represent sites upon which one can chemically or physically
bond a catalytic substance.

The physical surface area of the washcoat is the sum of all internal areas of the oxide
including all the walls of each pore. The catalytic components are bound upon these internal
walls and at the OH¯ sites. The catalytic surface area is the sum of all the areas of the active
catalytic components - in this example, platinum. The smaller the crystallite size of the active
catalytic material, the higher the catalytic surface area. It is also assumed that the higher the
catalytic surface area, the higher the rate of reaction for a process controlled by kinetics
rather than by mass transfer.

The washcoat and noble metal systems in today’s catalysts are more complex than that
discussed in the above example. As a rule, the washcoat includes several substances which
act as stabilizers or promoters. In such systems complex interactions may occur among the
washcoat components, as well as between the washcoat and the noble metal(s). These
interactions should be controlled and used to increase the catalyst activity and durability
rather than to cause a loss of activity.
The following interactions between washcoat and catalyst components can be used to
enhance the activity and/or durability of the catalyst system:

 stabilization

 synergism

 segregation.

Stabilizing substances increase the catalyst durability by preventing its deterioration due to
the loss of surface area at elevated temperatures.

Synergism between washcoat components occurs whenever the performance of the mixture
is better than that of each of the components separately. Many synergisms are utilized in
emission control catalysts. An example of a synergistic effect is the interaction between
gamma-alumina and barium oxide, a common washcoat stabilizer. After calcination at
1000°C, the specific surface area of alumina is about 90 m2/g. The corresponding surface
area of barium oxide is less than 1 m2/g. Due to the synergistic effect, a mixture of 3% BaO in
alumina can have a surface area as high as 130 m2/g [Tauster 1997].

Sometimes catalytic components have to be segregated to avoid undesired effects such as


alloying between noble metals or reactions between a noble metal and a washcoat
component. For example, an alloy can be formed between palladium and rhodium in a three-
way automotive catalyst. In a multicomponent washcoat, such alloying metals can be
physically segregated through impregnation on different washcoat species.

CONVERSION RATE IN HETEROGENEOUS CATALYST

The complete reaction process in a heterogeneous catalyst can be divided into several sub-
processes or steps. These sub-processes are of physical (mass transfer) and chemical
(reaction kinetics) nature, each of them with its own rate. The apparent overall rate of the
reaction is always lower than that of its lowest rate step. Depending on the particular system
and physical conditions, any of the sub-processes may be limiting the overall reaction rate.

Lets imagine a gas stream, which contains a reactant (A), flowing through a bed of
heterogeneous catalyst. In the catalyst, the reactant (A) is converted to a product (B), as
shown by Equation (8).

(8)A → B
The complete conversion process of A into B involves the following physical and chemical
steps:

1. Bulk diffusion of reactants. Reactant A must contact the catalytic sites at the outer
surface of the carrier. It must diffuse through the boundary layer of gas in the
proximity of the catalyzed carrier.

2. Pore diffusion of reactants. The majority of A molecules must diffuse through the
porous carrier network toward the internally dispersed active sites.

3. Adsorption of reactants. After the molecule A arrives at the catalytic surface , it must
chemisorb onto a catalytic site.

4. Catalytic reaction. Catalytic reactions, which occur at the active sites, involve
chemisorbed A, energetically located at the peak of the activation energy profile,
which then converts to product B.

5. Desorption of products. After the reaction, product B desorbs from the catalytic sites.

6. Pore diffusion of products. Product B diffuses through the pore network to the
outside surface.

7. Bulk diffusion of products. Having arrived at the outside surface, product B must
diffuse through the boundary layer into the bulk gas.

The above steps can be divided into the following three classes, which are characterized by
different mechanisms and reaction rates:

 Chemical reaction mechanisms include the catalytic reaction itself (step 4), as well as
the chemisorption of reactants and desorption of products (steps 3 and 5,
respectively). The kinetics of all these chemical steps follows an exponential
dependence on temperature; i.e., exp(-E/RT), where E is the activation energy. For
chemical reaction, as well as for chemisorption, E is typically greater than 40
kJ/mole (10 kcal/mole) [Heck 1995].

 Pore diffusion includes steps 2 (reactants) and 6 (products). The rate of these
processes depends primarily on the size and shape of both the pore and the diffusing
reactant and product. The rate of pore diffusion varies approximately with the square
root of k, where k is the rate constant for the chemical conversion reaction. Its
activation energy is approximately half of that of a chemical reaction or about 25-40
kJ/mole (6-10 kcal/mole).
 Bulk mass transfer includes steps 1 (reactants) and 7 (products). The mass transfer
rate in this regime is a function of the specific molecules, the dynamic of the flow
conditions, and the geometric surface area (external) of the catalyst. Bulk molecular
diffusion rates vary approximately with T3/2 and usually have apparent activation
energies of 8-16 kJ/mole (2-4 kcal/mole).

Any of the above 7 steps can be rate limiting and can control the overall rate of reaction in a
heterogeneous catalyst. The rate limiting mechanism can be determined based on the
differences in activation energies for the reaction, pore diffusion, and mass transfer. Figure 3
illustrates the conversion of a gas phase reactant A to product B as the temperature is
increased. At low temperatures, the rate of chemical reaction is slow relative to diffusion.
That area is a reaction kinetics controlled regime. As the temperature increases, the reaction
steps with high activation energies and exponential dependence increase the fastest, and
control of the overall rate shifts from chemical to pore diffusion. Finally, at high
temperatures, both the chemical and pore diffusion rates become sufficiently fast that bulk
mass transfer, having a small relative dependence on temperature, becomes rate limiting.

Figure 3. Regions for Kinetics, Pore Diffusion and Bulk Mass Transfer Reaction Rate
Control

The slope of the conversion curve, Figure 3, can give a qualitative picture of the rate
controlling mechanisms. The lower part of the curve, with fast increasing slope, is indicative
of chemical reaction control (steps 3, 4, or 5). The relatively flat temperature insensitive
portion reflects bulk mass transfer control (steps 1 or 7). The intermediate portion is
characteristic of pore diffusion control (steps 2 or 7).
The chemical reaction steps can be modeled using the following chemical kinetics equations.
Assuming a first order reaction, the reaction rate expression, Equation (2), can be simplified
as follows:

(9)r = k · C

where:
C - concentration of reactant A, kmole/m3

The reaction rate constant (k) increases with temperature, reflecting the fact that reactions
increase their rates at high temperatures. The relationship between k and T is given by the
Arrhenius equation:

(10)k = k0 · exp(-E/RT)

or

(11)ln(k) = ln(k0) - E/(RT)

where:
R - universal gas constant, R = 8.31434 kJ/kmole·K
T - absolute temperature, K.

Experimentally determined reaction rates can be plotted in coordinates of ln(k) vs (1/T), a


representation known as an Arrhenius plot. According to Equation (11), the experimental
data will form a straight line with a slope (-E/R) and intercept of ln(k 0). An Arrhenius plot is
commonly used to determine activation energies of chemical reactions. If reaction rates
corresponding to the data in Figure 3 were plotted in Arrhenius coordinates, the resulting
graph would show differences in activation energies for mass transfer, pore diffusion and
reaction kinetics. The differences would be visible as different slopes for line sections
corresponding to the respective regimes. Obviously, no straight lines would be observed in
transitional areas, e.g., for the transition between the reaction kinetics and pore diffusion
limited regimes.

For reactions in a monolithic, flow-through reactor, one can solve a material balance
equation to determine the relationship between the initial and final concentrations of the
reactants. The basic material balance equation across any reactor, assuming one dimensional
plug flow, steady-state and isothermal operation, is given by Equation (12).

(12)v (dC/dx) = -r
where:
v - velocity, m/s
x - length, m
r - rate of reaction or mass transfer, kmole/(m3·s).

If the process is reaction kinetics controlled and the reaction is of first order, the reaction rate
(r) is given by Equation (9). By combining and integrating the equations, one receives the
following relationship between the inlet (Ci) and outlet (Co) concentrations of reactant A:

(13)ln(Co/Ci) = -kτ

where:
τ - residence time (length of reactor/linear gas velocity), s.

Equation (13) can be used to calculate conversions in a catalytic reactor, which operates in
the reaction kinetics regime, provided the rate constant (k) and the corresponding activation
energy (E) are known. However, there is no universal method to calculate activation energy
for a reaction with a given catalyst. Rather, activation energies of reactions are determined
experimentally for each catalyst.

For a bulk mass transfer controlled process, the rate of mass transfer is given as:

(14)r = KgaC

where:
Kg - mass transfer coefficient, m/s
a - geometric surface area (of the monolith substrate) per unit volume, 1/m.

The relationship between the inlet and outlet concentrations of reactant A takes the following
form:

(15)ln(Co/Ci) = -Kgaτ

The fractional conversion of A can be then calculated as:

(16)(Ci - Co)/Ci = 1 - exp(-Kgaτ)

The mass transfer coefficient (Kg) can be calculated from various correlations developed from
experimental data. One of such correlations for high surface area catalysts on monolithic
substrates [Bonacci 1989] can be expressed as:

(17)NSh = 0.834 (dh/L) NRe0.544 NSc0.56


where:
dh - hydraulic diameter of the honeycomb monolith channel, m
L - honeycomb monolith length, m
NSh - Sherwood number, dimensionless
NRe - Reynolds number, dimensionless
NSc - Schmidt number, dimensionless.

The dimensionless numbers are defined as follows:

(18)NSh = Kga/D
(19)NRe = (W/AF) dh/μ
(20)NSc = μ/ρD

where:
D - diffusivity of the reactant in air, m 2/s
W - total gas mass flow rate, kg/s
μ - gas dynamic viscosity, kg/m·s
AF - open frontal area of the catalyst monolith, m 2
ρ - gas density at operating conditions, kg/m3.

From the above equations, the conversion of a given component can be estimated for a
known monolith geometry and exhaust gas flow. First, the diffusivity of the component (D)
has to be calculated using methods available in the literature (the system can be considered a
binary mixture of a given component and air; methods by Wilke and Lee or by Fuller give
acceptably accurate results [Reid 1987]). Next, NRe and NSc are calculated from Equation (19)
and (20). Honeycomb size parameters for several monolith configurations are listed in the
sections describing ceramic and metallic monolith substrates. Next, NSh is calculated from
Equation (18) and Kg from an experimental correlation, for example from Equation (17).
Finally, the fractional conversion can be calculated from Equation (16).

The accuracy of the overall calculation depends on the accuracy of the correlation for K g
which was used. It should be emphasized that the calculated conversions are valid for the
mass transfer controlled regime. Therefore, they are valid only for sufficiently high exhaust
temperatures. As the temperature decreases, pore diffusion and kinetics start controlling the
process and observed conversions will be lower than those calculated for the mass transfer
region.

Another factor which may impact conversion rates is the pulsation of exhaust gas flow. Most
models neglect pulsation—an approach which is probably justified for converters in
underfloor location. Closed-coupled converters, on the other hand, may be exposed to highly
pulsating flow. Theoretical models showed that in this case neglecting pulsation and
performing calculations based on the average gas flow rate resulted in overestimated
conversion rates.

* CATALYST DEACTIVIATION

Modes of Catalyst Deactivation


Theoretically, catalysts are not used or consumed during catalytic reactions. In all practical
applications, however, catalysts progressively deactivate with use. The following are major
modes of deactivation of monolithic emission control catalysts:

 poisoning and fouling

 thermal deactivation (sintering)

 reactions between catalytic species and the carrier

 washcoat losses.

Catalyst poisoning is one of the most important sources of catalyst deactivation. It occurs
when substances which are present in exhaust gases chemically deactivate the catalytic sites
or cause fouling of the catalytic surface. Poisons in exhaust gases from internal combustion
engines may be derived from lube oil components or from the fuels.

Thermal deactivation is caused by loss of active catalytic sites due to high temperature
sintering of the catalyst or of the carrier.

Interactions between different catalyst species or between catalyst species and carrier
components are another temperature induced mode of catalyst deactivation. An example of
such interactions is the reaction between rhodium and CeO 2 in the automotive three-way
catalyst. This type of problem can be prevented by using alternative carriers and special
washcoat technologies which physically separate the reacting components.

Catalyst deactivation may also occur due to a physical washcoat loss through erosion and
attrition. That mechanism may also be important for emission control catalyst because of the
high gas velocities, temperature changes, and differences in thermal expansion between the
washcoat and substrate materials.
In practical applications, deactivation of a catalyst is usually caused by a combination of
several of the above deactivation modes.

Poisoning
A catalyst poison is a substance which deposits on the surface of the catalyst, rendering it less
active or inactive. A common cause of catalyst deactivation results from fuel- or lube oil-
derived contaminants present in the exhaust gases. There are two basic mechanisms by
which poisoning occurs:

 selective poisoning

 nonselective poisoning.

In the selective poisoning, a catalyst poison directly reacts with the active site or the carrier to
render it less active or completely inactive. In the nonselective poisoning, masking of sites
and pores occurs due to a deposition of fouling agents onto the catalyst carrier. Fouling
results in a performance loss caused by decreased accessibility of reactants to active sites.
Catalyst poisons can be further classified as either temporary or permanent.

Figure 4. Selective Catalyst Poisoning

Selective poisoning is illustrated in Figure 4. In this discriminating process, a poison reacts


directly with an active site, decreasing its activity or selectivity for a given reaction. Some
poisons (e.g., Pb, Hg, Cd) chemically react with the catalytic component (e.g., Pt) to form a
catalytically inactive alloy. Such poisoning leads to permanent catalyst deactivation. Some
poisons merely adsorb (chemisorb) onto sites (e.g., SO2 on Pt) and block that site from
further reaction. These mechanisms are, at least partially, reversible. The catalyst activity can
be restored by desorbing the poison through heat treatment, washing, or simply by removing
the poison from the exhaust gas.
Selective poisoning of the catalyst species results usually in a loss of active catalytic sites with
no change in activation energy, because the remaining sites can function as before. When the
carrier reacts with a constituent in the gas stream to form a new compound (e.g., Al 2(SO4)3),
pores are usually partially blocked, resulting in increased diffusional resistance causing a
decrease in the apparent activation energy.

Figure 5. Catalyst Fouling

Nonselective poisoning (Figure 5) is also called fouling or masking. In this nondiscriminating


process, high molecular weight materials deposit physically on the outer surface of the
catalyst. It has a double effect of decreasing the available catalytic surface area and blocking
access to the pores. In effect fewer active sites are available for reactants and there is an
increase in the diffusional resistance in pores. An example of a masking agent is phosphorus
from the engine lube oil. Phosphorus compounds produce a glaze over the outer washcoat
surface leading to an irreversible loss of catalytic activity.

Thermal Deactivation
Thermal catalyst deactivation, known as sintering, can be attributed to changes in the
structure of the catalytic component (catalyst sintering) or to changes in the structure of the
carrier (washcoat sintering). These processes are influenced by high temperatures, but also
by the nature of the catalytic species, the carrier, and the exhaust gas environment.

Figure 6. Catalyst Sintering


The catalytic component in an active catalyst is highly dispersed on the carrier. It is common
that the fine catalyst crystallites are unstable and grow to better defined crystals. As this
process proceeds, the crystals grow larger. The catalytic surface area decreases, leaving fewer
catalytic sites on the surface of the crystal available to the reactants. Many active sites are
also buried within the crystal with fewer active sites participating in the reaction. Sites may
be altered, affecting catalyst selectivity as well. This mechanism is called catalyst sintering
(Figure 6). Sintering rates of Pt can be reduced by certain rare earths stabilizers (CeO 2,
La2O3) added to the washcoat.

Figure 7. Washcoat Sintering

Washcoat sintering is defined as a loss of the internal pore structure, leading to a decrease in
the physical surface area of the carrier. In the case of Al 2O3 washcoat, sintering is associated
with loss of H2O from the alumina surface. As washcoat sintering occurs, the pore openings
get progressively smaller, introducing greater pore diffusion resistance. Thus, a chemically
controlled reaction may gradually become limited by pore diffusion. The presence of this
phenomenon is again manifested by a progressive decrease in the apparent reaction
activation energy. In extreme cases, the pores close completely, making the catalytic species
within the pore inaccessible to the reactants. The presence of stabilizers such as BaO, La 2O3,
SiO2, and ZrO2 can slow down the rate of sintering in certain carriers. The stabilizers are
believed to form solid solutions with the carrier surface decreasing their surface reactivity.

Another mechanism for carrier change involves conversion to a new crystal structure. For
example, when gamma-Al2O3 converts to delta-Al2O3, there is a significant stepwise decrease
in the internal surface area from about 150 to below 50 m 2/g.

Certain catalytic sites may occur at interfaces between metal (e.g., Pt) and surface support
(e.g., alumina). As both phases sinter, the number and type of interactions changes, affecting
the catalyst performance.
Deactivation Modes and Conversion
Deactivation of catalyst changes the conversion versus temperature curve of a fresh catalyst.
Various possible shifts in the character of the curve are illustrated in Figure 8 [Heck 1995]. The
shifts in the profile can provide insight into the prevailing mechanism of deactivation.

Figure 8. Changes in Conversion for Various Deactivation Modes

The way different deactivation modes affect the catalytic conversion profile corresponds to
the different rate limiting mechanisms in a heterogeneous catalyst, which include reaction
kinetics, pore diffusion and bulk mass transfer.

Loss of active sites, associated with catalyst sintering or selective poisoning, will affect the
reaction kinetics. The activity will shift to higher temperatures, but the maximum conversion
may be unaffected. Catalyst deactivation associated with increased pore diffusion resistance
manifests itself in lower conversions (change in the profile slope) in the pore diffusion
control regime, which limits the process at medium temperatures. Catalyst performance at
low temperature may be unaffected. Masking is a combination of both loss of surface sites
and pore blocking. The conversion is shifted towards higher temperatures with a significant
change in slope.

You might also like