You are on page 1of 13

5.

Catalyst testing

The primary function of a catalyst is to accelerate a chemical reaction. Hence, the


primary objective of testing catalysts is thus the testing of the intrinsic kinetic property,
i.e. its propensity to accelerate the chemical reaction. Mass and heat transfer may
obscure the extent to which a catalyst may accelerate a reaction. Hence, precautionary
measures must be put in place to ensure the absence of mass and heat transfer
limitations, if measurement of the intrinsic catalytic activity is the objective (in some
cases the performance of catalyst pellets needs to be evaluated, i.e. in the presence of
mass transport limitations, in which case the hydrodynamics of the reactor set-up must
be considered carefully). Furthermore, catalyst activity testing can be hampered by
catalyst deactivation, which may differ for various catalysts.

5.1 General definitions


The performance of catalysts is tested experimentally by filling a reactor (vessel) with a
catalyst and adding reactant(s) to the reactor (see Figure 5.1). There are various types
of reactors, viz. continuous reactors, where a continuous flow of reactants is added to
the reactor and a continuous flow of products removed from the reactor, batch reactors,
in which the reactants are added initially to the reactor and the products are removed
after stopping the reaction, and semi-batch reactors, in which a continuous flow of
reactants is added to the reactor and the products are removed at the end of the
reaction.

Continuous flow reactors

System
Flow into Reactor Flow out of
volume
system, FA,0 system, FA

Batch reactors

Reactor Reactor
NA,0 NA

t=0 t=t

Figure 5.1: Continuous flow reactors and batch reactors

Typically, the molar flow rates (F) or number of moles (N) are being determined from
concentration measurements and volumetric flow rates or volume measurements.
Fi  Ci  v

124
Hence, the error in the molar flow rate or number of moles is the error in the
concentration measurement and the error in the volumetric flow rate or volume
measurement. Flow measurement can be replaced by the addition of an internal
standard, which is fed with a constant flow rate with the reactants.
 Ci 
Fi   
  Fint ernal standard
 Cint ernal standard  in a giv en stream

The conversion of a reactant (A) is defined as:


F N
X  1 A  1 A
FA,0 N A,0

The conversion is independent of the units taken, i.e. the conversion in mole-% is
identical to that in weight-%.

The determination of the molar flow rates or number of moles has a great influence on
the accuracy with which the conversion can be determined. If the concentration and the
volumetric flow rate or volume can each be measured with a relative error of 1% (i.e. the
molar flow rates is then determined with a relative error of 2% each), the absolute error
at a conversion level of 10% is 3.6%, i.e. the conversion is somewhere between 6.4 and
13.6%. The absolute error and the relative error in the determination of the conversion
level decreases with increasing conversion levels. Hence, accurate and repeatable
measurements of all variables determining the molar flow rate are required to obtain
reliable conversion levels. The conversion level determined by concentration and
volumetric flow rate/volume determination or with the aid of an internal standard must be
checked with an overall mass balance over the reactor taking into account the mass of
the products formed in the reaction to obtain more confidence in the determined
conversion level.

The amount of product (P) formed is typically related to the amount of reactant fed to the
reactor (yield). The molar yield of a product P is defined as:
F N
YP  P  P
FA,0 N A,0

It should be noted that for many reactions the yield is a function of the units chosen. In a
simple cracking reaction, e.g. C6H12  2 C3H6 the molar yield various between 0 and
200 mole-%, whereas the mass yield varies between 0 and 100 wt.-%. Reporting yields
in weight-% (or equivalent terms such as C-%) is recommended, since the limits of this
value are 0 and 100% as a consequence of the conservation of mass.

The relative error in the yield is directly related to the relative error in the concentration
measurement and the determination of the volumetric flow rates or volume. If the
concentration and the volumetric flow rate or volume can each be measured with a
relative error of 1% (i.e. the molar flow rates is then determined with a relative error of
2% each), the relative error in the yield is 4% (independent of the conversion level).

The selectivity of a reaction is defined as the amount of product formed relative to the
amount reactant converted:

125
FP NP
SP  
FA,0  FA N A,0  N A
Y
SP  P
X
The selectivity is just as the yield therefore also a function of the chosen units.

The relative error associated with the selectivity is now a function of the relative error in
the conversion (which increases with decreasing conversion) and the relative error in the
yield determination. Especially, at low conversions the relative error in the selectivity can
be large. If the concentration and the volumetric flow rate or volume can each be
measured with a relative error of 1% (i.e. the molar flow rates is then determined with a
relative error of 2% each), the relative error at a conversion level of 10% is 36% and the
relative error in the yield is 4%. Thus, the relative error in the selectivity is 40%. Thus, if
a selectivity of 80% is reported at a conversion level of 10%, the selectivity can be
between 48% and 100%. .

The integral rate of reaction for heterogeneously catalysed reactions is defined as the
amount of reactant converted per unit time relative to the mass of catalyst employed.
The rate of reaction is a function of temperature and concentration. The intrinsic rate of
reaction can be measured, if all catalytically active sites are exposed to the same
temperature and concentration:
FA,0  FA 1 NA
 rA   
mcatalyst mcatalyst t
It is assumed here that all sites are exposed to the same temperature and concentration

The turn-over frequency is another measure of catalytic activity describing the number of
molecules of reactant converted per unit time per active site. The turnover frequency can
be calculated knowing the intrinsic rate of reaction and the number of active sites per
unit mass (Nactive sites):
:
N Av
T.O.F.  rA 
Nactiv e.sites

The rate of reaction is dependent on the concentration and temperature prevalent at or


near the catalytically active site. A comparison based on the fundamental rate constant
(and inhibition constants) would present a much better, albeit much less practiced,
approach. This approach would require a sufficient detailed knowledge of the
dependency of the rate of reaction on the concentration of the reactants and product
compounds, which is often either not known or still controversially discussed. Hence, it is
advisable to not only report rate constants (or even turn-over frequencies), but also the
conversion obtained at specified reaction conditions.

5.2 Ideal reactors


Catalysts are tested in reactors, which are classified according to the flow pattern of the
fluid phase (a more exhaustive description of ideal reactors and their flow patterns is
given in [1-3]). Reactors for testing heterogeneous catalysts are have always two or
more phases present within the reactor, viz. a solid phase (the catalyst) and one or more

126
fluid phases (gas and/or liquid). Each of the phases has a different flow profile. The solid
phase typically remains within the reactor, and thus these reactors are batch reactors for
the catalyst (i.e. catalyst performance may change with time).

5.2.1 Batch reactors


The reactor might also be a batch reactor for the fluid phase, i.e. the reactant is initially
added and the product is removed at the end of the reaction. The fluid phase within the
ideal batch reactor is well-mixed. Thus, the concentration and temperature at any
position within the reactor are assumed to be the same. This is a critical assumption,
which requires some further exploration. However, this question has not been explored
to a large extent in literature. In heterogeneously catalyzed reactions, the conversion of
the reagent(s) and the formation of product compound(s) take place at defined places
within the reactor. Mixing of the fluid element around the catalyst particle with the rest of
the fluid will result in the equalisation of the spatial concentration gradients. The
important parameters affecting the spatial variation in the concentration of the reactants
and product compounds are the rate of the chemical reaction, the density of reaction
sites, and the mixing time. A minimal spatial variation might be expected when the time
constant for the chemical reaction is much larger than the mixing time, i.e. the rate of
chemical reaction is much slower than the rate at which the mixing process takes place.
The reciprocal value of a (pseudo) first order rate constant can be taken as the time
constant for the chemical reaction. The mixing time (mix) can be defined as [4]:
2
cons tan t 1  dimpeller 

 mix   
N 1  d 
 reactor 
P03
with N: stirrer speed (rpm)
P
P0: Power number P0 
3 5
 f luid  N  dimpeller
P: Power input (W)
dimpeller: diameter of impeller (m)
dreactor: diameter of reactor (m)
(a typical value for the constant is 5; hence, the mixing time in a vessel of a diameter of
0.2 m filled with water ( = 1000 kg/m3) stirred with 600 rpm (or 10 rotations per second)
and a power input of 500 W with an impeller with a diameter of 0.1 m, amounts to ca. 0.9
s).

A useful criterion to assess the possible occurrence of spatial concentration (and thus
also temperature) variations might be:
mix 
 mix   rA   1
reaction C A
(hence, only for reactions with a observed 1st order rate constant of much smaller than
0.1 s-1 can be the vessel described above be considered an ideal batch reactor).

The rate of reaction in a batch reactor can be obtained from a temporal mass balance
around the reactor loaded with mcatalyst mass of catalyst yields:
dN A
  rA   mcataly st
dt

127
In a constant volume batch reactor (e.g. carrying out the reaction in an autoclave or for
liquid phase reactions), the rate of reaction can be related to the change in the
concentration of the reactant by:
dC A mcataly st
  rA  
dt Vf luid.phase
A measurement of the change in the concentration as a function of time is thus required
to obtain the rate of reaction. Figure 5.2 shows schematically how to obtain the rate of
reaction from the measured concentration as a function of time data. The rate of reaction
typically declines with reaction time due to depletion of the reactant (unless the reaction
order is zero or less with respect to the reactant).

1
concentration of reactant,

0.8
CA(t)/CA(t=0)

0.6

0.4

0.2
rate =-dCA/dt

0
time
rate

rate

time 0 0.2 0.4 0.6 0.8 1


concentration of reactant, CA(t)/CA(t=0)

Figure 5.2: Obtaining activity data (rate of reaction) from batch reactor experiments

The obvious advantage of batch reaction studies is the amount of data gathered in a
single run, which can be used for a kinetic analysis. The rate of reaction or the change in
the concentration with time can be monitored as a function of the concentration of the
reactants and products. The reaction network can be elucidated from the concentration
and selectivity data as a function of reaction time of a single experiment. However,
assessment of the catalyst performance after a fixed reaction time severely limits the
validity of the test, due to the variation in the conversion (see 5.3).

Furthermore, the starting time of the reaction is for some reactions difficult to define,
especially for reactions occurring at elevated temperatures and/or pressures (see e.g. [5]

128
and Figure 5.3). The heating-up phase is of particular importance if the initial stage of
the reaction needs to be monitored. The reaction does already start (albeit at a slower
rate) at temperatures below the reaction temperature. For instance, the rate of reaction
at 175oC is only 4 times smaller than the rate of reaction at 200oC, if the reaction occurs
with an activation energy of 100 kJ/mol. An additional disadvantage of the batch reactor
is the difficulty to assess catalyst deactivation, due to the inherent change in the
measured activity with time.

250 100
Monoethylamine
(MEA)
200 80
Temperature, oC

Samines, C-%
Treaction-25oC
150 60

100 40
Diethylamine (DEA)
50 20 Triethylamine
(TEA)
0 0
t = 4 min 0 10 20 30 40 50
0 30 60 90
Time, min Ethanol conversion, XEtOH, mole-%

Figure 5.3: Ethanol ammination in a batch reactor (Vreactor = 600 cm3; catalyst (10 g):
10 wt.-% Co/SiO2; T = 473.15 K, nC2H5OH,initial = 1.88 mol; nNH3,initial = 0.93
mol, nH2,initial = 0 mol) [7]
Left: Heating up a batch reactor
Right: Selectivity of the liquid phase ethanol ammination as a function of
ethanol conversion

The intrinsic kinetics can be masked by internal and external mass and heat transfer
limitations. The internal mass and heat transfer limitations (in the absence of external
mass and heat transfer limitations) can be assessed by performing the reactions with
catalyst particles of a different size or size range.

The assessment of external mass and heat transfer limitations can be problematic. THe
relative velocity of the fluid to the solid particle must be varied in order to reduce the
fluid-solid mass and heat transfer across the boundary layer surrounding the solid
catalyst particle. Batch reactors are often stirred, and increasing the stirrer speed will
increase the average velocity of the fluid. However, if the density difference between the
fluid and the solid particle is minimal, as for instance in the case of heterogeneously
catalysed liquid phase reactions, the relative velocity of the fluid to the solid catalyst
particle will not change much with changing stirrer speed. The introduction of baffles in
the reactor may reduce the extent of external mass and heat transfer limitations in this
case. Improved external mass transfer may be obtained by adhering the solid catalyst
particles to the impeller, e.g. in a basket attached to the impeller forcing the liquid
through the catalyst bed.

5.2.1 Packed bed reactors


In a packed bed reactor the fluid is flowing over the catalyst bed. This is quite a
commonly used reactor type for catalyst testing due to its simplicity in operation. The
packed bed reactor acts like a batch reactor for the solid catalyst particles and is
typically assumed to operate as a plug flow reactor for the fluid.

129
The characteristics of the ideal plug flow reactor are the absence of any radial
concentration and temperature gradient. The concentrations of the various reactant and
product compounds do vary along the axial position of the bed. Hence, a single
measurement at a given flow rate will yield a single data point (conversion/selectivity)
unless the reactor set-up has been designed with sampling points along the catalyst
bed.

In a fixed bed reactor, catalyst pellets are packed closely to each other introducing a
resistance to the fluid flow resulting in a pressure drop along the catalyst bed (typically
given by the Ergun equation). A minimal pressure drop over a catalyst bed in the case of
heterogeneously catalyzed reactions involving a gas phase is desired since the pressure
drop will affect the intrinsic kinetics for these reactions.

The reactor wall also affects the pressure drop over the catalyst bed [6,7]. The presence
of catalyst near the reactor wall results in a large void surrounding these catalyst pellets,
which can be as large as half of the pellet diameter [8,9]. At high Reynolds numbers,
when the mass transfer boundary layer is thin, the increased local porosity increases the
local fluid velocity and hence decreases pressure drop over the catalyst bed. However at
low Reynolds number, the additional friction introduced by the reactor wall may reduce
the local velocity and thereby increasing the pressure drop across the catalyst bed. The
extent of the reactor wall effect is expected to be a function of the ratio of reactor to
particle diameter (Dreactor/dpellet), since Reynolds number is a function of pellet diameter, A
correlation containing over 2300 experimental point [10] shows that the reactor wall
effect on pressure drop across the catalyst bed becomes negligible when Dreactor/dpellet is
greater than 15 for cylindrical pellets.

Plug flow behaviour implies that there is no mixing along the axial direction of the
catalyst bed, i.e. the flow of the reactants/products is uni-directional. This is only possible
in theory, since some degree of dispersion will always be present. A low dispersion
number typically indicates a good approximation to plug flow behaviour. The extent of
dispersion is a function of particle size in relation to the length of the reactor and the
velocity of the fluid through the catalyst bed [1, 11-13] (see Figure 5.4). Hence, a longer
catalyst bed would be required for small catalyst particles than for large catalyst particles

100
Dispersion number

10 Axial dispersion

0.1
Radial dispersion

0.01
0.01 0.1 1 10 100
Reynolds number

Figure 5.4: Axial and radial dispersion as a function in packed bed reactors [11,12]
(calculated for spherical particles; Sc = 0.9, bed = 0.4)

130
operating at the same linear velocity to ensure plug flow behaviour. Investigating the
intrinsic kinetic behaviour of the catalyst however requires utilization of small catalyst
particles. The opposite requirements can be circumvented by utilizing a diluent of a
given (large) particle size and introducing a pre-heating zone in the reactor [13].

The presence of internal and external mass and heat transfer effects can be easily
investigated using a packed bed reactor. The former can be investigated by measuring
the activity of a catalyst as a function of the particle size used in the study (keeping the
linear velocity and the space velocity constant). The presence of external mass and heat
transfer limitations can be investigated by changing the linear velocity of the fluid flow
over the catalyst bed keeping the space velocity constant. This requires a set of
experiments in which the flow rate of the fluid flow is systematically increased with a
corresponding increase in the catalyst mass to keep the space velocity constant.

The kinetic analysis of the performance of a catalyst typically requires a set of


experiments, in which the space velocity of the reactant is changed. The observed
change in the conversion is then related to the catalyst activity. A packed bed reactor is
an integral reactor, which means that the catalyst particles experience a whole range of
concentrations (from the inlet concentration to the outlet concentration). Strictly
speaking, the evaluation of the kinetics of a catalyst in a packed bed reactor requires the
same approach as the approach in a batch reactor. The rate of reaction in a packed bed
reactor can be obtained from::
dX A
  rA 
 mcatalyst 
d 
 F 
 A,0 
Hence the variation in the conversion with reciprocal value of the space velocity yields
the rate of reaction, which must be related to the average concentration change in this
space velocity interval.

Two alternative approached can be taken in order to simplify the kinetic evaluation of
catalysts in a packed bed, viz. the integral kinetic evaluation and the differential kinetic
evaluation. In the integral kinetic evaluation a rate expression is assumed yielding an
expression of the outlet concentration or conversion as a function of space velocity by
integrating the mole balance for the reactant A over the bed length. For example
assuming a first order rate expression:
dX A
  rA   k  C A
 m cataly st 
d 
 F 
 A,0 
And further assuming that the contraction/expansion due to the reaction is minimal (e.g.
for equimolar reactions, reactions with a highly diluted reactant or reactions in the liquid
phase):
dX A
 k  C A,0  1  X A 
 mcataly st 
d 
 F 
 A ,0 
C A,0
ln1  X A   k   mcataly st
FA,0

131
mcataly st
ln1  X A   k 
v inlet

The differential approach presumes that the differential can be expressed as the
difference in the conversion as a function of the reciprocal value of the space velocity
(also known as the space time):
dX A X A X A,outlet  X A,inlet X A,outlet
     rA  This approach
 m cataly st   m cataly st   m cataly st   m cataly st 
d       
 F   F   F   F 
 A,0   A,0   A,0   A,0 
presumes that the change in the concentration of the reactants/products over the
catalyst bed does not affect the rate of reaction. The typical question that arises is up to
what level of conversion is this approach valid. This approach is valid for zero order
reactions at any level of conversion, since the rate of reaction is not a function of the
concentration of the reactants/products and thus independent of the level of conversion.
An error will be introduced for reactions with reaction orders differing from zero. Some
level of uncertainty can be allowed in the kinetic data. It can be shown that this approach
leads to an error in the rate constant of less than 10% up to a conversion of 20% for 1st
order reactions. The error becomes more severe for reactions with a higher reaction
order.

5.2.2 Continuous stirred tank reactors (CSTR)


A number of reactors can behave as reactors, in which the catalyst is exposed to a
single concentration of the reactants. This can be achieved in a fixed bed by minimizing
the level of conversion (the so-called differential reactor). Alternatively, the fluid may be
recycled over the catalyst bed, so that the conversion per pass is minimal. This can be
achieved for instance in a Berty reactor [14], in a jet-loop reactor [15] or in a slurry
reactor.

The reaction conditions in a CSTR are given by the exit concentration of the components
in the fluid phase, and thus the catalyst particles experience (in the ideal case) only
these conditions. This is in contrast to a packed bed reactor, where the catalyst particles
experience all concentrations from the inlet concentration to the exit concentration.
Hence, CSTR-type of reactors are ideally suited for measuring the rate of reaction which
is given by
XA
  rA 
 m cataly st 
 
 F 
 A ,0 
Although these type of reactors are ideally for determining the rate at a particular
concentration of reactants (and products), performing kinetic measurements in this type
of reactors can be difficult since exact control over the concentration of the relevant
compounds is difficult to achieve.

5.3 Catalyst comparison


The most obvious observable in catalyst testing is the decline in the amount of reactant
present in the reactor or reactor effluent, i.e. the conversion of the reactant. Hence, a
comparison of catalyst activity based on direct measurement of the effluent composition
is tempting and often the only possible route forward, if more detailed knowledge

132
regarding the kinetics and/or the nature of the active sites is unknown. However, in these
situations some basic aspects should be kept in mind.

Catalysts should not be compared under conditions yielding conversion levels of 100%
or even close to 100%. Under those conditions uncertainly exists about how many of the
active sites are actually involved in the catalytic reactions. Figure 5.5 depicts two
situations with 100% conversion; in one case 99% conversion of the reactant is obtained
after 60% of the catalyst bed, whereas the other case shows a situation where 99%
conversion is obtained after 30% of the catalyst bed. Measurement of the catalytic
activity based on the effluent composition would rank both catalyst equally active (100%
conversion), whereas in reality catalyst B is more active than catalyst A.

CA,0 CA,0=0 CA,0 CA,0=0


Catalyst A Catalyst B
XA=100% XA=100%
CA/CA,0

CA/CA,0
Bed length Bed length

Figure 5.5: Comparison of the concentration profile of the reactant A over two
catalysts with different activity resulting in both cases to 100% conversion

The accuracy with which the conversion level can be determined than also plays an
important role in the ranking of the catalyst. A better distinction in the level of conversion
is obtained by testing catalyst at between 40 and 60% conversion (see Figure 5.6). The
ultimate difference in the conversion obtained over catalysts with different activity is a
function of the level of conversion, the reaction kinetics and the difference in catalytic
activity. Hence, comparison of catalytic activity must be interpreted carefully.
Furthermore, discrimination between catalyst differing in catalyst activity less than 20%
may become experimentally challenging.

In the same vein, catalysts for equilibrium limited reactions should not be tested close to
the equilibrium conversion level (hence, it is important to establish a priori the position of
the equilibrium conversion). It should be realised that close to equilibrium the difference
between catalysts of varying activity vanishes, since catalysts catalyze both the forward
and the reverse reaction to the same extent. Hence, the difference between the catalysts
in terms of obtained conversion level becomes indistinguishable.

A comparison of catalytic activity on the basis of conversion for deactivating catalysts is


problematic, since the conversion over the catalysts change with time on line. Figure 5.4
show the obtained conversion as a function of time on stream for two different,
deactivating catalysts. A correlation between the activity of the catalyst and the rate of
deactivation can be expected for some types of deactivation, i.e. an initially highly active
catalyst may deactivate faster. The time span over which the activity is compared
becomes important here as well. As demonstrated in Figure 5.7, catalyst A shows
superior activity after relatively short time on line (e.g. t 1) as typically measured in
laboratories. Comparing the activity however at time t2 results in the reverse conclusion,
i.e. catalyst B having a superior activity to catalyst A. The choice of the “best” catalyst is
now strongly dependent on the possible engineering solutions to deal with a deactivating

133
catalyst. However, it should be kept in mind that a stable operation (i.e. dealing with a
less deactivating catalyst) is from an engineering point of view often more desirable, i.e.
a lower rate of deactivation is preferable over a high activity.
50

Difference in conversion
between less active and
more active catalyst, %
0th order reaction
40

30
1st order reaction
20
2nd order reaction
10

0
0 20 40 60 80 100
Conversion with less active catalyst, XA, %

Figure 5.6: Difference in the conversion level obtained with a catalyst twice as active
as the less active catalyst (dotted line for 1st order reaction with catalyst
only 20% more active than the less active catalyst) as a function of the
conversion obtained over the less active catalyst in a plug flow reactor or
a batch reactor

100

80
Conversion, XA, %

60 Catalyst B

40
Catalyst A

20

0
t1 t2 end-of-life
Time on line
Figure 5.7: Comparison of deactivating catalysts.

Selectivity is often of greater importance than activity. Selectivity is a measure of the


amount of product formed relative to the amount of reactant converted, and hence a
proper comparison of the selectivity is of great industrial value. However, the comparison
of selectivity obtained over catalysts with different activity is fraught with difficulty, since
the route of product formation must be kept in mind when comparing catalyst selectivity.
The selectivity of a reaction can be strongly dependent on the conversion level, if
secondary reactions do take place (see Figure 5.8). Thus, strictly speaking selectivity of
a reaction over a variety of catalysts can only be compared at constant conversion level.

134
Reactions in parallel Consecutive reactions

Selectivity for the formation of product P, Sp, %


100 100
80 80
60 k1 = 2 60
40
P 40
A k1 = 2 k2 = 1
20
k2 = 1 20 A P S
S
0 0
0 20 40 60 80 100 0 20 40 60 80 100

100 100
k1 = 2
80 80
Ak =2
P k1 = 2 k2 = 1
60 -1 60
Ak =2
P S
40 k2 = 1 40 -1

20
A S 20
0 0
0 20 40 60 80 100 0 20 40 60 80 100

100 100
80 80
60 k1 = 2 60
40 A P 40 k1 = 2 k2 = 1
k2 = 1 A P S
20 A S 20 k-2 = 1
k-2 = 1
0 0
0 20 40 60 80 100 0 20 40 60 80 100
100 100
k1 = 2
80 Ak =2
P 80
-1
60 60
40 k2 = 1 40 k1 = 2 k2 = 1
20 A S 20 Ak =2
P k-2 = 1
S
k-2 = 1 -1
0 0
0 20 40 60 80 100 0 20 40 60 80 100

Series-parallel type of reactions


100
80
60
k1 = 2
40 P
k3 = 1
20 A k2 = 1
0 S
0 20 40 60 80 100

Conversion of reactant A, XA, %


Figure 5.8: Selectivity of a variety of 1st order reactions as a function of conversion
(solid lines: batch or plug flow reactor; dotted lines: CSTR; graph adapted
from [1])

135
Reactions, which only show primary reactions, i.e. reactions for which the secondary
conversion of the product compounds are slow, can be compared at all levels of
conversion. The occurrence of secondary reactions to an appreciable extent may
complicate comparison of catalysts based on selectivity measurements significantly.
This has been in particular noticed in the selective oxidation of the rather unreactive n-
paraffins, where high selectivities have been reported at very low level of conversion (i.e.
negligible extent of secondary reactions), but poor selectivities at high conversion (see
e.g. [2]).

References:
[1] O. Levenspiel, „Chemical Reaction Engineering“, 2nd ed., John Wiley&Sons, New
York (1972).
[2] H.S. Fogler, “Elements of Chemical Reaction Engineering“, 3rd ed., Prentice Hall,
New Jersey (1999).
[3] M.E. Davis, R.J. Davis, ‘Fundamentals of chemical reaction engineering’,
McGraw-Hill, Boston (2003).
[4] A.W. Nienow, Chem. Eng. Sci. 52 (1997), 2557.
[5] L.D. Biquiza, MSc-thesis, University of Cape Town (2002).
[6] P.C. Carman, Trans. Inst. Chem. Eng. (London) 15 (1937), 150.
[7] D.A. Nield, AIChE J. 29 (1983), 688.
[8] L.H.S. Roblee, R.M. Baird, J.W. Tierney, AIChE J. 4 (1958), 460.
[9] R.F. Benenati, C.B. Brosilow, AIChE J. 8 (1962), 359.
[10] B. Eisfeld, K. Schitzlein, Chem. Eng. Sci. 56 (2001), 4321.
[11] D.J. Gunn, Chem. Eng. Proc. 32 (1993), 333
[12] D.J. Gunn, Chem. Eng. Sci. 42 (1987), 363
[13] R.J. Berger, J. Pérez-Ramirez, F. Kapteijn, J.A. Moulijn, Appl. Catal. A: General
227 (2002), 321
[14] J.M. Berty, Catal. Chem. Eng. Prog. 70 (1974), 78.
R. Brosius, J.C.Q. Fletcher, Chem. Eng. J. 161 (2010), 196.
[15] K.P. Moller, J.C.Q. Fletcher, C.T. O’Connor, R.J. Becker, Chem. Eng. Commun.
137 (1985), 111

136

You might also like