Sunteți pe pagina 1din 119

.

ADVANCES IN CATALYSIS VOLUME 28

The Kinetics of Some Industrial


Heterogeneous Catalytic
Reactions
M . I . TEMKIN
Karpov Institute of Physical Chemistry
Moscow. USSR
I.Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 173
I1.The Measurement of Reaction Rate . . . . . . . . . . . . . . . . 174
111. Mass Transfer in Heterogeneous Catalysis . . . . . . . . . . . . . 178
IV .Ideal Adsorbed Layers . . . . . . . . . . . . . . . . . . . . . . 184
V. The Routes of Complex Reaction . . . . . . . . . . . . . . . . . 188
VI .The Steady-State Conditions . . . . . . . . . . . . . . . . . . . 192
VII .Some General Relations for Steady-State Reactions . . . . . . . . . 197
VIII .Reversible Many-Stage Reactions . . . . . . . . . . . . . . . . . 203
IX. Nonuniform Surfaces . . . . . . . . . . . . . . . . . . . . . . 207
X. Adsorption Equilibrium and the Kinetics of Reaching it on
Nonuniform Surfaces . . . . . . . . . . . . . . . . . . . . . . 213
XI . The Kinetics of Reactions on Nonuniform Surfaces . . . . . . . . . 223
XI1. Oxidation of Ethylene into Ethylene Oxide . . . . . . . . . . . . . 230
XI11. Hydroxylamine Synthesis . . . . . . . . . . . . . . . . . . . . . 239
XIV . Reaction of Methane with Steam . . . . . . . . . . . . . . . . . 244
XV . Ammonia Synthesis-Simple Kinetics . . . . . . . . . . . . . . . 250
XVI . Ammonia Synthesis-Complicated Kinetics . . . . . . . . . . . . . 257
XVII . Carbon Monoxide Conversion . . . . . . . . . . . . . . . . . . 263
XVIII . Isotopic Exchange between Steam and Hydrogen . . . . . . . . . . 267
XIX . Phosgene Synthesis . . . . . . . . . . . . . . . . . . . . . . . 270
X X . Reactions of Carbon with Carbon Dioxide andsteam . . . . . . . . 273
XXI . Ammonia Oxidation . . . . . . . . . . . . . . . . . . . . . . . 279
References . . . . . . . . . . . . . . . . . . . . . . . . . . . 287

I . Introduction

The phenomenon of catalysis is a subject of chemical kinetics. as follows


from Ostwalds definition of catalysis. Therefore. the accumulation of
data on the kinetics of concrete catalytic reactions favors the progress of the
theory of catalysis.
It is expedient to choose for kinetic studies the reactions of industrial
importance since the kinetics of these reactions are not only a source of in-
173 Copyright 0 1979 by Academic Press. Inc .
All rights of reproduction in any form reserved.
ISBN 0-12-007828-7
174 M. I. TEMKIN

formation for the theory, but also find immediate practical applications in
process optimization and reactor design. The calculations of this kind are
performed now with electronic computers on a large scale.
This chapter presents an account of the main results of studies of kinetics
of some industrial heterogeneous catalytic reactions. The studies have been
carried out by the author with his co-workers at the Karpov Institute of
Physical Chemistry (Moscow, USSR). The presentation is not chrono-
logical ; the reactions are arranged based on the character of interpretation
of their kinetics.
For compactness and to avoid repetition, some general aspects of experi-
mental methods and theoretical principles developed in the course of our
investigations are formulated prior to the discussion of individual reactions.
A more detailed analysis of many of the subjects dealt with in this part of
the chapter can be found in the contributions by Kiperman (1) and Snagov-
skii and Ostrovskii (2).
The limits of this article did not allow inclusion of the results of the com-
parison of kinetic equations with experimental data. They are given in the
referenced original papers.

II. The Measurement of Reaction Rate

The rate of a reaction is its intensity expressed quantitatively. Let a reac-


tion be described by a chemical equation of the form
aA + aA = bB + bB (1)
where A and A are reactants, B and B are products, and a, a, b, b are
stoichiometric coefficients. If a molecules of substance A and a molecules
of substance A have reacted and formed b molecules of substance B and
b molecules of substance B, we shall say that there has been one run of the
reaction along (1). The total number of reaction runs divided by the Avo-
gadro number ( N = 6.02 x mol-) is called the extent of reaction
(3); this variable, which was introduced into chemical thermodynamics
by de Donder, is usually denoted as 5. The extent of the reaction, as the
amount of matter, is measured in moles.
Some authors state that the reaction rate is dt/dt where t stands for time.
But dt;/dt is proportional to the size of the reactor and, hence, is an extensive
property like t, and not an intensive property, as should be the reaction
rate, according to the definition of the term. The derivative dt/dt is to be
called the reactor productivity, but not the reaction rate.
In general, when reaction rate, Y, varies with time and is not the same in
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 175

different parts of the reaction space, it is defined as follows. If the reac-


tion is homogeneous
r = a2gat a0 (2)
where u is the volume in which the reaction occurs. For a heterogeneous reac-
tion
r = a2t/at as (3)
where s is the area of the surface on which the reaction occurs. Thus the rate
of homogeneous reaction is the density of the reactor productivity, and that
of heterogeneous reaction is the surface density of the reactor productivity.
It is more convenient for theoretical considerations to define the reaction
rate on the basis of the number of its runs without dividing the number by N .
Engineering calculations often need the rate of heterogeneous catalytic
reaction to be referred not to the surface, but to the mass of the catalyst or
the volume of the bed of the catalyst grains. The relation between different
expressions of a heterogeneous catalytic reaction rate is determined by the
values of specific surface area, o, of the catalyst and the bulk density, P b ,
of the catalyst bed. The total surface of the catalyst, s = om, where m is
the mass of the catalyst; m = PbU, where u is the volume of the bed.
The notion reaction rate will be discussed more completely later (see
Section VI), including the cases when the reaction simultaneously follows
several pathways with the formation of different products.
In most of the investigations described below the reaction rates were
measured by the circulation flow method proposed in 1950 (4). This method
offers a possibility of realizing a steady-state heterogeneous catalystic reac-
tion without any concentration and temperature gradients; i.e., it belongs
to the group of methods which were called nongradient (5). The scheme
of the circulation flow system is shown in Fig. 1.
The apparatus for measurements at atmospheric pressure is made of
glass. The mixture of gaseous reactants is fed into the system at a constant
rate through pipe 1. The converted gas mixture flows out through pipe 2 ; it
is analyzed, e.g., chromatographically. Pump 3 provides for the intense
circulation of the mixture. The hollow glass piston of the pump is packed
with iron and is actuated by an electric magnet; the valves of the pump are
arranged in such a way that both strokes of the piston are utilized. Reactor
4 (containing catalyst grains) is placed in an oven. Before contacting the
catalyst, the gas mixture is heated in a coil also immersed in the oven. Cir-
culation rate (i.e., the velocity of the passing of the gas mixture through
reactor 4) is many times higher than the flow rate (the velocity of the gas
mixture in pipes 1 and 2).
176 M. I. TEMKIN

- - II

FIG.1. Circulation flow system.

Some time after the beginning of the operation, the system comes to a
steady state and the composition of the outgoing gas mixture in pipe 2 ceases
to change. In essence, it is sufficient to pass through the system a volume of
the gas mixture 5 times that of the system. After that the reaction continues
indefinitely under constant conditions. Because of the large circulation rate,
the degree of conversion at each single passage of the gas mixture through
the reactor is small; therefore, the composition and temperature of the
mixture at the reactor inlet and outlet are almost the same. Conversely, the
composition of the mixture at the cycle inlet and outlet can differ consider-
ably; the overall degree of conversion possibly being high since the gas mix-
ture repeatedly passes through the reactor and is renewed to only a small
extent by the feedstock.
It has already been mentioned that the experiments of this sort realize a
practically nongradient course of reaction. Specifically, large-scale non-
gradient conditions exist in the reactor, viz., the composition of the gas mix-
ture and its temperature in the space between catalyst grains are virtually
identical in the whole volume of the reactor. But this does not preclude
considerable temperature and concentration gradients directed from the
surface to the center of the catalyst grain. These gradients can be eliminated,
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 177

however, if sufficient small grains are used. On the other hand, experiments
with grains of the size which is used in industry can provide data on the effect
of diffusionin the pores of the catalyst and heat transfer in the grains on the
course of reaction under the industrial conditions.
The measurements with small grains characterize a reaction in the kinetic
region. The composition of the gas mixture flowing out of the cycle, which is
determined by analysis, is just the composition of the reaction medium.
The reaction rate is easily calculated from the same results of analysis; the
extent of the reaction, t, during some time, t , is equal to the increment of the
amount of any product or to the decrement of the amount of any reactant
when the gas mixture has passed through the system divided by the respective
stoichiometric coefficient. Since the reaction is steady state, 5 is proportional
to t ; and since the reaction proceeds under nongradient conditions, 5 is
proportional to the surface of the catalyst, s. As a result, the general Eq. (3)
reduces to
r = lyts. (4)
If the reaction is conducted in a closed system and an identical course of
the reaction on the whole surface of the catalyst is achieved, e.g., by circula-
tion, then
r = (l/s)(d(/dt). (5)
If the reaction is steady state and carried out in a flow reactor, then
r = (l/t)(d</ds). (6)
The comparison of (5) and (6) with (4)demonstrates that the reaction rate
in the circulation flow system is found in the simplest way.
By changing flow rate and the composition of the gas mixture at the inlet
of the cycle, we vary the composition of the reacting gas mixture and deter-
mine the corresponding variation in reaction rate. We may also add dif-
ferent amounts of the reaction products, inert gases, or catalytic poisons to
the initial gas mixture.
The advantages of the circulation flow system over static or flow systems
become particularly essential in the studies of the reactions occurring simul-
taneously along several pathways with the formation of a variety of products.
The circulation flow method is applicable for the studies of almost any
heterogeneous catalytic reaction. At high pressures steel equipment is used
instead of glass. The first circulation flow reactor for high pressures was
designed by Sidorov (6);the gas mixture is circulated in this reactor by means
of steel bellows that are actuated by a rod introduced into the reactor also
through bellows, without any packing.
After the circulation flow method had been developed, various authors
178 M. I. TEMKIN

proposed other solutions of the problem of carrying out steady-state hetero-


geneous catalytic processes under nongradient conditions. A number of
arrangements are described by Korneichuk and Streltsov (7). Additionally,
a convenient design of a high-pressure reactor with rotating baskets contain-
ing the catalyst (8)is to be mentioned.
Frequently, particularly from the viewpoint of the technological applica-
tion of a heterogeneous catalytic reaction, the conditions of experiments
in a flow reactor are characterized by space velocity or contact time
values. Space velocity, V, is the ratio to the volume of the catalyst bed of
the volume of a gas mixture, reduced to the normal conditions (OC, 760
Torr), passed through the reactor per hour. If the reaction involves a volume
change, inlet and outlet space velocities should be distinguished. The recipro-
cal of V is of the dimension of time. Contact time (conventional contact
time), t,, is a value proportional to V-. It is defined as the ratio of the
catalyst volume to the volume of the gas mixture passed per unit time, the
gas volume being not under normal conditions but at temperature and pres-
sure in the reactor. Usually, t, is expressed in seconds. It follows from the
definitions given that
t,/sec = (V-/hr)(P,T/3600 PT,) (7)
where P and T are pressure and temperature in the reactor, Po and To are
normal pressure and temperature. In the case of a reaction with volume
change, t, has no exact meaning, if only because of the difference in the inlet
and outlet values of V. Moreover, the inlet (or outlet) V value does not
rigorously determine reaction conditions either, since depending on the
degree of compactness, the unit of bed volume may contain a somewhat
different amount of catalyst. On the other hand, V and t , are engineeringly
convenient because it is often important to know the productivity of a reac-
tor of a given volume, particularly if this is a high-pressure reactor.
When expensive catalysts are used, e.g., platinum, the knowledge of the
productivity of unit mass of the catalyst is more important than that of a
unit volume of catalyst bed.

111. Mass Transfer in Heterogeneous Catalysis

Purely physical phenomena, namely the diffusion of reactants to the


catalyst surface (in the majority of cases, to the inner surface of pores), the
diffusion of products from this surface, and the transfer of heat evolved or
absorbed in the course of the reaction are indispensable components of the
heterogeneous catalystic process. To find the kinetics of a reaction, it is
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 179

desirable to perform reaction rate measurements in the kinetic region; i.e.,


when the physical processes do not affect the reaction rate. Therefore, it is
substantial to be able to judge whether the kinetic region has been reached
at the given experimental conditions. For reactions on porous catalyst grains,
external diffusion (in the bulk of a gas or a liquid to the outer surface of the
grains) and internal diffusion (in the pores of the catalyst to its internal sur-
face) are to be distinguished. In the case of gas reactions, if internal diffusion
does not affect the reaction rate, external diffusion has no influence either,
because effective diffusion coefficients in porous bodies are essentially (e.g.,
10 times) lower than bulk diffusion coefficients; and moreover the diffusion
path inside the grain is longer than the thickness of the diffusion layer at
the outer surface of the grain. This is particularly true for a circulation flow
system since in it the gas mixture passes through the bed of the catalyst grains
at a large velocity, which results in a reduction of the thickness of the difTu-
sion layer.
A purely experimental criterion of the kinetic region, namely independence
of the experimentally observed reaction rate on the size of catalyst grains,
is often used. However, this approach does not guarantee the investigator
against erroneous conclusions, for the intrinsic catalytic activity of grains
of different sizes can differ; e.g., because of the difference in the conditions
of the catalyst reduction. The conclusions based on theoretical calculations
of the possible role of diffusion are frequently more reliable and less time
consuming. The effect of mass and heat transfer on the course of heterogene-
ous catalytic reactions has been paid much attention in the physical-
chemical literature. Referring the reader to the valuable monographs by
Frank-Kamenetskii (9) and Satterfield (lo),we shall give here only some
additional information.
The effect of diffusion on the rate of a heterogeneous catalytic reaction
is characterized by the efficiency factor of a catalyst, q, defined as the
ratio of the actual reaction rate to the rate that would be in the kinetic region
under the same conditions.
Let a single-route steady-state reaction occur in the spherical porous
catalyst grain. It is convenient to transpose all terms of (1) to the right side,
i.e., to represent the reaction equation in the following form:
0 = CbiBi.
Then if Bi is a reactant, the corresponding stoichiometric coefficient, bi,
will be negative; if Bi is a product, bi will be positive. Reaction rate calculated
per unit of the grain volume (not the volume of the bed of grains) will be
denoted as o.Thus,
LO = ropg (9)
180 M. I. TEMKIN

where r is reaction rate per unit surface, t~ is the specific surface area, and
pg is the grain density.
Effective diffusion coefficient of substance Bi in the grain will be denoted
as Di* ; for the simplicity we assume Di* to be constant.
In the case of a kinetic equation of an arbitrary form, the well-known
differential equation describing a combined progress of a reaction and
diffusion cannot be integrated. The first terms of series expansion of the
efficiency factor in powers of the grain radius, a, can be found. A rather
cumbersome calculation gives (ZI)

Here ci is the concentration of the substance Bi;index u shows that the corre-
sponding value refers to the periphery of the grain (i.e., to the distance from
the center of the grain equal to a).
The term of the series (10) containing u2 is, as a rule, negative; therefore,
diffusion usually inhibits the reaction. Indeed, if Bi is a reactant, then bi < 0
and, usually, do/&, > 0; if, on the other hand, B , is a reaction product,
bi < 0 and, usually, do/&, < 0.Thus, all nonvanishing products bi(8w/dci),
are usually negative (although exceptions do exist). The term of (10) con-
taining u4 is, as a rule, positive.
The greater q differs from 1, the slower the series (10) converges; when
the difference is large, the series becomes divergent. Limiting the expansion
as given above, i.e., terminating it by the term with u4, one may obtain q
with a sufficient accuracy at the level not less than ca. 0.85. In order to judge
whether the kinetic region is achieved, a formula adequate for q values still
nearer to 1 is sufficient; therefore, the series can be terminated by the term
with u2. Let us introduce a symbol
Y = - a2 1(bi/Di*)(do/dci),.
i

The minus sign in the definition makes Y usually positive. For q close to
unity
q x 1-Ar. (12)
Therefore, the effect of diffusion may be neglected if the absolute value of
the dimensionless criterion, Y, is sufficiently small. Thus, if /TI6 0.3, then,
according to (12), 0.98 < q 6 1.02.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 181

In a number of cases the kinetics is described by an equation of the form


w =k n cYi
i

where ni is the order of the reaction with respect to substance Bi,k is the
rate constant. Then, according to ( l l ) ,
r = -azw, ci (binjoi*(ci),). (14)

For first-order reactions, w = kc, n = 1, b = - 1 ; therefore, T = a2w,/D*c.


Thiele (12) has obtained an exact expression for this case:

where
cp = a(k/D*). (16)
Therefore, cp = f l .
There was an opinion that a2w,/D*c [or ~ ( w l , l D * c ) / ~is] a generally
applicable criterion of the attainment of the kinetic region. However, this
is not always true. In the case of a reversible first-order reaction
A P B. (17)
w = k + c A - k - c , where k , and k - are rate constants for the forward
and reverse reactions, bA = -1 and bB = 1. It follows from (11) that
T = a2(k+/DA*+ k-/D,*). Since the problem is linear, an exact solution
can be obtained; if we have
cp = a(k+/D,* + k-/DB*)2, (18)
then Eq. (15) is again valid (13). According to (15) and (18), the efficiency
factor, q, is constant regardless of the degree of approach to the equilibrium.
It follows that can be much smaller than unity no matter how small the
reaction rate on the periphery of the grains, 0,. Thus, in the case of a re-
versible reaction, one cannot conclude that the reaction occurs in the kinetic
region only on the basis that its rate is small; for a correct judgment, the
value of the criterion Y should be found as defined by (1 1).
One should take into account the specific features of gas diffusion in
porous solids when measuring effective diffusion coefficients in the pores
of catalysts. The measurements are usually carried out with a flat membrane
of the porous material. The membrane is washed on one side by one gas
and on the other side by another gas, the pressure on both sides being kept
182 M. I. TEMKIN

identical. Often hydrogen and nitrogen are used. The amount of one gas
passing through the membrane per unit time is determined from its content
in the stream of the second gas after the membrane and the velocity of the
stream. Having known the thickness and the area of the membrane, the
effective diffusion coefficient of the first gas is calculated. Starting from this,
further calculations of effective diffusion coefficients of participating gases
for the required values of temperature and pressure are performed, calcula-
tion depending on the interpretation of diffusion; namely whether it is of
the Knudsen, bulk, or intermediate type.
Bulk diffusion coefficients in binary gas mixture are almost independent
of the ratio of components of the mixture. Therefore, it was supposed that
if diffusion in the measurements described above is of the bulk type, i.e.,
the free path of molecules is much lesser than the diameter of pores, then the
first gas diffuses into the second gas at the same rate as the second gas dif-
fuses into the first.
However, as early as in 1833 Graham described in a number of articles
(14) the results of his experiments on diffusion of various gases in porous
bodies, mainly gypsum. He concluded that gases diffuse into each other not
at equal velocities, but at velocities that are inversely proportional to the
square roots of their densities or, in modern terms, to the square roots of
molecular weights. In the concluding paragraph Graham wrote :

The law at which we have arrived (which is merely a description of the appearances, and
involves, I believe, nothing hypothetic), is certainly not provided for in the corpuscular
philosophy of the day, and is altogether so extraordinary, that I may be excused for not
speculating further upon its cause, till its various bearings, and certain collateral sub-
jects, be fully investigated.

Later, after the works of Knudsen on gas flow at low pressures, the result
of Graham was explained supposing that in his experiments the free path of
molecules exceeded the diameter of pores [see, e.g., the book by Herzfeld
(IS)].Only in 1955 Hoogschagen proved experimentally that even with pores
considerably larger than the free path of the molecules, the velocities of the
counterflows of gases are inversely proportional to the square roots of
molecular weights (16). Subsequently, this result was confirmed by a number
of authors (17, 18); none of them, however, like Hoogschagen, mention
Grahams work. In justice, this relationship must be called Grahams
law, as named by Maxwell (19).
The explanation of Grahams law given by Hoogschagen is not com-
plete, as subsequent authors (17,18) stated. However, the attempts of these
authors to give a more complete explanation for the law are not convincing.
It is known that at conventional measurements of diffusion coefficients in
binary gas mixtures using wide capillaries, equal velocities of counterdif-
fusion of the components are observed. From the considerations developed
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 183

in the referenced articles, the reason for the difference between narrow
capillaries where Grahams law is followed and wide capillaries where this
law is not obeyed cannot be seen.
Let steady-state diffusion of components A and B of a gas mixture occur
in opposite directions at a capillary diameter that exceeds many times the
free path of the molecules, the pressure along the capillary being constant.
The concentration of A decreases from left to right and that of B from right
to left. Let us assume that the gas mixture at the beginning does not move.
A small section of the wall of the capillary is impinged upon by the molecules
hitting the wall from the distance of the order of the average free path. Since
the concentration of the molecules of the component A on the left is larger
than on the right, more such molecules will be coming from the left. As a
result, the section will be subjected not only to the normal force of partial
pressure of A, but also to a tangential force directed in parallel to the axis
of the capillary. The tangential force owing to component B has the opposite
direction. However, if the molecular weights of A and B are different, these
tangential forces do not equilibrate. According to the law of action and
counteraction, the gas mixture in this case will be affected by a force equal in
value to the resultant of these tangential forces, but oppositely directed.
Such forces will arise at all elementary sections of the capillary surface.
Under their action the gas mixture will be accelerated until, as the result of
addition of the velocity of the gas mixture as a whole and velocities of mole-
cules relative to the gas mixture, the velocities of the molecules relative to
the capillary will change so that the total momentum transfer to the wall
would vanish. Then the force applied to the gas mixture and, consequently,
the acceleration will also vanish and the gas mixture will move at a constant
velocity. This gas flow, the necessity of which has not been recognized by
the previous authors, may be called concentrational effusion, in analogy to
a long known phenomenon of thermal effusion. The flow occurs in a form
completely different from the usual flow of a gas in a capillary under the
effect of pressure difference, viz., the velocity of the gas mixture in the case of
concentrational effusion is identical at all distances from the axis of the capil-
lary, except for the layer adjacent to the wall, the thickness of which is of
the order of the free path, in which the velocity decreases up to zero. The
requirement that the tangential momentum transfer to the wall per unit
time be equal zero immediately leads, as it has been already demonstrated
by Hoogschagen, to Grahams law.
Such is the explanation of Grahams law from the viewpoint of corpuscu-
lar philosophy; from it follows that this law should be valid for a capillary
of any width on the condition that the pressure along the capillary is constant.
Why then does Grahams law not manifest itself in wide capillaries?
The reason is that the requirement of the constancy of pressure becomes
stricter, the wider the capillary. It is known that, according to Poiseuilles
184 M. I. TEMKIN

law, the volume flowing through the capillary per unit time is proportional
to d4 where d is the diameter of the capillary. A small and therefore imper-
ceptible difference in pressures in wide capillaries creates a Poiseuilles flow
compensating the concentrational effusion. It can be calculated, for instance,
that in the case of counterdiffusion of H2 and N, at 1 atm and 20C, this
compensation at d = 0.01 mm requires the pressure difference, AP = 3.4
Torr and at d = 2 mm it is sufficient to have AP = Torr. Let a vessel
containing hydrogen be connected through a valve and a capillary of 2 mm
diameter to a vessel containing nitrogen and let, as long as the valve is closed,
the pressure in both vessels be exactly 1 atm. When the valve is opened,
concentrational effusion directed towards the vessel with nitrogen starts ;
at the moment the pressure in the vessel with nitrogen becomes higher than
in the vessel with hydrogen by Torr, effusion will be compensated by
Poiseuilles counterflow and the gases will counterdiffuse at equal velocities.
The discussion of diffusion in pores of a catalyst requires taking into ac-
count the specific features of diffusion in narrow capillaries.

IV. Ideal Adsorbed Layers

The regularities of reactions on the catalyst surfaces are of a very com-


plicated nature and their description is only possible on the basis of schematic
and simplified physical models. A model of this kind should, on the one
hand, reflect the main features of the phenomenon and, on the other hand,
result in comprehensible mathematical expressions. The model of an ideal
adsorbed layer or, in terms of the author of the model, Langmuir, simple
adsorption (20)is the simplest and historically the first of the models retain-
ing their importance until now.
It is assumed that the surface of a solid consists of a definite number of
areas of atomic size, sites; each of these sites is capable to bind one particle,
atom or molecule, in an adsorption process. The adsorbed layer is ideal if
(1) all surface sites are identical and (2) the interaction between adsorbed
particles may be neglected.
An ideal adsorbed layer possesses the properties of a perfect (ideal con-
centrated) solution formed by adsorbed particles of one or several species
and free sites. Therefore, mass action law for the rates of surface reactions
and corresponding equilibria is formulated quite similar to the law for
volume reactions in ideal systems with the only difference being that the
equations may also contain, along with surface concentrations of substances,
surface concentrations of free sites.
Mass action law cannot, however, be applied to a reaction rate if two (or
more) adsorbed particles participate in an elementary act when the surface
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 185

mobility of the particles is low and the exchange of particles between the
surface and the volume is not sufficiently rapid. Then, as a result of the reac-
tion, the probability of encountering the particles of reactants on neighboring
sites is lower than the probability at an entirely chaotic distribution of the
particles on the surface. This reservation is not important for the examples
we shall discuss below; mass action law holds true for them.
In contrast to spatial distribution, the equilibrium energy distribution of
adsorbed particles cannot be violated to any substantial degree by reaction
since energy is rapidly transferred between adsorbed particles and solids.
Therefore, the activated complex method may be applied to rates of surface
reactions. For this we consider the activated complex (transition state) of a
surface reaction as a likeness of adsorbed particle (21). But, assuming that
each adsorbed particle occupies only one site, it is necessary, even in the
simplest kinetic model, to consider that activated complexes are able to
occupy not only one, but also several surface sites (21). For example, the
usual picture of a reaction between two particles adsorbed on neighboring
sites involves, in fact, the notion that the activated complex occupies both
sites. When the activated complex occupies several sites, this does not create
any difficulty for the theory since the surface concentration of activated
complexes is an infinitesimal quantity, and so the possibility of overlapping
the required sites is excluded.
If particles enter the surface activated complexes directly from the volume,
e.g., when the reaction occurs as a result of impact of molecules from the gas
phase upon the adsorbed molecules, the expression for the reaction rate
will contain, together with surface concentrations, the values of volume
concentrations. These impact mechanisms were long ago proposed by
Langmuir (22) for the reactions of CO and H, with 0, on the surface of
platinum; the reaction occurs at the impact of a CO or H, molecule against
an adsorbed 0 atom. Such reactions seem to be numerous (23). Along with
this, the above-mentioned adsorption mechanisms that involve the reaction
between two adsorbed particles are possible. Elementary acts of surface
reactions in which more than two particles participate are hardly probable.
It is convenient to consider the processes of adsorption and desorption,
being the constituents of heterogeneous catalytic reactions, as particular
cases of surface reactions. Thus, the reversible process of adsorption and
desorption of substance A can be expressed as follows :

AfZ#ZA (19)
where Z is the free site of the surface and ZA is the site occupied by the mole-
cule A. Applying mass action law, we obtain
186 M. I. TEMKIN

for the rate of adsorption and


r - = k_[ZA] (21)
for the rate of desorption. Here PA is the partial pressure of substance A in
the gas phase, [a and [ZA] are surface concentrations of free and occupied
sites, k, and k- are rate constants. We shall in the following denote by [Z]
and [ZA] the fractions of the free and occupied surface, respectively; this
only results in a change in the numerical values of k, and k- compared t:,
the case when surface concentrations are expressed as number particles on
the unit surface area. With this notation
[Z] + [ZA] = 1. (22)
Equations (20) and (21) are Langmuir expressions for adsorption and desorp-
tion rates. If the equilibrium is reached with respect to the process (19), then,
in accordance with mass action law,
[zAI/PA[zl =a (23)
where a is the equilibrium constant of adsorption (19) which is usually called
adsorption coefficient. From (22) and (23) it follows that
[ZA] = aP,/(1 + UP,), (24)
i.e., the Langmuir adsorption isotherm; [ZA] is usually denoted as 8, or
simply 8 if only one substance is adsorbed.
The value a, being an adsorption equilibrium constant, is related to the
standard Gibbs energy of adsorption, AGOo,by the equation
AGOo= -RTlna. (25)

Here the state with [ZA] = [ZIis taken as a standard state of the adsorbed
layer; thus, in the case when only one gas is adsorbed, the layer is in the
standard state at the coverage 1/2. It can be easily seen that l/a is the equi-
librium pressure at [ZA] = [Z], i.e., at the standard state of the adsorbed
substance. This value may be called desorption pressure; we shall denote
it as b. It is analogous to vapor pressure or dissociation pressure in mono-
variant systems (24). Indeed, in the case of equilibrium of liquid with its
vapor, the surface from which evaporation occurs is equal to the surface
for condensation; the same equality is realized at the adsorption equilibrium
if the fraction of the occupied surface is equal to that of the free surface.
This analogy explains the applicability of the Nernst approximate formula
to desorption pressure (24):
log,, b = -(AHO/4.58T) + 1.75 log,, T + i* (26)
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 187

where AHa is molar adsorption heat (isosteric), z* is a conventional chemical


constant. The latter, as was suggested by Eucken, can be taken as 3 for all
molecules except H, and 1.5 for atoms and H, . With (26) the order of mag-
nitude of b can be easily estimated if AHa is known, or, vice versa, AHa
estimated from b.
An elementary reaction may be, in the general case, represented as follows:
aA + aA + iZI + iZ1 + zZ+products.
Here A and A are the substances entering the reaction from the gas phase,
I and I are the substances entering the reaction from the adsorbed state,
a, a, i, i are stoichiometric coefficients,z is the number of free sites required,
in addition to the sites occupied by particles rand I,for the formation of the
activated complex. In reality, the elementary reaction involves no more than
+ +
three particles including free sites, Z; therefore, a + a i i + z I 3 and
some of the values in the left-hand part of the inequality equal zero. If the
gas phase and the adsorbed layer are ideal, then the rate, r, or the reaction
will be given by the equation

where k, is the rate constant, CA and CAt,are the concentrations of the


substances A and A in the gas phase. We shall express CA and C ., by the
number of molecules in the unit volume. Reaction rate, r, will be defined
as the number of reaction acts per unit area per unit time. According to the
theory of absolute rates of surface reactions (22),
k, = K(k,T/h)L(gf/fAafdt~~~~,)e-(~akB~). (28)
Here K is the transmission coefficient, for the adiabatic reactions (in the
quantum mechanical sense) K z 1, k, is the Boltzmann constant, h is the
Planck constant, L is the number of sites for adsorption on the surface of a
unit area, g is the number of positions of the activated complex when one
of the sites occupied by the complex is fixed (when the activated complex
occupies two sites, g = 2, 4, or 6 depending on the symmetry of the two-
dimensional lattice of the crystal face),fA, etc. are partition functions (sums
over states) of particles A, etc. calculated with the energies of microstates
measured from zero energies of given gaseous or adsorbed particles, i.e.,
including only the energy of thermal motion; for gas particlesf values refer
to unit volume; f # is the partition function of the activated complex not
including the factor corresponding to the reaction path coordinate, E,
is the energy barrier of the reaction (activation energy at T = 0) for one re-
action act.
It is not difficult to obtain from the formula (27) containing concentrations
the corresponding formula with partial pressures of the components of the
188 M. I. TEMKIN

gas mixture, by means of the equations :


CA = PA/kBT, CAP = PAr/k,T. (29)
The constant k, is substituted then by the rate constant k ; according to
(271,
k = k,-/(kBT)"+"'. (30)
The expression for k obtained from (28) and (30) can be put in the follow-
ing form :
k = tc(kBT/h)Lge-AG*'kBT (31)
where AG' is the Gibbs activation energy of the reaction (a standard one,
but it is not denoted since for the process of activation other AG values are
never considered). Referring to one reaction act,
c+

For an estimation of the order of magnitude ofthe preexponential factor


in the Arrhenius equation for a surface reaction between simple molecules,
the termsf' and&, may be taken as equal to 1 since they include only vibra-
+
tional modes that are degenerate if hv k J ( v being the vibration frequency)
(21). The values offA for molecules in the gas phase are calculated in the
usual manner.

V. The Routes of Complex Reactions

We define an elementary reaction as a multitude of reaction acts identical


in the nature of participating particles and in the direction of change. A set
of different elementary reactions occurring jointly and related each to
other by having some of the participating species in common will be called a
complex reaction. Catalytic reactions, homogeneous and heterogeneous (as
well as chain reactions, etc.), are complex. The result of a complex reaction
as given by chemical analysis is described by one or several stoichiometric
equations.
For instance, the reaction of ethylene with oxygen on the surface of silver
results in the formation of ethylene oxide, carbon dioxide, and water vapor;
it is described by the equations
C,H4 + $0,= C,H40
C,H4 + 30, = 2C0, + 2 H 2 0
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 189

Such equations will be called overall chemical equations, reactants and


reaction products will be termed reaction participants.
Along with reaction participants, the chemical equations of elementary
reactions comprising the complex reaction include other species that do not
appear in the overall equations. They are called intermediates.
For the purpose of the kinetic analysis of a complex reaction, its elementary
reactions are grouped into stages (or steps). A stage, first of all, is a pair of
mutually reverse elementary reactions or one elementary reaction if it is
irreversible. Such stages will be called simple stage. Sometimes several
simple stages may be united into one complex stage; this is permissible when
the rates of the constituent elementary reactions are extremely high com-
pared to the rate of a complex reaction as a whole. Chemical equations of the
stages contain, like the chemical equations of elementary reactions, not only
reaction participants but also intermediates.
Overall reaction equations are linear combinations of chemical equations
of stages, i.e., they are obtained by the addition of chemical equations of
stages multiplied by certain numbers (positive, negative, or zero). The num-
bers must be chosen in such a way that the overall equations contain no
intermediates. According to Horiuti, such numbers are called stoichiometric.
To illustrate this notion by a simple example, let us assume that the mecha-
nism of oxidation of SOz on the surface of a solid catalyst, e.g., platinum, is
described by the following scheme :
1. 2 2 + o2P 2 Z O 1
2. zo + sozP Z + so, 2 (33)

0, + 2SOZ = 2s0,
In this scheme, as always in the following, Z denotes a site on the surface of
the catalyst, the arrows denote elementary reactions, and in the overall
equation the equality sign is employed. Stoichiometric numbers by which
the stage equations should be multiplied to obtain the overall equation are
given on the right side. In this instance intermediates are Z and ZO. Another
scheme equivalent to the previous one may be used :
1. 22 + o2 P 2 Z O 4 (34)
2. zo + so* # Z + so, I

$0, + SOz = SO,

Thus fractional stoichiometric numbers are also allowable (although it


is always possible to write overall equations so that all stoichiometric num-
bers are integers).
190 M. I. TEMKIN

The overall equation of scheme (34) is obtained from the overall equation
of scheme (33) when it is multiplied by 1/2. The possibility of multiplying
overall equations by arbitrary numbers is shown by the use of the equality
sign. Such an operation is not permissible for the equations of simple stages;
e.g., the equation
Z++02#ZO
is senseless: it would mean that 1/2 of the 0,molecule participates in an
elementary act. On the contrary, the equation
:o, + so, = so,
means only that from n 0, molecules and 2n SO, mdlecules 2 n SO3 mole-
cules are formed, where n is some number; the equation describes the stoichi-
ometry of the process, but not its mechanism.
In this example the exclusion of the intermediates could be effected,
essentially in one way only, since the column of stoichiometric numbers (i2)
is obtained from the column (i)by multiplication by an arbitrary factor;
this holds for any other suitable column of stoichiometric numbers. However,
this is not always the case-even when the reaction is described by one over-
all equation. Thus, a possible mechanism of oxidation of hydrogen on plat-
inum is (25)
Mi) 2

1. +
02 2Z+2ZO 1 1
2. H, +
2ZFt2ZH 0 2
3. ZO + ZH+ZOH Z + 0 2 (35)
4. ZOH + ZH + H 2 0 + 2 2 0 2
5. ZO + H, + Z H 2 0 + 2 0

N);M2);0,+ 2H2 = 2Hz0


This scheme means that H,O can be formed as a result of an impact of a
H, molecule upon an adsorbed 0 atom, stage 5, or in the reaction between
particles adsorbed on the adjacent sites of the surface, stage 4.
A set of stoichiometric numbers of the stages producing an overall reac-
tion equation is called, after Horiuti, a reaction route. Thus, in scheme
(35) there are two reaction routes, N ) and N(); route N ) cannot be ob-
tained from N(l) through multiplication by a number. Therefore, routes
N ( )and N() are essentially different, although their respective overall equa-
tions are identical.
Let us denote the stoichiometric number of stage s of route p as vp. If
the stoichiometric numbers of all stages of route M3)are obtained from the
stoichiometric numbers of routes N()and N ( 2 ) according
, to the equation
vs( 3 ) = c I v s( l ) + Czv;2 (36)
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 191

where C, and C2 are arbitrary constant numbers, we shall say that route
N 3 )is a linear combination of routes N()and N() and express this relation-
ship by a symbolic equation
3 = C,1 + C,? (37)
Such a definition can, evidently, be extended to any number of routes. It is
clear that if N ( l ) ,N(), N 3 ). . . areroutes of a given reaction, then any linear
combination of these routes will also be a route of the reaction (i.e., will
produce the cancellation of intermediates). Obviously, any number of such
combinations can be formed. Speaking in terms of linear algebra, the reac-
tion routes form a vector space. If, in a set of reaction routes, none can be
represented as a linear combination of others, then the routes of this set are
linearly independent. A set of linearly independent reaction routes such that
any route of the reaction is a linear combination of these routes of the set
will be called the basis of routes. It follows from the theorems of linear
algebra that although the basis of routes can be chosen in different ways,
the number of basis routes for a given reaction mechanism is determined
uniquely, being the dimension of the space of the routes. Any set of routes
is a basis if the routes of the set are linearly independent and if their number
is equal to the dimension of the space of routes.
It may occur that when stoichiometric numbers of stages are chosen to
form a route, the cancellation of one intermediate at the summing up of
the stage equations inevitably leads to the cancellation of another. Thus, in
the mechanism (33) this is the case with the intermediates Z and ZO. This
results from the equality
[Z] + [ZOI = 1. (38)
We shall call such relationships balance equations. If there were two types
of sites on the surface, we would have two balance equations.
It can be shown algebraically that the number of basic routes, P,is deter-
mined by
P=S+ W-J (39)
where S is the number of stages, W is the number of balance equations, and
J is the number of intermediates. In the example (35) S = 5, W = 1, and
J = 4 (intermediates Z, ZO, ZH and ZOH); therefore, P = 2. Since N)
and MZ) are linearly independent and P = 2, N1) and N2)form a basis.
The relationship (39) was given in a somewhat different form by Horiuti
(26).
A linear independence of routes does not imply linear independence of
the respective overall equations. For instance, as mentioned, the basic routes
of the reaction (35) result in the same overall equation. A more complicated
192 M. I. TEMKIN

example is offered by the reaction of oxidation of ethylene on silver; although


its stoichiometry is described by two equations given above, three basic
routes correspond to the reaction mechanism discussed later in Section XII.
A set of linearly independent overall equations such that any other acceptable
overall equation of the reaction in question is a linear combination of equa-
tions of this set forms a basis of the overall equations.
Let the number of basic overall equations be Q;it is evident that Q IP.
Let us denote the number of substances participating in the reaction as M
and the number of independent components, in the sense this notion is
used in the Gibbs phase rule, as C ; then
Q=M-C. (40)
In the reaction (35) M = 3 (0,,H,, and H,O), and 0, and H, can be taken
as independent components; thus C = 2. Therefore, Q = 1.
In the cases when P > Q the basis of routes can be always chosen in such
a way that the P - Q routes will have overall equations 0 = 0. Such routes
do not result in any chemical transformation, they will be called empty.
Thus, instead of the basis of routes in the scheme (35), a basis may be used
consisting of route N ( ) (or W 2 ) )and an empty route [ M 2 )- N ) ] . We
should not be embarrassed by the negative sign of the stoichiometric number
of stage 5 in the route [N) - N] indicating that the stage follows the re-
verse direction. Although stage 5 in the scheme (35) is shown to be irrevers-
ible, it implies only that the rate of the reaction in the reverse direction is so
small that it can be ignored; in general, all elementary reactions are revers-
ible and the reaction along the empty route under discussion is possible.
Moreover, this is of no importance since the operation of the substitution
of one basis of routes for another is of a purely algebraic nature.
The basis of routes with Q nonempty and P - Q empty will be called
stoichiometric basis of routes because, being a basis of routes, it deter-
mines simultaneously a basis of stoichiometric overall equations.

VI. The Steady-State Conditions

In deriving the kinetic equations of heterogeneous catalytic reactions, the


surface concentrations are assumed to be steady state (or stationary), as
has been done by Langmuir in the previously mentioned study of the reac-
tions of CO and H, with 0, on platinum (22). The treatment of surface
reactions as including adsorption equilibria widely used by Hinshelwood
and other authors is a particular case of this more general approach of
Langmuir.
Catalytic reactions in industrial reactors are, as a rule, steady state.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 193

Laboratory studies of the reactions at steady-state conditions have the


advantage of the much simpler mathematical analysis of the results com-
pared to nonsteady processes since the problem of deriving kinetic equa-
tions corresponding to a given reaction mechanism is reduced to the solution
of a set of algebraic equations instead of differential equations in the general
case of a nonsteady reaction.
Since mass action law for elementary reactions in ideal adsorbed layers
(including also adsorption and desorption processes) coincides in its form
with mass action law for elementary reactions in volume ideal systems,
general results of the theory of steady-state reactions are equally applicable
to volume and to surface reactions. They are very useful when the reaction
mechanism is complicated.
The concepts of the theory of steady-state reactions are better explained
with examples, but here they will be formulated mainly in a general form;
the examples will be found below in the discussion of concrete reactions.
A reaction is at steady state if the concentrations of all species in each
element of the reaction space (i.e., volume in the case of a homogeneous reac-
tion or surface in the case of a heterogeneous reactions) do not change in
time. Evidently, this is possible only in an open system such as a tubular
reactor or a circulation flow system. It has been stated in Section V that the
species participating in elementary reactions that comprise a complex reac-
tion are grouped into two categories : intermediates and reaction partici-
pants. In a steady-state reaction, the concentration (volume or surface) of
an intermediate does not change in time because the sum of rates of formation
of this species in elementary reactions is equal to the sum of rates of its con-
sumption in other elementary reactions. The concentration of a reaction
participant does not change in time because, along with the formation and
consumption in elementary reactions, this species enters an element of the
reaction space or leaves it; in the first case this is a reactant, in the second
case this is a product.
For an elementary reaction, the concept run introduced in Section I1
is equivalent to a single act of the reaction. The number of runs of a simple
stage is defined as the difference of the number of runs of the forward and
the reverse elementary reactions comprising the stage. We define the run
along a route as a set of runs of stages in numbers equal to the stoichiometric
numbers of these stages for the given route.
For a steady-state reaction the formation of a molecule of an intermediate
in an act of an elementary reaction must be compensated by consumption
of this molecule in some other elementary reaction. If a new molecule of the
same or another intermediate is formed in this reaction, it must also be
consumed. Sooner or later a moment comes when a complete compensation
of the formation and consumption of the intermediate molecules occurs in
194 M. I. TEMKIN

a certain group of reaction acts. This implies that a run along a certain route
of a given complex reaction is completed, the route by no means being an
obligatory basic one.
Now each route of the reaction can be represented as a linear combination
of its basic routes; therefore, a run along any route can be expressed as a
corresponding linear combination of runs along the basic routes. In accord-
ance with this, let us substitute for the runs along nonbasic routes (i.e., the
runs that are the actual constituents of a reaction) the equivalent runs along
basic routes. This means, for example, that if there was a run along a route
[C,N() + C2iV2)+ C,M3)] where N), M2), and M3) are basic routes, we
shall say that there was C , runs along the route N), C2 runs along the
route M2), and C , runs along the route M3). Let us consider a sufficiently
long period of time to neglect the runs along the routes that have begun
before this period and not terminated before it, or the routes that have begun
during this period and not terminated before its end. It is always possible,
since the reaction is at steady state and, hence, the period of time is not
limited. Then, neglecting uncompleted routes, we shall have all eventuated
reaction acts distributed singularly and without remains between basic
routes, forming runs along these routes. The ratio of the number of runs
along the basic route obtained in this manner to the period of time and the
size of the reaction space (volume or surface) is the rate of the reaction along
this basic route.
The notion rate along a basic route has a rigorous sense only for steady-
state reactions; in the general case the runs along stages cannot be singular
and without remains divided between basic routes. The rate of a steady-state
reaction is determined by the rates along the basic routes, as a vector in the
usual three-dimensional space is determined by its components along the
coordinate axis.
Let us denote the rate of the sth stage as rlsl;if this is a simple stage, then
rlsl= r, - rPs (41)
where r, and r - s are the rates of the forward and the reverse elementary
reactions comprising the stage. The rate along the basic route W)is denoted
as d p ) ; according to its definition there are v$)#~)runs of the sth stage per
unit time per unit reaction space, which was the share of the NP) route after
the distribution of the runs of stages between basic routes. The total number
of runs of the sth stage per unit time per unit reaction space is obtained by
summation over all P basic routes. Thus,
P

This equation was advanced originally by Horiuti and Nakamura (26) in a


INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 195

formal manner, the rates along basic routes being defined as the coefficients
in linear relations connecting vip) and rlslvalues.
In the most general case the rates of all elementary reactions are mutually
comparable, then all stages are simple and not at equilibrium. Equation
(42) is applicable to each stage. Expressing rls,in this equation through the
concentrations of species (volume of surface concentrations) in accordance
with (41) and mass action law, we obtain S equations according to the num-
ber of stages that relate rates along basic routes with concentrations. To
these S equations, W balance equations are added. J concentrations of in-
termediates and P rates along basic routes are the unknowns. According to
(25)
J + P = S + W, (43)
so the number of equations is equal to the number of unknowns.
If, for some stage r, x r - , B rlslthe stage is called fast, it may also be
called a quasi-equilibrium stage. In addition, strictly equilibrium stages are
also possible; namely, if stoichiometric numbers of a stage for all basic
routes are zero, then r, = r - , (ie., rlsl= 0). Herewith r, and r - s may be
small. Reversible adsorption of a catalytic poison may serve as an example
of such a stage. Quasi-equilibrium and strictly equilibrium stages will be
united under the general name equilibrium stages.
If a stage is an equilibrium one, then instead of the steady-state condition
(42) the equilibrium condition according to the mass action law is employed;
this does not change the number of equations and unknowns. Fast non-
equilibrium stages do not give equations, but then the number of unknowns
decreases correspondingly since some concentrations of intermediates are
not considered. Thus, (42) always solves completely the problem of the rate
of a steady-state reaction, at least in principle. The problem is reduced to
the solution of a set of algebraic equations. We shall call (42) stage steady-
state conditions. In fact, we need to apply this equation only to nonequi-
librium, and hence, simple stages; therefore we may substitute for rlsl the
difference r, - r - s without loss of generality and use the equation in the
following form :
P

When a reaction occurs in an ideal system (i.e., in ideal gas mixture, ideal
solution, or ideal adsorbed layer), then r, and r - , in (44)are determined by
simple mass action law. We shall call linear the stages whose rate, rlsl=
r, - r - , , depends linearly on the concentrations of intermediates (including
free sites of the surface); the stages whose rate depends nonlinearly on the
concentrations of intermediates (i.e., includes squares of concentrations of
196 M. I. TEMKIN

intermediates, their products, etc.) will be called nonlinear. A stage is


linear if each of the two elementary reactions comprising a reversible stage
(or the only elementary reaction of an irreversible stage) include only one
particle of an intermediate.
A reaction mechanism, all stages of which are linear, will be called linear.
For such a mechanism, (44)produces a linear set that always has one solu-
tion. This solution can be obtained algebraically in an explicit form. If the
reaction mechanism is nonlinear (i.e., if it includes nonlinear stages along
with linear ones), the existence of several solutions of a system of (44)(i.e.,
of several steady states of the reaction) is possible in some cases (28). Some-
times the steady-state course of a reaction is not reached at all and sustained
oscillations of the rate (29) or a continuous acceleration of the reaction of
an exponential type (30) occur. In the industrial catalytic processes dis-
cussed below, these possibilities are not realized even in the cases when the
mechanisms are nonlinear; therefore, it is not expedient to discuss these
possibilities here in more detail.
The original more formal proof (31) of (44)consisted in the demonstra-
tion of its equivalence to the steadyLstate condition in the Bodensteins
form; i.e., the condition that the rates of formation of intermediates are
equal to zero.
Let xis be the stoichiometric coefficient of an intermediate X j in the
chemical equation of a stage s. It is understood that xjs > 0 if X j is being
formed, xjs< 0 if X j is being consumed in stage s, and xjs = 0 if Xjdoes not
participate in this stage. By definition, stoichiometric numbers for the
route N(P)satisfy the equation

s:
s= 1
vpxjs= 0. (45)

Let r ( X j ) be the rate of formation of an intermediate Xj; i.e., the number of


molecules X j formed per unit time per unit reaction space. Then
F

Substituting rISlfrom (42)into (46),we obtain

V(Xj) = c c
S

s=l
xjs
P

p=l
v:p)r(p)
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 197

From this and (45) it follows that r ( X j ) = 0. This conclusion holds for all
intermediates. Thus, if (42) is fulfilled for all stages, the reaction is at steady
state.
The conditions of steady state in the form of Bodenstein are obtained
having r ( X j ) = 0 in (46). In contradistinction to (42), these conditions may
be called the intermediates steady-state conditions. They define only a part
of the unknowns, viz., the concentration of intermediates, Xj, but there is
usually no difficulty in the determination of the reaction rate from these.
In a number of cases the application of states steady-state conditions for
obtaining kinetic equations is more convenient. Besides that, these condi-
tions are a source of certain general relations to be discussed later.
If the number of basic overall equations, Q, is smaller than that of basic
routes, P,it is expedient to use a stoichiometric basis of routes for a descrip-
tion of the reaction. In this case we need to know only Q rates along non-
empty basic routes; the remaining P - Q rates are not required (but this by
no means implies that they equal zero).

VII. Some General Relations for Steady-State Reactions

The choice of a basis of routes for a given reaction is ambiguous and,


therefore, to a certain degree arbitrary. Let us consider the following prob-
lem: having known the rates r), d2), . r() along the basic routes N1),
e

N2),. . N p ) ,find the rates r), r() . . . r) that answer to another basis
of routes, N),N2). . MP)(31, 32).
Let the passage from the first basis (not primed) to the second (primed)
be defined by equations of the form
i = c1 1 v()
s + c12 v s( 2 ) + . . . C,,v:p+

v p = cp 1 v(sl) + cPZ P)
s + . . . + cppv:p
where C l l , etc. are coefficients. Then the sought for rates will be found
by solving, with respect to r(),P),. . . r(), the set of equations
= c11r(l) + c . . . + Cplr(P)
2 1r ( 2 ) +
r(2) = + . . . + cP2+P)
C1 2r(1) + cZ2+2) (48)
. . .
rp = c1Pr(l)+ cZ P r(2) + ... + Cppr(P)

It is composed in such a way that the coefficients of each row of the set (47)
198 M. I. TEMKIN

are situated in a corresponding column of the set (48). Indeed, according to


the equalities (48)
D

,..

+ (CPlv:l) + CP2v:') + . . + Cppvip))r(p)'


Hence, and from (47)
P P

According to (42), the s u m in the left-hand side of (49) equals rlsl;therefore,


the right-hand side of the equation also equals rlsi.This means that the set
of rates r(')',r(2)',. . r"" satisfies the stages steady-state conditions (i.e.,
+

it is the sought for solution).


It has been already noted that the rate of a steady-state reaction can be
regarded as a vector in the P-dimensional space specified by its components,
which are the rates along the basic routes. In terms of linear algebra, the
above result means that when the basis of routes is transformed the reaction
rate vector along these routes is transformed contravariantly.
The equations determining the kinetics of a steady-state reaction com-
prise, together with unknown rates along the basic routes, unknown concen-
trations of intermediates. Only the rates as functions of the concentrations
of reaction participants are usually required; therefore, the unknown con-
centrations are to be excluded from the equations. In many cases this is
made easier by application of an equation that is obtained as follows (33).
We form an identity including the rates of in stages with numbers, sl,
s2, . . . s,,,, chosen arbitrary from the total number of stages, S:

(rsl - r-sl)rszrs3* . . rs, + r-&s2 - r-s2)f-s3 - - rs,


+ r-s,r-sz(rs3 - r - s 3 ) * . . rsm + * * + r-s,r-szr.-s3. *(rs, - r - s m )
= r s l r s z r s 3* ~ ram - rslr-szr-s3. . . r-s,. (50)
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 199

It is easy to be satisfied that this identity is correct by opening the paren-


-
theses. Expressing the rates of the stages rsl - r - s l - r(s1lY rs2 - r - s 2 = rls*l9
etc. in accordance with the steady-state condition (44)and regrouping the
terms one obtains
...

= r s1y s2 rs 1 rs, - r-s1r-s2r-s3 . . * r-s,. (51)


The order of stages in (51) is arbitrary, their number 1 Im I S. At m = 1
(51) becomes (44). The direction opposite to that accepted in the record of
the reaction mechanism may be assigned to a stage, but then the signs of the
stoichiometric numbers of this stage should be changed (- instead of +
and + instead of -). It can be easily seen that the rule on writing the left-
hand side of (51) is as follows: the items of the factor at the rate along the
basic route rp) are obtained from the product r,,r,, * . . r,, by a successive
substitution of v::), v$), etc., for the first, second, etc., factor, the stage
number subscripts acquiring a minus sign in the factors preceding that
replaced. Equation ( 5 1) will be called the steady-state reaction equation.
We shall denote the rate of a single-route reaction as r and the stoichio-
metric numbers of its stages as v,. The equation of single-route steady-state
reactions follows directly from (51):

In a number of reaction mechanisms groups of stages can be separated,


such that the stages participate not in all, but only in a part of basic routes.
This means that the stages of such groups have nonzero stoichiometric
numbers only in these routes and are related to the remaining basic routes
only via intermediates. Such group of stages will be called a block.
If a reaction mechanism consists of blocks, it is expedient to apply steady-

This term is used in the theory of graphs (34).


200 M. I. TEMKIN

state reaction equation to each block separately including in it only the


stages of this block. Since stoichiometric numbers of stages not comprising
a given block equal zero, the equation will contain only the basic routes of
the block and have such a form as if there were no other blocks. In particular,
if a block includes only one route, then the single-route steady-state reac-
tion equation can be applied.
A general method of application of (51) for the elimination of unknown
concentrations with the help of a graph of a reaction mechanism is described
elsewhere (27).
Here we limit ourselves to the application of (52) to a simple two-stage
mechanism :
1. Z + A , # Z I + B , 1
2. ZI + A , P Z + B2 1 (53)

A, + A, = B, + B2
In scheme (53) I denotes an intermediate particle adsorbed on site Z,A,,
A,, B,, and B, are the molecules of gaseous reaction participants.
The rates of the elementary reactions are
rt = kl[Z]PAl; r - l = k-l[ZI]PB,
(54)
T2 = k,[zI]p~,; Y-2 = k_2[z]P~,.

In addition we have a balance equation:


[Z]+ [ZI]= 1. (55)
Let us take s1 = 1, s2 = 2; then from (52) we obtain (since v1 = v 2 = 1)

Canceling [ZI],
we find
lCZpA2 -k k-lPB,
[Z]= r (57)
klPAlk2PA2 - k-1PB1k-2PB2
Taking s1 = 2, s2 = 1, in accordance with (52)

Now we can cancel [Z]; this gives

[ZI]= r klPA~ + k-2PB2 (59)


klPAlk2PA2 - k-lPB,k-2PB2'
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 20 1

Adding (57) and (59), we obtain

whence, according to ( 5 9 ,

This deduction illustrates the method of application of (52) that also can
be used in more complicated cases. The intermediate steady-state conditions
lead to the same result (35).
If both stages of scheme (53) are irreversible, then, having k - = k - 2 = 0
in (61), we obtain
k l P A l k2PA2
r=
klPAl + k2PA2'
This equation can be written as follows:
llr = ( l / k l p A ~ ) + (l/kZpAz)* (63)
It can also be easily obtained directly from the stages steady-state condi-
tions. Indeed, applying these conditions to each stage, we have

Therefore,
[z] = r/klPAlr [a]= r/k2PA2.
As a result of (55), we obtain (63) by summing these equations.
In general, (51) and (52) are useful for the deduction of kinetic equations
of such reactions, all stages of which (or at least a considerable part of
those stages) are reversible. If all stages are irreversible, it is convenient to
use stage steady-state condition (44).
It is easy to see how the obtained kinetic equations are to be changed
when, in a reaction analogous to that described by the scheme (53), the
number of gaseous participants differs from that in the scheme. If, e.g., two
species A, and A,' participate in the forward direction of the stage 2, there
will be the product P A I P A 2 t , instead of PA2,in the kinetic equation. If species
A2 is absent, 1 must be substituted for PA2,and so on.
The steady-state rate of a many-stage reaction is not attained immediately;
202 M. I. TEMKIN

we shall consider the establishment of a steady state using a simple two-


stage catalytic mechanism as an example (36).
Let the concentrations of A,, A2, B1, and B, be constant and let the
coverage of surface [ZI] at a certain initial moment be not equal to the
coverage that would correspond to a steady-state reaction at the given
concentrations. A simple calculation (36) shows that the reaction rate, r,
will change in time t , according to the law
r = r, + (ro - r,)e-rc (66)
where r , is the steady-state reaction rate corresponding to (61), which is
attained asymptotically (formally, at t = co). ro is the value of r at t = +O
(i.e., at t = E where E is a positive value approaching zero) and z is the re-
laxation time of the reaction rate

Here L is the number of catalytic centers (i.e., sites for adsorption of par-
ticles I) on unit surface.
The ratio r/L is called turnover number of a catalytic site. The reverse
value u = L/r is of the dimension of time. It may be called turnover time;
it is the average time required for a catalytic center to complete the cycle
of transformations and to return to the initial state. The designation u
corresponds to the German word die Umschlagszeit [i.e., the time of turn-
over (of stock)]. On the basis of (61) and (67), it can be demonstrated that

t <$U, (68)
where u, = L/r,; the sign of equality is obtained if, first, the stages of
4
mechanism (53)are irreversible and, second, LZ] = [ZI] = [see Temkin
(36)l.
Thus the u value, which can be easily calculated directly from experi-
mental data, permits one to estimate the upper limit of z. It follows from
the meaning of u that in a mechanism more complex than (53) t also usu-
ally does not exceed u.
A reaction in a closed system cannot be strictly steady state. If, however,
during a period of the order of relaxation time the concentrations of reaction
participants vary negligibly little, the rate of the reaction at each moment
is practically the same as it would be at constant concentrations equal to
the concentrations at this moment. It is said in this case that the reaction
is quasi-steady state.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 203

All equations that are strictly correct for steady-state reactions are also
approximately correct for quasi-steady-state reactions. The method of ap-
proximate treatment of quasi-steady-state reactions as steady state is due
to Bodenstein.

VIII. Reversible Many-Stage Reactions

A reaction is reversible as a whole only if all its stages are reversible.


Let us write (52) for a reversible single-route reaction so that all its stages
are included (their number is designated as S ) :
rlr2v3...rs - r - 1 r - 2 r - 3 . * ' v - s
r= . (69)
v1r2r3---rs+ r-,v2r3**.rs + r-,r-,v,-~.r, + . ~ - + r - , r - 2 r - 3 ~ ' * ~ S
According to (69), r is the difference of two quantities,
r = r+ - r - , (70)
where

and
r- lr-2r-3 '..rPs
r- = vlr2r3"'rs 4-r-lv2r3..ars+ r-,r-,v,.,-r,+..' + r-1r-2r-3**-vs (72)

If but one of quantities r l , r2, - - a , r, equals zero, then r + = 0; and if but


one of quantities r - l , r - 2 , .. *, r - s equals zero, then r - = 0. Therefore, it
is natural to consider r+ as the forward reaction rate and r- as the backward
reaction rate (33). It can be shown that this interpretation of r+ and r - is
in accordance with the determination of the forward and backward single-
route reaction rates by means of labeled atoms (37). It follows from (71)
and (72) that (31)

Equations (70) and (73) taken together unambiguously determine r+


and r - .
204 M. I. TEMKIN

Let us apply these equations to a reaction that follows the mechanism


(53). Equations (73) and (54) give

_+ -- - lPB1 k - 2'B2
(74)
r- klpA1k2pA, '

From (70), (74), and (61)


kl k2PA2
r+ = (75)
klPAl + k2PA2 + k - I P B 1 + k-2PB,
and
k-1PB1k-2PB,
r- =
klPAl + k2PA2 + k - l P B 1 + k - I P B 2 '
If any of simple stages is an equilibrium one, then designating its number
as e, we have re x r - e , and these factors cancel in (73). A complex equi-
librium stage consists of simple equilibrium stages. Thus it is sufficient to
include in (73) not all, but only nonequilibrium stages. Such conclusion is
also true for (71) and (72). Indeed, let some simple stages be equilibrium
ones, their numbers being e l , e,, etc. Then rel z r-el, re, z r-ez, etc. These
values are much larger than the rates of elementary reactions of irreversible
stages. Therefore, in the denominator of the right-hand sides of (71) and (72)
the terms that do not contain re,, r - e l , re2,r-e2, etc. may be neglected; then
the values re and r...e cancel.
If only one stage of a single-route reaction is a nonequilibrium, it is called
a rate-determining stage. We shall denote the number of this stage as d .
If there is a rate-determining stage, (71) and (72), as a result of cancellations,
are reduced to
r + = rd/vd (77)
and
r- = r-d/Vd. ' (78)
The meaning of (77) and (78) is evident.
Let us consider a single-route reaction
aA + a'A' = hB + b'B' (79)
where A, etc. are substances and a, etc. are stoichiometric coefficients. We
define the average stoichiometric number, P, as follows :
- v,AGI + v2AG2 + * * . vSAGs +
AGI -k AG2 -k - * . + AG,
v = '
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 205

Here AG, is the Gibbs energy change in the stage number s. According to
(80), 5 is the weighted average of the stoichiometric numbers v, with their
weights proportional to AGs. For equilibrium stages, AGs x 0; thus only
nonequilibrium stages are to be taken into account in the determination
of i . In particular, if only one stage number d is nonequilibrium and, there-
fore, is rate determining, then 5 = vd.
It follows from the theory of absolute rates of reactions that
AG, = - RT ln(r,/r-,). (81)
From chemical thermodynamics it is known that
vlAGl + v2AGz + . * . + vSAG = AG
= - RT In K + RT ~I[(B)~(B')~'/(A)~(A')"']
(82)
where AG and K are the Gibbs energy change and the equilibrium constant
of the reaction (79), (A), etc., are the activities of substances A, etc.; for the
reactions in ideal gas mixtures they are equal to partial pressures PA, etc.
Equations (73), (80), (81),and (82) imply that (31)

The number V is in the general case variable since the reaction conditions
affect the values AG,. However, 5 cannot exceed the largest of stoichiometric
numbers of stages, v,, and since v, are of the order of 1, ij is a number of the
same order. Therefore, (83) could possibly aid in deciding, without knowing
the reaction mechanism, whether r - may be neglected as compared with r + .
If stoichiometric numbers of all the slow stages are the same, then i
equals these numbers and, consequently, is constant. In these cases it is
always possible to write the overall reaction equation in such a way that
5 = 1. If this is done,

as in a single-stage reaction.
Generally speaking, the concept of forward and reverse reaction rates
is inapplicable to many-route reactions, but it can be applied in some par-
ticular cases. Such is the case of a double-route reaction with one of basic
routes being at equilibrium (38).
Indeed, let one of two basic routes, N") and N('), e.g., route N2),be an
equilibrium one. Let us only include in (51) all nonequilibrium stages of
the reaction. All stages participating in route N") are equilibrium ones;
206 M. I. TEMKIN

hence, for all nonequilibrium stages viZ) = 0. Therefore, the factor at r(2)
in (51) equals zero. Thus the equation obtained coincides with the equation
of a steady-state single-route reaction including all nonequilibrium stages ;
it is just the equation used above for the introduction of r+ and r - values.
Another case occurs if a mechanism of a many-route reaction includes
a block that has only one basic route; then, as was explained in Section
VII, the equation of single-route steady-state reactions and, consequently,
the concept of forward and reverse rates can be applied.
It may be noted that a rate of a many-route reaction can generally be
represented as a sum of rates of some imaginary irreversible reactions;
the number of these items, however, may exceed two (39).
Anticipating somewhat, we shall discuss now the relationship between
the forward and reverse rates of a single-route reaction on a nonuniform
surface. The surface under consideration consists of several kinds of sites
differkg in their rate of reaction; their number can be very large.
Let us denote the contributions of a site number i to the values of r+ and
r - as r$ and r?. Let us suppose first that intermediates do not migrate
from sites of one kind to sites of another; thus the reaction proceeds on
sites of each kind as if there were no others. Then according to (83)
]
(B)~(B)~lIp
(A)(A) .
If the same stages are equilibrium ones on all sites and the stoichiometric
numbers of nonequilibrium stages are the same (namely, are equal to a
certain number v), then (85) is applicable to all sites with the same i j = v
so that the ratio r?/rV is identical for all sites. Then

Thus the ratio r - / r + is determined in the same manner as for a uniform


surface.
Equation (83) may also be considered as applicable for a reaction on a
nonuniform surface with different stoichiometric numbers of nonequi-
librium stages ; in this case, however, the average stoichiometric number
in (83) is determined not by (go), but in a more complicated manner ; the
definition must involve averaging with respect not only to stages, but also
to the kinds of the surface sites.
Let us now give up the assumption that there is no surface migration of
intermediates. In this case we deal, in fact, not with a set of simultaneous
independent reactions, but with one complex reaction ; the processes of
migration of intermediates from sites of one kind to sites of another being
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 207

stages of this complex reaction. We can ascribe zero stoichiometric num-


bers to these stages for all basic routes. To each kind of sites of the surface
there is a corresponding basic route of the complex reaction under consid-
eration; the mechanism of the reaction as a whole consists of blocks each
involving one of these basic routes. Since these are single-route blocks,
(69) is applicable to each of them; the rate along the route correspond-
ing to the ith kind of sites, can be represented as the difference between
r$) and r?, and (85) will hold for the ratio of these values. Therefore, the
consideration given above can be extended to a general case of a reaction
with a nonuniform surface with a possible surface migration of intermediates.
The presentation of the rate of reversible reaction, r, as a difference of
two values, r+ and r - , helps in the derivation of kinetic equations; instead
of deriving the equation for r, one may derive the equation for r+ by means
of (71), which is simpler, and then obtain at once the equation for r - , apply-
ing (73).
In analysis of experimental data on the rate of a reversible reaction that
should be single-route, we use the following equation resulting from (70)
and (83):
Y
I

r+ = 1 - (K-~[(B)~(B)*/(A)(A)I) (87)

This equation is particularly useful if V = v (i.e., if V is constant). Having


determined 5 in advance by means of labeled atoms or any other method
or having assumed a certain value of 5 as a hypothesis, we pass over from
experimental r values to r + values that depend in a simpler way on partial
pressures (or concentrations) of substances. This facilitates the determina-
tion of a kinetic equation. For instance, often the following form of equa-
tion is applicable :
r+ = k+PAAPA+rEBPB?! (88)
where k , is rate constant and nA, etc. are the reaction orders with respect
to A, etc. (integers or fractions, positive or negative). Then nA and other
values can be found by means of plots in log-log coordinates.

IX. Nonuniform Surfaces

The model of an ideal adsorbed layer considered in Section IV leads to


consequences that disagree with experimental data. Thus differential heat
of adsorption, as a rule, is not constant, but decreases with the increase of
surface coverage; the rate of adsorption of a gas is described not by the
208 M. I. TEMKIN

Langmuir equation for a uniform surface, but by the Zeldovich and Rogin-
skii equation or by the Bangham equation. Adsorption equilibrium is de-
scribed not by the hyperbolic Langmuir isotherm, but by the Freundlich
isotherm or the logarithmic isotherm (40).
The kinetics of catalytic reactions in a number of cases also agrees only
qualitatively with the equations obtained on the basis of the model of an
ideal adsorption layer; experimental data often lead to fractional reaction
orders (such as cannot be accounted for by dissociative adsorption).
These facts require a complication of the adsorbed layer model. The
experimental agreement can be improved by discarding at least one of the
initial assumptions, viz., either adsorption sites should be taken different
or a mutual influence of adsorbed particles is to be assumed. In the literature
all surfaces for which the assumptions of the ideal adsorbed layer model do
not hold are often called nonuniform. If surface sites are different, this
is according to Roginskiis terminology, biographical nonuniformity ;
if there is some mutual influence of adsorbed particles, this is the induced
nonuniformity. For the description of kinetics of heterogeneous catalytic
reactions, the model of biographical (i.e., true) nonuniformity was ap-
plied most often. We will discuss this model.
The notion of nonuniformity of catalytic surfaces has been originally
advanced by Langmuir (20)and particularly by Taylor. The physical nature
of nonuniformity is insufficiently clarified. To some extent, it results from
the difference in properties of crystal faces, from dislocations, and other
disturbances of crystal lattice. It is possible that admixtures of some foreign
substances is of greater importance. The particles of admixtures change
adsorption energy on adjacent surface sites. The model of nonuniform
surface probably describes the overall result of the effect of particles of
admixtures on adsorbed particles and of the mutual influence of adsorbed
particles (i.e., to an approximation the model takes into account not only bio-
graphical, but also induced nonuniformity).
There are conceivable CI priori different functions determining the frac-
tion of the total number of surface sites characterized by definite adsorption
energy of a given substance, various combinations of these functions at
simultaneous adsorption of two or more substances and, finally, various
combinations of adsorption energy with kinetic characteristics of surface
sites with respect to adsorption and elementary reactions.
The diversity of possibilities of the general picture of a nonuniform surface
makes it indefinite and hence scarcely suitable for concrete applications.
On the basis of experimental data on heat of adsorption, adsorption equi-
libria, adsorption and desorption kinetics, the kinetics of reactions and
electrode processes, a special model of a nonuniform surface has been
formulated (41) that describes the experiments satisfactorily and without
excessive mathematical complexity.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 209

Since this formulation includes standard values of thermodynamic func-


tions, we shall define the standard state of a particle adsorbed on a nonuni-
form surface and the standard state of a free site. A uniform surface with one
substance adsorbed is at its standard state at coverage 1/2 (Section IV). In
order to apply this definition to each site of a nonuniform surface, we shall
restate it as follows: a surface site is in its standard state if the probability
that it is occupied is 112 (in other words, if it is occupied and is vacant
equally often).
We accept the following assumptions :
1. For an elementary reaction of the kind

A+Z+B+ZI (89)
the Gibbs activation energy, AG, is a linear function of the standard Gibbs
energy of adsorption of particles I that is denoted as GOo:
AG = .(AGOo) + const; 0 ICI I 1 . (90)
The constant a is called transfer coefficient; often a = 1/2. It can be
easily seen that (90) is equivalent to a linear dependence of AG* on the
standard change of the Gibbs energy at elementary reaction (89), AGO :
AG = a(AGo) + const. (91)
Indeed, the change in AGO corresponding to a passage from one site of the
surface to another is equal to the change in AGOo.
Equation (89) includes, as a particular case, the process of adsorption,
namely when B is absent and I coincides with A. In this case (90) and (91)
are identical.
2. The values of GOoon different sites of the surface lie between a certain
minimum value, (AGao)min,and a certain maximum value :

(AGoo)min 5 GOo (AGao)max. (92)


The number of sites, dl, with the standard Gibbs energy of adsorption within
the limits of AGOoand AGOof d(AGOo) provided that AGOosatisfies (92) is
given by the formula
dl = CeAGaoIRed(AGOo) (93)
where C and 0 are constants. (Gas constant, R, is introduced in order to
have 0 of the dimension of temperature.) The constant 0 can be positive,
negative, and equal infinity; in the latter case (93) becomes
dl = Cd(AGao), (94)
i.e., the exponential distribution becomes even.
The assumptions given above are principal; two more are added to them.
210 M. I. TEMKIN

3. If at the adsorption of substances I and J on the same site the value of


does not differ much from that of (AGOo)],these values change equally
when passing to any other surface site. In other words, the value of AGO
for the reaction of adsorption displacement
ZI + J = ZJ + I (95)
is the same on all surface sites [since for reaction (95)AGO = (AG,)] - (AGao),].
4. If there are several nonequilibrium stages of a reaction, the transfer
coefficient, a, may be taken identical for different stages. This assumption
is justified by the fact already that often a = 1/2.
The supplementary assumptions (3 and 4) in a number of cases simplify
considerably the derivation of kinetic equations and for this reason are
included into the model, although they, presumably, not always hold.
A linear relationship between AG and AGO similar to (91) was found by
Brqhsted for reactions of homogeneous acid-base catalysis, by Polanyi for
the series of analogous homogeneous reactions of atoms or free radicak2
and by Frunkin for electrode processes. A theoretical interpretation of this
relationship was given by Polanyi and Evans (42). The linear relationship
between AG and AGO was extended to the phenomena of adsorption and
heterogeneous catalysis by the present author (40). Since a is called transfer
coefficient, we may call the relationship under discussion rule of transfer,
meaning the transfer of changes of equilibrium to the rate of elementary
process.
Distribution functions (93) and (94) are substantiated by the fact that
they lead to the most frequently observed adsorption isotherms for the region
of medium surface coverages, viz., the Freundlich isotherm corresponds
to distribution (93) with 0 > 0 and T < 0 as it has been demonstrated by
Zeldovich (43) and the logarithmic isotherm corresponds to distribution
(94) (40). Besides that, if we accept Assumption 1, then distribution (94)
will give the equation of adsorption kinetics by Zeldovich and Roginskii
and distribution (93) will result in the equation of adsorption kinetics by
Bangham. Finally, if Assumption 1 is correct, then distribution (93), in-
cluding distribution (94) as its particular case, follows from the kinetics of
fractional order reactions (44).
The assumptions of the special model of a nonuniform surface were
formulated above in terms of changes of the standard Gibbs energy, AGO,
and the Gibbs activation energy, AG. It can be assumed that standard
entropy, So, of each kind of adsorbed particles and activated complexes
on different sites of a nonuniform surface does not differ substantially; in
this case there can be given practically equivalent formulation of the model
* Later the result of Polanyi was again obtained by Semenov.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 21 1

involving (instead of AGO and AG) the variation of internal energy change
A U (which in an adsorbed layer virtually coincide with enthalpy change
AH) and activation energy E. Then Assumption 1 (i.e., rule of transfer)
takes the form of a linear relationship between E and AU, etc. To avoid
the necessity of making assumptions about entropy, we prefer the formula-
tion in terms of the Gibbs energy.
Let the gas pressure that corresponds to adsorption equilibrium be de-
noted asp. The value o f p for the standard state of a site is denoted as b and
is called desorption pressure of the site (Section IV). Each site of a non-
uniform surface is characterized by a certain b value or by an adsorption
coefficient, a = I/b (for a given temperature). At adsorption equilibrium,
the probability that a site is occupied
0 = p/(p + b) = ap/(l + up). (96)
It is often more convenient to consider not b or a but ( = In b = -In a ;
it was proposed to call this value the desorbtion exponent. According
to (25)
4; = AG,O/RT. (97)
Let us number the sites of a surface of unit area in the sequence of in-
creasing b or ( (i.e., in the sequence of decreasing adsorbtion power). We
shall denote the number of a site as I ; its ratio to the total number of sites
on a section of unit area,
s = l/L, (98)
will be called relative number of the site. The value of s varies from 0 to 1.
Experimentally observed quantities pertaining to the whole surface, such
as the amount of adsorbed substance, heat of adsorption, reaction rate, are
sums of contributions of surface sites or, since the number of sites is extremely
great, the respective integrals. As 5 increases monotonously with s, each of
them can be taken as variable for integration; both methods of calculation
are used. If 4; is chosen as an independent variable, a differential function of
distribution of surface sites with respect to desorption exponents, (p((), is
introduced such that p(()d( is the number of sites with AG:/RT values
within the limits of ( and ( + d5. Let us denote the largest and the smallest
values of 4; as (,, and C1, respectively, and introduce the notation
y = T/O. (99)
Then by virtue of the second assumption the function ~ ( 4 ; )is expressed as
follows :
212 M. I. TEMKIN

at to < 5 < t1 p(t) = AeYC


at 5 > l 1 445) = 0
where A is a constant.
The value of A is determined by the normalization condition

where L is the whole number of sites on the unit surface. According to


(100) and (102) the condition (103) may be written thusly:

Substituting expression (101) with y # 0, we obtain


A = yL/(eYt' - eyto). (105)
At y = 0, i.e., at an evenly nonuniform surface, p({) = A, and (104) gives
'4 = U(5, - 50). (106)
The same result is obtained from (105) evaluating the indeterminate form
in the usual manner.
Some statements referring to nonuniform surfaces are valid irrespective
of the form of the distribution function of surface sites with respect to their
AG,' values. Using the relative number of site, s, as an independent variable,
we obtain the following expression (40) for the equilibrium surface cover-
age, 8:

where a is a function of s (monotonously decreasing). Let the equilibrium


gas pressure, p, be so small that
a,p <1 (108)
where a. is the value of a at s = 0 (i.e., the highest value of a). Then we
can neglect up as compared to 1 in the denominator of the integral (107)
so that

e =p lo1
ads. (109)

Thus, at sufficiently small pressures, irrespective of the character of non-


INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 213

uniformity, surface coverage is proportional to pressure (i.e., the Henry


law holds). Condition (108) means that the probability of being occupied
is small even for the most strongly adsorbing sites. On other sites this
probability is still smaller, thus 8 4 1.
For a site at adsorption equilibrium, the probability of being free is
1 - a = 1/(1 + up). (1 10)
Hence, the fraction of free surface
1
1 - 8 = Jo (1 - a)ds = Jol [1/(1 + ap)]ds.
If gas pressure, p , is so high that
alp 9 1, (1 12)
where a, is the value of a at s = 1 (i.e., the lowest value of a), then, neglect-
ing unity as compared to up, we obtain
f l

Thus, at sufficiently high pressures, irrespective of the character of non-


uniformity, the fraction of free surface is inversely proportional to pressure.
Condition (112) shows that in this case the probability of being occupied
is high even for the sites with the smallest adsorbing power, so 8 is near
to unity.
The conclusion that at low coverages the Henry law holds and at high
coverages free surface is inversely proportional to p is also valid for a uni-
form surface. Thus, in the regions of small or of large coverages the distinc-
tion vanished between uniform and nonuniform surfaces regarding the
shape of adsorption isotherms. The same result can be obtained for kinetic
laws. Therefore, a nonuniform surface differs qualitatively from a uniform
one only in the region of medium coverages. It is essential that in the case
of high degree of nonuniformity (i.e., when a, greatly differs from ao), the
region of medium coverages includes a considerably broader range of values
o f p than when the surface is uniform.

X. Adsorption Equilibrium and the Kinetics of Reaching it


on Nonuniform Surfaces

When a gas adsorbed on a nonuniform surface is in equilibrium with gas


phase where its pressure is p, the surface coverage, 8,is determined by (1 07).
214 M. I. TEMKIN

In the case of an evenly nonuniform surface (94) gives


1 = CAG, + C (1 14)
where C is an integration constant. Introducing s from (98), we obtain
AGa = (L/C)S- (CjC) (115)
and on the basis of (25)
In a = -(L/CRT)s + (C/CRT). (116)
Introducingfand a, by means of
f = L/CRT
lna, = C/CRT
and using (1 16), we have
a = aoe-Js. (119)
As in the previous section, a, is the largest value of a. The smallest value
will be denoted as a, . According to (1 19)
a, = a,e-f. (120)
Substitution of (1 19) in (107) and integration gives (40)
0 = (l/f)lnCU + a o M 1 + alp)]. (121)
The value of f i s expressed on the basis of (120), through a, and a, :
f = ln(ao/a,>. (122)
This value is called the nonuniformity exponent. At high p values, when
a$ 9 1 and a l p 9 1, unity may be neglected as compared to aop and a$
and then from (121) 8 = l/fln(ao/a,), and from (122) 8 = 1 (i.e., the surface
is saturated). In contrast, at low p , when a$ Q 1 and a# 4 1, we expand
In( 1 + a d ) and In( 1 + a l p ) in series and, restricting to the first terms, obtain
0 = (a, - a l ) / f ) p (i.e., the Henry law). If a surface is strongly nonuniform
(ie., a, B a,), then a range of p values exists where a,p p 1 and simul-
+
taneously a l p 1. This means that for the most strongly adsorbing sites
the probability of being occupied is almost 1 and for the most weakly
adsorbing sites it is almost 0. This will be called the region of medium
coverages. We obtain for this region from (121)
8 = ( 1I f ) Waop). (123)
According to this equation, 0 depends linearly on In p ; therefore, we
call (1 23) logarithmic adsorption isotherm. This isotherm, as it was men-
tioned in Section IX, describes experimental data well in many cases.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 215

Isosteric molar heat of adsorption, q, is determined by the equation


(a InplaT), = q/RT2. (124)
Sincefdepends on T according to (1 17), it follows from (123) that q de-
pends on 8 linearly:
q = 40 - RTf8 (125)
where
q, = RT2[dIn (l/ao)/dT]. ( 126)
Thus qo is the heat of adsorption on a uniform surface with adsorption
coefficient a,. According to (1 25), nonuniformity exponent may be regarded
as the ratio to RT of the slope of the straight line in the plot of q as function
of 8. For strongly nonuniform surfacesf is much greater than 1.
If a homonuclear diatomic molecule (e.g., H, , 0, , N, , etc.) is adsorbed
with dissociation into atoms, the adsorption isotherms (121) and (123) give
8 as the functions of partial pressure of atoms in gas phase, pat,that would
correspond to dissociation equilibrium at the partial pressure of diatomic
molecules, p. These values are linked by
P 2 P = K9
where K is the dissociation constant. Thus, (Kp)'', must be substituted in
the equation of isotherm for p. Instead of (123) we obtain
9 = (1/2f) WOKP)
(i.e.,the linear dependence of 8 on Inp is retained for dissociation adsorption).
To derive adsorption isotherm at the exponential nonuniformity as
described by (100)-(102) at 1 > y > 0, it is convenient to use the desorption
exponent, t = -In a, as integration variable. The probability of a site
+
being occupied is B = ap/(l up) = 1/(1 + l/ap), and at adsorption equi-
librium

For the distribution function considered

Let us introduce the symbol


216 M. I. TEMKIN

Inasmuch as u = (1 - a)/a, the meaning of u is the ratio of the probabili-


ties for a site to be free or occupied. According to (129),

d< = du/u (130)


and since
= Y Y
P U , (131)
(128) takes the form

1 l h p
e = ( I / L ) ~ Y1 loop [uY-'/(1 + u)]du. (132)

This integral generally cannot be expressed through elementary functions.


Because of it, we consider medium coverages only when alp 4 1 and a,p % 1.
In this case 0 and 03 may be substituted for the limits of the integral without
an appreciable error. We assume that 0 < y < 1. It is known that at 0 < h < 1

Jom [uh- '/( 1 + u)]du = n/sin hn. (133)

As etl = l/al, eco= l/a, and a, % a,, we may neglect eyro compared to
eYc1in (105). Then AIL = yaIYand, hence
t? = (yn/sin yrt)(alp)Y. (134)
Therefore
e = cpy (135)
where
C = (yn/sin yn)aIY. ( 136)
Equation (135) is the well-known Freundlich adsorption isotherm. In a
number of instances this isotherm accurately describes experimental data.
The interpretation of the Freundlich adsorption isotherm as resulting from
exponential nonuniformity of surface is due to Zel'dovich (43).
A kind of nonuniformity of surface with negative y is also conceivable.
If -1 -=
y < 0, then a similar derivation gives for the region of medium
coverages the following adsorption isotherm (44) :
e = 1 - (C/rlYI) (137)
where Iy( is the absolute value of y.
To derive equations of the rates of adsorption and desorption we shall
assume that the rate of migration of adsorbed molecules along the surface
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 217

is sufficiently high for the equilibrium within the adsorbed layer to exist
even when the layer is not in equilibrium with the gas phase. Then the
probability for each surface site to be occupied is determined by a certain
value p that is one and the same for the whole surface, p being the pressure
in the ideal gas phase that would be in equilibrium with the surface at the
given coverage. This value will be called fugacity of adsorbed substance.
When there is no equilibrium with gas phase, p differs from the actual gas
pressure, which we shall denote as P ; if p < P, the amount of adsorbed
particles increases with time; if p > P, this amount decreases. We shall
consider the rate of change of number of adsorbed particles on unit surface,
r ; it may be positive or negative, as it is the difference between the rates of
adsorption (i.e., condensation),r + ,and desorption (ie., evaporation), r - :
r = r+ - r - . (138)
On a uniform surface
Y+ = K + p ( 1 - 6) = (K+P)/(l + Up) (139)
where K + is the adsorption rate constant. Thus, the contribution of one
+ up),
site of a nonuniform surface to the rate of adsorption is (l/L)(~+p)/(l
and the rate of adsorption on a nonuniform surface is

Since we assume the linear relationship between AG+* and AG,, (91),
we obtain for a nonuniform surface, on the basis of (25) and (31),
K + = K+Oe-a(C-Co)
(141)
where K + is the value of K at ( = to(i.e., on the most strongly adsorbing
sites). Applying (101) for q(l),we obtain from (140) and (141)

r+ = (142)

Equation (142) contains a new symbol, m :


m=u-y. (143)
We now turn to the integration variable u determined by (129). With the
help of (105) we find that

y+ =-. (144)
218 M. I. TEMKIN

where
n=1-m. (145)
For the region of medium coverages we approximate the integral substitut-
ing integration limits 0 and co for uo and u l . Making use of (133) under the
condition that 0 < n < 1 (and hence, 0 < m < 1) and bearing in mind
that t1 - to= f and sin nn = sin mn, we obtain
n Y K+OP
r + =-*-.
sin mn eyf - 1 (sop)"'
Equation (146) is also valid at y = 0 if, for the factor y/(eyJ- l), its limit
at y = 0 (i.e., l/f) is substituted; additionally, at y = 0, m = a. Therefore,
for evenly nonuniform surface we have

Thus, the special model of a nonuniform surface that we consider gives


for r , the expression
r + = k,P/p" f 148)
where k , and m are constants; m satisfies the inequalities 0 < m < 1. Sub-
script ''&" is used at k because, as will be seen later, the expression for
desorption rate, r - , contains the same constant, k , . It is evident that at
low coverages the adsorption rate is proportional to P and is independent
of p ; this result can be obtained from (148) with m = 0. For large coverages
the probability that a site is not occupied is proportional to l/p; therefore,
the adsorption rate is proportional to P / p ; this is obtained from (148) with
m = 1. Thus, (148) can be regarded as valid for 0 < m < 1.
Equations (146) and (147) express the dependence of the rate of adsorp-
tion on the fugacity of the adsorbed layer, p . In order to obtain r + as a func-
tion of 8, p should be expressed in terms of 0 by means of a corresponding
adsorption isotherm. For an evenly nonuniform surface, (123) and (147) give

Y+ = (n/sin an)(K+/f)Pe-"f' (149)


or

r + = k+Pe-"f' ( 150)
where
k + = (n/sin an)(lc+O/f). (151)
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 219

Equation (150) is the well-known equation for adsorption rate found by


Zel'dovich and Roginskii (45) (sometimes it is erroneously called the "Elo-
vich equation"). This equation was experimentally confirmed for many
cases of chemisorption. It follows from (135) and (146) that on a surface
where the Freundlich isotherm is valid, the adsorption rate is proportional
to P/Om. Inverse proportionality between r + and a fractional power of 0
was found by Bangham (46).
Equations for desorption rate, r - , could be obtained by calculations
similar to those used for adsorption rate, r + , but there is a simpler method
based on the following considerations that are analogous to those developed
in Section VIII. On a uniform surface at gas pressure P and fugacity of the
adsorbed layer p , adsorption rate is
r+ = K + P (-~0) = K + P / ( ~ up)+ (152)
and desorption rate is
r- = K - 0 = K-ap/(l + up). (153)
Hence
r-b+ = (K-/K+)(UP/P). (154)
Since at equilibrium p = P,
K+/K- =u (155)
and, consequently,
r-/r+ = p/P. (156)
This equation holds for the contributions of each site of a nonuniform
surface to adsorption and desorption rates, and hence, for total r+ and r -
values on a nonuniform surface. Therefore, (148) with n taken from (145)
gives directly
r- = kip". (157)
It may seem strange that (148) and (157) contain the same constant k ,
since the comparison of (148) and (146) shows that
k, = (n/sin mn)y/(eYf - l)[~+'/(a~)'"] (158)
(i.e., the expression for k , contains constants K+' and m that describe the
process of adsorption). But since m + n = 1, sin mn = sin nn and, accord-
ing to (155),K+'/K-' = uo,so (158) is equivalent to the following expression:

k, = (n/sin nn)y/(eYf- l)lc-'(u,)". (159)


220 M. I. TEMKIN

Equation (159) contains constants rc-' and n that describe the process of
desorption.
For an evenly nonuniform surface we express p in terms of 0 by means
of (123). Introducing the frequently used symbol
j?=l-Cr, (160)
we have in this case n = B and
r- = (n/sin fin)(.- '/f)ePfe (161)
or
r- = k-esfe
where
k- = (n/sin Bn)(rc-'/f).

Desorption rate depends exponentially on surface coverage [i.e., according


to (162)], as Becker and Langmuir observed experimentally.
It can be easily seen that if adsorption equilibrium follows the Freundlich
isotherm, then r- will be proportional to 8". According to (143), (145),
and (160),
n=j?+y. (164)
Let us consider now the case (47) when there are molecules of two differ-
ent gases on the surface, the ability to be adsorbed not differing greatly
for the two kinds of molecules; we shall make use of Assumption 3 formu-
lated in Section IX (i.e., assume that the change in the standard Gibbs
energy of adsorption at passing from one surface site to another is the
same for both kinds of molecules).
If gas A is being adsorbed from a mixture with gas A , the probability
that at adsorption equilibrium a certain surface site is occupied by a particle
A is
UP
a=
1 + up + a'p'
where p and p' are the equilibrium pressures of gases A and A , a and a' are
their adsorption coefficients for this site ; i.e., the constants of the equilibria
Z+A=ZA (166)
and
Z + A' = ZA'. (167)
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 22 1

We shall denote the equilibrium constant of adsorption displacement


ZA 4- A = Z A i- A (168)
as c. Since (168) is the difference between (167) and (166),
aria = c . (169)
According to the assumption made, c is identical for all surface sites. Thus,
adsorption coefficients of gases A and A are mutually proportional. Equi-
librium surface coverage by gas A is
apds
1 + up + ap
or, since in the case of proportionality of adsorption coefficients c is inde-
pendent of s,
P

The comparison of this equation with (107) shows that the integrand of
(171) can be obtained from the integrand of (107) by substitution of p cp+
for p . Therefore, if (107) leads to some adsorption isotherm of a pure gas A,
0 = F(p), (172)
where F is a function determined by the form of dependence of a on s,
then for the adsorption of A from a mixture with A we have

It can be easily seen that the equilibrium surface coverage by gas A

These results can be generalized naturally for the case of adsorption equi-
librium of more than two gases with proportional adsorption coefficients.
Thus, the passage from an adsorption isotherm of a pure substance to
the corresponding adsorption isotherm of a mixture is very easy, supposing
that the model of a nonuniform surface is applicable and adsorption coeffi-
cients are proportional. If, for instance, adsorption of pure gas A is described
by the Freundlich isotherm (135), then for adsorption of A from mixture
222 M. I. TEMKIN

with A', according to (173), we have

e=c P
(p + cp')' - y'
(175)

Rate of adsorption, r + , of gas A from mixture with gas A will be con-


sidered on the assumption of high surface mobility. This assumption allows
us to consider the probability for a site to be free as determined by the
fugacities, p and p', of the substances A and A in the adsorbed state, the
fugacities being identical for all sites. Then
u+Pds
+ +
1 u p u'p'
where P is the pressure of gas A, K + is the adsorption rate constant of A,
which is a function of s. In accordance with (169)

r+ = jo+ 1
Ic+Pds
+
u(p cp')'
(1771

If adsorption coefficients are proportional (i.e., if c is independent of s),


this equation differs from the equation for the rate of adsorption of gas A
in the absence of A only by substituting ( p + cp') for p . It follows that if
for pure gas A
r+ = m P ) (178)
where G(p)is some function of p ; then for adsorption from a mixture with A'
r + = PG(p + cp'). ( 179)
If, for example, (148) holds for r+ in the case of adsorption of pure gas A,
then for adsorption of A from a mixture with A'
r+ = k,P/(p + cp')". (180)
In a similar manner from

r- =
' K-upds
1 + u p + alp"
if the rate of desorption from a surface containing only particles of A is
r- = H(p) (182)
where H ( p ) is a function of p , then for desorption from the same surface
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 223

containing particles of A and A

It follows from (173) and (174) that


e + el = ~ ( +p cp).
Hence,
p + cp = + el)
~ - 1 ( e (185)
where F - is a function inverse to F. Substituting this into (179), we have
r+ = PG[F-(d + el)]. (186)
Equation (186) means that the rate of adsorption of A from a mixture of
A and A depends on the total coverage of the surface by both gases in the
same manner as it depends on the coverage of the surface by gas A when
only A is being adsorbed. Thus, in the case of adsorption of A and A on
an evenly nonuniform surface we have for the rate of adsorption of A,
instead of (150),
= k+pe-am+u
(187)

Xi. The Kinetics of Reactions on Nonuniform Surfaces

The larger the decrease in the Gibbs energy at a chemical reaction, the
larger the accompanying loss in the ability to perform a useful work, and
hence the less perfect the process is from the viewpoint of thermodynamics.
Therefore, it seems not to be accidental that in commercial large-scale pro-
cesses reactions are utilized that occur at conditions close to equilibrium
[i.e., reactions with small (-AG) values]. Such are the industrial processes
of ammonia synthesis, sulfur dioxide oxidation, hydrogen production by
reactions of methane and carbon monoxide with steam, methanol synthesis,
ethylene hydration, stepwise butane dehydrogenation to butadiene, and a
number of other reactions.
If a system that is in equilibrium with respect to all stages of a complex
reaction is being shifted to states more and more distant from equilibrium
by gradual variation of concentrations of substances participating in the
reaction, then, since the rates of elementary reactions may differ to any
224 M. I. TEMKIN

extent, it is natural to expect that not all the stages will become nonequi-
librium simultaneously. First of all, it will happen to the stage whose ele-
mentary reactions are the slowest. It is not surprising, therefore, that the
rate of single-route reactions at conditions near equilibrium is determined,
as a rule, only by one stage; other stages are at equilibrium. In such cases, as
the numeration of the stages of a cyclic steady-state reaction is arbitrary,
we shall ascribe the number one to the rate-determining stage; equilibrium
stages, if there are several of them, will be united into one common complex
stage, the number of which will be two.
We shall confine ourselves to the consideratiofi of a simple two-stage
mechanism when the surface contains in a significant amount only inter-
mediate particles of one kind, the degrees of coverage by other particles,
if there are any, being small. Such mechanism may be described by the fol-
lowing scheme (48):
I. + Z $ B1 + ZI
A, 1
(188)
2. a2A2+ a2A2+ ZI = b2B2 + b2B2 + Z 1

A, + a2A2+ a2A2= B, + b2Bz + b2Bz

Here A,, A,, etc. are substances, and a,, etc. are stoichiometric coefficients.
The sign = denotes equilibrium (or quasi-equilibri~m).~
The reaction direction, taken as forward, in particular cases may corre-
spond to either the forward or reverse direction of stage 1 of scheme (1 88).
Therefore, we distinguish the directions of stage 1 as follows. The direction
that results in the occupation of a free site on the surface is called the ad-
sorption direction, and the direction that results in a site becoming un-
occupied is called the desorption direction. The rate of stage 1 in adsorp-
tion direction [i.e., in the forward direction in our record (188)] is denoted
as r, and the rate in desorption direction [i.e., in the reverse direction in
scheme (188)] is denoted as rB. When applied to concrete reactions one of
these values will stand for the forward reaction rate, r , , and the other will
stand for the reverse reaction rate, r - . Transfer coefficients for adsorption
and desorption directions will be denoted as a and p, respectively; so
a+P=l.
Adsorption and desorption processes are particular cases of stage 1;
namely, when substance B, is absent, and I coincides with A , , r, is the rate
of adsorption and r, is the rate of desorption. Since the equilibrium of
stage 2 is maintained as a result of mutually reverse elementary reaction,
the particles I very frequently leave some surface sites and appear on others.
Previously the author had proposed a more expressive sign, (the sign of constellation
Libra or balance), for this purpose. It turned out, however, that its use caused difficulties in
typesetting.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS ' 225

Consequently, they are distributed on the surface randomly even if there


is no surface migration. Therefore, in contrast to the discussion of adsorp-
tion and desorption processes in Section X,the mechanism (188) needs no
assumed rapid surface migration (it is evident, of course, that if the migra-
tion occurs, it does not affect the results). The analysis of adsorption and
desorption rates given in Section X needs only minor alterations to its
application to stage 1 ; namely, products KBPB, would be substituted for
rate constants of desorption on separate surface sites, IC- , and the fugacity
of adsorbed particles I, p I for p . Therefore, in analogy to (148) and (1 57), we
obtain

where k , and k , are constants and m and n, as before, are defined by (143)
+ +
and (164) (i.e., m = CI - y and n = /3 y, so that m n = 1).
+
At equilibrium r,, = r,; since m n = 1 , this gives k A / k , = pB&/PA1.
Hence
kA/kB = KA (191)
where k Ais the equilibrium constant of the reaction of formation of particles
,
I from A in the gas phase :
A, = B, + I. (192)
The equilibrium of stage 2 can be considered as the difference of gas phase
chemical equilibrium
I + a2A2 + a2'A2' = b2B2 + b,'B,' ( 193)
and adsorption equilibrium
I + z = ZI. ( 1 94)
The term "adsorption chemical equilibria" has been suggested for equi-
libria such as those shown by stage 2; they can be observed directly by
experimentation (49-52).
Denoting the equilibrium constant of (193) as K,,we have

whence
226 M. I. TEMKIN

and
rB = (kBIK?) P, ,(Pg2Pg2',/p"A',PA';,
)". ( 197)
Equations (196) and (197) are valid under the condition that 0 I m 4 1 or
under the equivalent condition that 0 I n 5 1.
The results obtained indicate the interpretation of fractional reaction
orders frequently obtained in experimental studies of heterogeneous catalytic
reactions.
If in the course of a reaction, not one, but several adsorption chemical
equilibria exist, it is always possible to write the chemical equations of the
equilibria in such a way that the stoichiometric number of one of these equa-
tions be l and that of other equations be zero. Confining ourselves to two
equilibria, we have the following mechanism instead of (1 88) (48):

1. + Z 3 B, + ZI
A,
'0
1

2. a2A2 + a2'A2' + ZI = b2B2 + b2'B2' + Z 1


3. a3A3 + a,'A,' + Z = b3B3 + b,'B,' + ZJ 0

A, + a2Az + a2'A2' = B, + b2B2 + b2'Bz'

We suppose that only substances I and J occupy essential parts of the sur-
face; then the form of equations for r, and rB results from the equation for
the simultaneous adsorption of two gases (1 80).
Let us consider now the kinetics of a reaction following the simple two-
stage mechanism (53) on a nonuniform surface with none of the stages of
the reaction being assumed to be at equilibrium (53).Since, in contradistinc-
tion to scheme (188), both stages of scheme (53) are equivalent, we may
always ascribe number one to the stage whose positive direction is of the
adsorption type. Therefore, without loss of the generality of the analysis,
we can distinguish the directions of the elementary reactions in scheme (53)
as forward and reverse. Accordingly, the rate constants of the elementary
reactions of stage 1 will be denoted as K~ and K - ~and those of stage 2 as
K~ and K - ~ For
. <
brevity we introduce A = - t o .Its largest value is - c1
5 0 ="f
We accept the simplifying Assumption 4 of Section IX (i.e., assume that
the transfer coefficient c1 is identical for both stages). Then, from the linear
'
relationship between AG+ and AG,' for each stage and bearing in mind
that rate constants K , and K - ~refer to adsorption directions, and K - ~
and K~ refer to desorption directions, we obtain, in a manner similar to
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 227

that used in deriving (141), the following equations :


K~ = Kloe-u'
K- = KO
(199)
K~ = KZoePA

K - 2 = K!Ze-".

Here lcl0 is the value of x1 at 1 = 0, etc.


We assume a rapid surface migration; this allows us to introduce the
fugacity of adsorbed particles I, p,, common for all surface sites, although
at a reaction following scheme (53), in contrast to mechanism (188), the
coverage of the surface by particles I is in the general case not an equilibrium
one, but a steady-state one.
According to (75), the contribution of a surface site, p + , to r+ is deter-
mined by the equation

Since

J 50
where the distribution function, cp(<), is given by (101) and (105),we obtain,
passing over to the variable 1,

Here we use once again the symbols m = CI - y and n = B + y.


Let us turn to the variable u determined by equation

Equations (57) and (59) show that u, as in Section X, is equal to the ratio
of the probabilities for site to be free and to be occupied. It follows from
(199) and (203) that
228 M. I. TEMKIN

Dividing the numerator and the denominator of the integrand in (202) by


( K ~+ ~KOZPB2)e-mL
P ~ ~ and using (204), we obtain

In the region of medium coverages uo x 0, u1 x co ; thus, on the basis of


(133) and by substituting sin mn for sin nn, we obtain for this region

We use (83) to find r - . Since stoichiometric numbers of both stages of


scheme (53) equal 1, i~= 1 and
r - / r + = K - 1(PBIPB2/PAIPA2). (207)

According to (199), K ~ / K = - ~(iclo/K!l)e-L and K ~ / K =


- ~t c 2 0 / K ? , ) e L ; there-
~ / K Thus,
fore, K ~ K J K - 1 ~ - 2= K ~ ~ I ~ ~ l~!z. ' ? the product of equilibrium con-
stants of the stages, K I K z , is identical on all surface sites; it is easy to see
that this product equals K,equilibrium constant of the reaction
K10Kz0/K?1K!2 =K. (208)
From this equation and (207)
(209)
r - / r + = K!? 1P~IK!~P~2/K~oP~IK~oP~2.
Multiplying (206) and (209), we obtain

Equations (206) and (210) taken together give

In the case of an evenly nonuniform surface, y = 0 and m = a ; then (211)


transforms into (35)

Equations (206) and (210) are transformed correspondingly.


INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 229

Taking into consideration (208), we write (21 1) in the following form :

where k and A are constants. A = I C ~ ~or,/ according I C ! ~ to (199), A = K ~ / K - ,


for any site. Thus, A is the ratio of rate constants of adsorption directions
of the first and the second stages; this ratio is identical for all sites since it
was accepted that the transfer coefficient, a, is the same for both stages.
Equation (213) holds with m = 0 for small coverages and with m = 1 for
large coverages; at these coverages it coincides in form with (61) since,
according to (59) and (6l), small coverages (i.e., [ZI] < 1) are realized at
+
k l P A , + k - , P B , G k,PA2 k-lPB1and large coverages (i.e., [ZI] x 1) at
k l P A l + k - , P B , 9 k l P A 2+ k - l P B , . Thus, we may take (213) to be valid
atOImI1.
Equation (213) is simplified when the reaction is irreversible. This requires
at least one stage to be irreversible. Let it be the first one (we cannot take
the second stage to be irreversible since it would imply A = CO). As K-' = 0,
(213) becomes

It is to be noted that (213), which describes the kinetics of the reaction in


the region where it is reversible, does not coincide with the equation obtained
by multiplying the right-hand side Of (214) by [1 - K - 1 ( ~ ~ l ~ ~ 2 / ~ ~
This occurs because the expression for r+ in the region of reversibility of the
reaction does not coincide with that for r in the region where the reaction is
irreversible.
If the second stage is much faster than the first, then resulting from the
above-stated meaning of the constant A we may have A = 0. Equation (214)
takes the form
= kPAI(PA2/PB2)m. (215)
The same result is obtained from (189) and (195) at irreversible stage 1
since in this case rl coincides with Y .
,
If both stages are irreversible, then with K! = K! = 0, we obtain from
(206) or (21 1)
r = [y/(eyf- l)](n/sin mn)(Kl o P A , ) ' - m ( ~ Z o ~ A 2 ) m . (2 16)
It is easy to see that the form of the equation will hold also at m = 0 (small
coverages) or m = 1 (large coverages).
230 M. I. TEMKIN

Thus the kinetic equation of a simple two-stage reaction on a nonuniform


surface, if both its stages are irreversible, is of the form

where 0 s n I 1.
The comparison of (63) and (217) that answer the same simple mechanism
reveals graphically the difference in the features of kinetics on uniform and
nonuniform surfaces.

XII. Oxidation of Ethylene into Ethylene Oxide

Oxidation of ethylene with the objective of obtaining ethylene oxide is


conducted on the surface of silver. To the best of our knowledge, on no other
heterogeneous catalyst can ethylene oxide be obtained in more than trace
amounts. Partial oxidation of ethylene into ethylene oxide

is always accompanied by its complete oxidation into carbon dioxide and


water:
C2H4 + 302 = 2C0, + 2H20. (219)
No other reaction products are observed. Like other reactions of organic
substances with oxygen, the oxidation of ethylene is irreversible.
Silver is used in the form of porous pellets or, more often, supported on
an inert carrier with wide pores (e.g., corundum). Commercially, the reac-
tion is performed with ethylene-oxygen and ethylene-air mixtures at about
25OoC, 10-25 atm and time of contact of the order of 1 sec. Higher pressures
facilitate the subsequent separation of ethylene oxide from the exit gas
mixture. Small amounts of some elements are added to silver; this increases
the selectivity [i.e., the fraction of ethylene converted according to (218) in
the total amount of ethylene converted].
Selectivity of a catalyst process is improved by also doping of the catalyst
in other cases; usually it has been ascribed to "partial poisoning" (i.e., to
the poisoning of the catalyst by the dope with respect to undesirable reactions
without poisoning it, or poisoning to a smaller extent, with respect to the
reaction of formation of the desirable product, implying that different reac-
tions are catalyzed by different surface sites). Our studies showed that such
an explanation cannot be applied to the oxidation of ethylene.
Electronegative elements, namely C1, S , Se, and P in the anions SO:-,
Se0:- (or SeOi-), and PO:-, promote oxidation of ethylene. Irrespective
of the method of their introduction into silver, these elements concentrate
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 23 1

on its surface at the reaction temperature; this is proved by the possibility


of rinsing these additives with water or 2% KOH (55). In the experiments
referred to, elements labeled with radioactive isotopes were used ; so very
small amounts of them could be determined. The effect of additives is
revealed at surface coverages considerably smaller than monolayer. As an
example, Fig. 2 gives the results of experiments at 1 atm and 218C with
selenium introduced in the form of silver selenate. The abscissa shows
selenium content and the percentage of silver surface covered with selenium
(in logarithmic scale); the ordinate shows the ratio of the rate constant of
reaction (218) (calculated according to a simplified kinetic equation) to the
rate constant for the catalyst without additives. Figure 2 also shows the
variation in selectivity, SCZH4.It can be seen that at small surface coverages
with selenium the reaction rate is several times higher than that on pure
silver; there is practically no change in selectivity. Hence, the reactions
(218) and (219) are accelerated almost equally. At higher surface coverages
with selenium the reaction rate diminishes and the selectivity increases
owing to a greater retardation of the conversion of C2H, according to (219)
compared to that according to (218). At surface coverages with still much
less selenium than monolayer, the reaction stops completely. Roginskii

3
h

z
i
Lo
J

- 0.2

I I
!!I+,
10-3, IO't, lo-', at. %
0.001 0.01 0.i LO e,,
FIG.2. Modifying of the silver catalyst with selenium. (SC2H,)0x 0.5.
232 M. I. TEMKIN

proposed to use the term modifying for the phenomenon of catalyst


activation with small amounts of an additive and its poisoning by larger
quantities of the same additive. The effect of electronegative additives to
silver at ethylene oxidation is a typical example of catalyst modifying. The
nature of this modifying is elucidated by the results of studies of isotopic
exchange of oxygen on silver containing various amounts of selenium
(56-58). It follows from these results that the increase in surface coverage
with selenium results in a lowering of the bond strength of reactive (weakly
bound) oxygen with the surface. This facilitates the participation of oxygen
in the reaction, but at an excessive diminution of the bond strength the
coverage of the surface with reactive oxygen becomes small and results in a
deceleration of the reaction.
The catalyst modified with selenium is most suitable for the studies of
reaction kinetics since this element, in contrast to chlorine usually used as
promoter in the commercial processes, does not volatilize from the surface
of silver under the reaction conditions. We studied the kinetics of ethylene
oxidation under gradientless conditions (Section 11) using a circulation flow
system in the experiments at atmospheric pressure (59-61) and a reactor
with rotating baskets for the catalyst (8) at elevated pressures (62).
To evaluate the effect of the reaction products on its rate, we conducted
a series of experiments using traps in the cycle of a circulation flow unit.
These traps were cooled with solid carbon dioxide or filled with chemical
absorbents. The experiments showed that all three reaction products inhibit
the reaction, as the removal of these products from the cycle increased the
reaction rate. The inhibiting effect decreases in the series C2H40> H,O >
CO, . When the reaction is being carried out at a high P02/PC2H40 ratio, the
removal of C2H40from the cycle does not change the selectivity. This means
that CO, and H,O are formed as a result of oxidation of C2H4 and not that
of C2H40. At small O2 contents in gas mixture, the selectivity noticeably
decreases. The variation in ethylene concentration in a very broad range
does not affect selectivity substantially; approximately the same selectivity,
viz., SCZH4 x 0.65 can be obtained in gas mixtures with large 0, excess vs.
C2H4 or with large C2H4 excess vs. 0,. These features of the reaction must
be taken into account when formulating the scheme of its mechanism.
Apart from kinetic measurements, the studies of oxygen adsorption on
silver (63, 64, electron work function variations accompanying oxygen
adsorption (65-67), heat effects of adsorption (68), reactivity of oxygen
adsorbed on silver (67), and oxygen isotopic exchange on silver (56-58)
were used for the elucidation of the mechanism of the reaction. The papers
cited contain references to the works of other authors that were used for
the formulation of the reaction mechanism on a level with our results. Here
we shall mention, first of all, the work by Twigg (69),who has ascertained
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 233

the parallel course of partial and full oxidation of CzH4 on Ag and a very
fast oxidation of acetaldehyde CH,CHO to C 0 2 and H 2 0 at the condi-
tions of C2H, oxidation reaction. This allows us to consider CH,CHO
as an intermediate product of complete oxidation (however, Twigg himself
supposed that formaldehyde, HCHO, was the intermediate product).
Important features of the reaction mechanism were clarified by Orzechow-
ski and MacCormack (70). These authors simplified the problem by con-
sidering the initial reaction rates for (218) and (219), obtained by extrapola-
tion (i.e., the rates in the absence of reaction products). They arrived at a
conclusion that the reaction is basically composed of irreversible adsorp-
tion of oxygen and the removal of adsorbed oxygen by ethylene with the
formation either ethylene oxide or an intermediate substance, most likely
acetaldehyde, that is rapidly oxidized to C 0 2 and H 2 0 .
The reaction scheme given below is more complex since it takes into ac-
count the retardation of the reaction by reaction products and changes in
selectivity. It is a development of a number of earlier schemes (59-62,71,72),
but still it cannot be regarded as final.
N"' "2' 1v(3'

1. +
ZO2 C2H4 --* ZO C2H40 + 1 0 1
2. ZO, +
C2H4 --C ZO CH3CHO + 0 1 0
+
2'. CH3CHO 5ZO2 2C02 2HZO + 5 2 + 0 1 0
+
-+

3. Z C2H4O ZC2H4O 0 0 1
4.
4'.
5. Z02 +
+
z o2--C ZO2
2 2 0 z zo2
-+ +
C2H40 = Z 0 2 . C 2 H 4 0
i
0
3
3

0
0
1
0
(220)
6. Z02 +
H2O 3 ZO2' H2O 0 0 0
7. zo2 + co2= zo2'co2 0 0 0
8. Z02 +
C2H4 = Z02'C2H4 0 0 0
9. ZC2H,O-+ZO +
C2H4 0 0 1

N1); C2H4 + to2 = C2H4O


N"'; C2H4 + 3 0 , = 2C02 + 2 H 2 0
"3'; 0 = 0.
The sign = in (220) shows that the stage is at equilibrium (see Section XI).
Numbers with primes (2' and 4') denote fast (instant) irreversible stages;
there is no need in assigning independent numbers to these stages since in
the kinetic equations there will be no characterizing constants. Stage 2' is,
of course, complex; i.e., it is a sum of several simple stages. Their nature is
of no importance, however, because the mechanism of fast stages is of no
significance to the kinetics of the reaction.
Letter Z usually denotes a site on the catalyst surface that can adsorb one
molecule; in the case under discussion it is assumed that Z is a molecule of
the surface compound Ag',")O (superscript s denotes a surface atom). In
contrast to the bulk oxide, Ag20, the surface oxide, Agf'O, is very strong
234 M. I. TEMKIN

as shown by the measurements of the heat of oxygen adsorption on silver.


Because of this, the oxygen in Ag',")O does not participate in the reaction.
Adsorption of O2 on Z gives Z 0 2 (i.e., Ag',)OO,),stage 4.Oxygen in Z 0 2 is
bound weakly. At the impact of the C2H4 molecule from the gas phase upon
Z 0 2 , an act of stage 1 or 2 occurs and a C2H40 or CH,CHO molecule is
liberated into the gas phase. Acetaldehyde is rapidly oxidized further to
C 0 2 and H 2 0 (stage 2'). According to the scheme, the surface oxide ZO
(i.e., Ag',")02)is unstable and disproportionates very rapidly with the forma-
tion of Z and Z 0 2 (stage 4'). Because of this, ZO does not take part in the
reaction.
The interaction of the same particles, Ag'i'O, and C2H4, can have two
different results : the formation of ethylene oxide, C2H40, or acetaldehyde,
CH,CHO (or possibly hypothetical vinyl alcohol that instantly isomerizes
to acetaldehyde). It may be visualized that the results depends on the orien-
tation of a C2H4 molecule at the moment of impact against Ag',")O,. The
two transition states differ in the configuration of atoms. The heights of the
corresponding energy barriers are approximately equal and hence the rate
of stages 1 and 2 are of the same order.
The intermediate of complete oxidation (in our case, supposedly, acetalde-
hyde) would have been a product of partial oxidation if it had not under-
gone further oxidation. Thus, the difference between partial and complete
oxidation is due not to the substantial difference in the nature of the primary
act of oxygen addition, but to the difference in the capability of primary
products for further oxidation under the conditions of reaction. It is evident
that no hydrocarbons can be oxidized to C 0 2 and H 2 0 directly without
intermediate steps.
It has been repeatedly suggested that the parallel course of partial and
complete oxidation of ethylene and of other organic substances is due to
the presence of two forms of chemisorbed oxygen on the surface of the
catalyst; i.e., two different surface oxides. One of the oxides can participate
in partial oxidation, and the other can be involved in complete oxidation.
However, this is not necessarily the case, as it follows from the above dis-
cussion. If two surface oxides participated in the reaction, then in the gen-
eral case it would be expected that each oxide takes part, to a degree, in
both partial and complete oxidation.
It should be emphasized that according to scheme (220) we deal not with
two reactions (partial and complete oxidation), but with one complex reac-
tion. The mechanism under discussion is linked : intermediates Z 0 2 and ZO
participate both in stage 1 (giving C2H40)and in stage 2 (leading to C 0 2
and H 2 0 ) ;Z 0 2 is formed in stages 4 and 4' common to both directions of
the reaction, etc. Therefore, partial and complete oxidation should not be
considered separately; from the viewpoint of kinetics, the reaction must
be considered as a whole.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 235

While the reaction is described by two overall equations, the number of


basic routes is three. The basis of the routes in scheme (220) is chosen so
that the overall equation of route M3)is 0 = 0; i.e., M3)is an empty route.
Thus, the basis of routes is stoichiometric. Therefore, kinetic equations are
required only for two rates; namely that along routes N") and A@).
Calorimetric data indicate that in the case of oxygen adsorption on
oxygenated silver (the surface sites of which we denote as Z) [i.e., in the case
of the process corresponding to stage 4 of scheme (220)], the surface behaves
in such a way as if it were uniform, in accordance with Assumption 3 for-
mulated in Section IX. Thus the model of an ideal surface layer may be used
to obtain the kinetic equations. Applying the stage steady-state conditions
(44) to the slow stages 1, 2 , 4 , and 9, we have
r'" + r'3' = kl[Z02]PCzH4 (221)
r(') = k,[ZOZ]PC,H, (222)
+ 3 P ) = k9[Z]Po, (223)
r(3)= k,[ZCzH40]. (224)
Mass action law applied to equilibrium stages 3 and 5-8 gives
[ZCZH~O]
= K~[Z]~C,H~O (225)
iZo2 ' C2H4OI = K5[Z021PC2H40 (226)
[ZO, . HzOl = K6tZ021PH2O (227)
[Z02. COZ1 = K7[Z021PC02 (228)
tzoz - CZHJ = ~,tZO,I~C,",. (229)
Here K , is the equilibrium constant of stage 3, etc.
Additionally, we have the balance equation
[ZOJ + [Z] + [ZCzH,O] + [ZOZ CZH,O] + [ZOZ.HZO]
+ [ZO, CO,] + [ZOZ. C2H4] = 1.
9
(230)
This equation does not include [ZO] since, due to the instant stage 4',
[ZO] < 1. (23 1)
236 M. I. TEMKIN

and (221) gives


'" = kl[ZoZ1PC2H, - k9K3[ZlPC2H40. (234)
Substituting r(') from (234) and 6,)from (222) into (223), we obtain

From (232) and (235) [Z] and [ZO,] can be found. Let us introduce the
following symbols :
L = (k, + 6k2)/2k4 (236)
M = 3k9K3/k4 (237)
and

From this equation and (232) and (238) we have

From (242) and (243) we find the ratio r(l)/r(') that fixes selectivity:
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 231

To compare this equation with experiments, we introduce selectivity with


respect to oxygen, So,, defined as the fraction of consumed oxygen that
goes to ethylene oxide :
So, = r(')/(r(') + 6r(2)). (245)

The usually considered selectivity with respect to ethylene

'C2H4 = r(')/(r(') + r(2))= 6S0,/(1 + 5S0,).


It is convenient to consider So, since, as follows from (244) and (245),

i.e., So, is a linear function of the ratio PC2H40/Po2. This relationship is


confirmed by experimental results (61,62).
If stage 9 is excluded from mechanism (220) [i.e., if k9 = 0, and hence,
according to (237), M = 01, then (244) will give r(1)/r(2)= k , / k 2 , selectiv-
ity will not depend on the composition of the gas mixture. Stage 9 is in-
troduced into the mechanism to account for the experimentally observed
decrease in selectivity occurring with an increase of the Pc,H,o/Po,ratio.
L. P. Levchenko, N. V. Kul'kova, and the author passed C,H,O at 240C
over a silver catalyst promoted with chlorine or containing no promoters
and observed in the exit gas C2H4 and CH,CHO in small, approximately
equal, amounts (ca 1%). The formation of C2H4 corresponds to stage 9
and the formation of CH,CHO is the result of isomerization of C2H40.
Our experiments differ from those of Twigg (69)and Kenson and Lapkin
(76) who also had observed the isomerization of C2H40 into CH,CHO
on an ethylene oxidation catalyst, in that we employed silver without
carrier. This excluded the possibility that the isomerization took place not
on the silver surface, but only on the surface of the carrier. At the same time
our other experiments have shown that carriers such as A120, and SiO,
can additionally catalyze C2H40 isomerization. If ethylene oxide isomeriza-
tion is assumed to be the main cause of the dependence of selectivity on
PC2H40/P02 ratio, the following stage must be substituted for the stage 9
of scheme (220).

Z.C,H,O -P Z + CHSCHO. (248)


This substitution has only a small effect on the form of kinetic equations;
in particular, the linear character of the dependence of So, on Pc2H40/Po2
is preserved.
238 M. I. TEMKIN

If we have PC2H40 = PHz0= Pco, = 0 in (238), (242), and (243), we obtain


for the initial reaction rates

= 1 PC2H4
(249)
+ K8PC2H4 + L(pc2H4/po2)

Equations (249) and (250) coincide in their form with equations proposed
by Orzechowski and MacCormack for the initial rates.
The comparison of more complete kinetic equations (242) and (243) with
experimentation is hampered by the instability of activity of silver catalysts
(59). The effects arising from the penetration of oxygen into the subsurface
silver layer (63) and the formation of a polymer film on the surface (70),
an extremely high sensitivity of the catalyst to the traces of compounds of
such elements as S and C1 that may be present in the reactants as impurities,
can be the sources of this instability.
To reduce the number of parameters in the kinetic equations that are to
be determined from experimental data, we used the following considerations.
The values k, , k 2 , and k, that enter into the definition of the constant L,
(236), are of analogous nature; they indicate the fraction of the number of
impacts of gas molecules upon a surface site resulting in the reaction. So the
corresponding preexponential factors should be approximately the same (if
these elementary reactions are adiabatic). Then, since k, , k, , and k, are of
the same order of magnitude, their activation energies should be almost
identical. It follows that L can be considered temperature independent.
Similar considerations are applicable to M . Indeed, according to (237),
-,
M = 3k,k,/k4k . Elementary reactions 9 and - 3 (i.e., the reverse direc-
tion of 3), are of the same type; since k, and k-, are similar in order of
magnitude, we may take the ratio k9/kW3as temperature independent.
Analogously, we arrive at the conclusion of approximate temperature inde-
pendence of k,/k4. Thus M , as well as L, may be considered as temperature
independent.
K , ,Ks, K 6 , Kl ,and K , appearing in the kinetic equations are adsorption
equilibrium constants and their dependence on temperature is determined
by respective heat of adsorption. Since these constants are of the same order
of magnitude, the heat of adsorption should be approximately the same.
Thus the temperature dependence of K , , K , , K s , K l , and K , can be de-
scribed with some common value of the heat of adsorption.
It is easily seen that constants k, and k2 with the rates referred per unit
mass of the catalyst should be proportional to the specific silver surface
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 239

area; conversely, L, M , K 3 , K5,K 6 , K , , and K, should be independent of


the surface area (for a given coverage of the surface by a promoter).
These considerations being taken into account, extensive experimental
data (59-62) confirm sufficiently well the above kinetic equations. For the
sake of simplicity, stage 8 was neglected in these calculations ;i.e., it was as-
sumed that K , = 0.
For an unsupported catalyst with specific surface area 0.15 m2/gpromoted
with selenium (0.001 at. %), the constants at 240C for 1 g of catalyst have
the following values: (1 atm = 1 kgf/cm2): k , = 1.83 cm3 C2H4 (NTP)/g
sec atm; k2 = 0.57 cm3 C2H4 (NTP)/g sec atm [see Ionov et al. (62)];corre-
sponding activation energies (average values for experiments at 1- 16 atm)
are El = 19 kcal/mol, E2 = 21.5 kcal/mol [see Ostrovskii et al. (61)].The
other constants are: M = 3.6, L = 1.O (irrespective of temperature), K3 = 37
atm-', K , = 245 atm-', K6 = 37 atm-', and K , = 6.1 atm-' (at 240C).
The values of K 3 , K , , K 6 , and K7 are proportional to &IRT where Q = 7
kcal/m~l.~
For the same catalyst with the same surface area promoted with chlorine,
at 240C, k , = 0.81 cm3 C2H4 (NTP/g sec atm, k2 = 0.33 cm3 C2H4
(NTP)/g sec atm (average values for experiments at 1-16 atm) [see Ionov
et al. (62)].The other constants have the magnitudes given above.

XIII. Hydroxylamine Synthesis

Catalytic synthesis of hydroxylamine from nitrogen oxide and hydrogen


is widely used in industry as a constituent part of caprolactam production.
The reaction is conducted in aqueous sulfuric acid solution saturated with
NO and H, at 40C and a pressure of approximately 1 atm. Platinum sup-
ported on porous graphite, in the form of fine particles suspended in the
intensely stirred solution, is used as a catalyst, The main direction of the
reaction is

2N0 + 3Hz + HzS04 = (NH,OH)Z.HISO,. (251)


Along with hydroxylamine, by-proddcts, namely ammonia and nitrous
oxide, are formed
2N0 + 5Hz + HzSOL = (NH,),SO, + 2HzO (252)
2 N 0 + Hz = NzO + H 2 0 . (253)
This value, which had been calculated from insufficient data, disagrees with an estimate
from (26). Q = 22 kcal/mol.
240 M. I. TEMKIN

Our kinetic experiments were performed at atmospheric pressure. The gases,


NO and H,, were fed into solution through glass filters. The solution was
agitated very intensely: 3000 rpm of the stirrer. Energy dissipation at stirring
was measured by determining the rate of heating of the solution caused by
the rotation of the stirrer. From this, using the correlation based on the work
by Sherwood et al. (78),the rate of diffusion of dissolved gases to the external
surface of catalyst particles (the size of which was less 0.1 mm) was calculated.
It was found that external diffusion did not limit the reaction rate. A calcu-
lation using (10) showed that the reaction rate was not affected essentially
by diffusion in catalyst pores either. These conclusions were confirmed in
experiments with catalyst particles of different size and in experiments with
variable intensity of stirring. The latter experiments have also demonstrated
that the reaction rate was not limited by the rate of dissolution of gases.
Thus the results may be considered as referring to the kinetic region (at 25C
and lower temperatures).
The catalyst contained 10% platinum deposited on activated carbon.
This high platinum content (20 times more than in the commercial catalyst)
made the activity of the catalyst more stable, thus facilitating the studies of
kinetics. The surface of platinum in 1 g of the catalyst measured by hydrogen
chemisorption was 2 m2. Since the concentrations of H, and NO in the

0.1 0.2
PNO/otm

FIG.3. The rates of hydrogenation of NO to NH,OH(r)), NH,(r@)), and N,0(r(3)) on


platinum in 10% sulfuric acid solution; 25C, PH1= 0.35 atm.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 24 1

solution are determined by their partial pressures over the solution, P,, and
P,, , we shall consider reaction rate as a function of PHIand P,, .
The kinetics of the reaction was found to be unusual. The rate of hy-
droxylamine formation, r('), and that of ammonia formation, I-(,),at con-
stant P,, and increasing PN, , has a maximum at the PNovalue that we shall
denote as (PNO)max. This value does not depend on P,, and temperature, and
equals 0.11 atm. At P,, = 2(PN&,axthe formation of NH,OH and NH3
virtually ceases (Fig. 3).
The rate of formation of nitrous oxide by (253), I - @ ) ,features quite a dif-
ferent behavior, viz., it permanently increases with increasing PNo (Fig. 3).
The regularities observed are accounted for by the following scheme of
reaction mechanism :
"1) "2' N(3)

1. Z + H2 = ZH, $ $ 4
2. ZH2 + NO -+ ZH,NO 1 1 1
3. ZHZNO + ZH, 4 2 2 + (H) + H2NOH 1 0 0
4. 2(H) + Z ZH2 f -f f (254)
5. ZHZNO + ZH, + Z + ZOH + NH3 0 1 0
6. ZOH + (H) Z + H 2 0
-+ 0 1 0
7. 2ZH2NO 2(H) + H2O + N2O + 2 2
-+ 0 O f

N"'; NO + 3Hz = HzNOH


; + $H, = NH3 + H z 0
M 2 ) NO
N'3'; NO + f H 2 = f N 2 0 + fH,O

Scheme (254) is based on the assumption of the formation of chemisorbed


intermediate H,NO that covers a considerable part of the surface. Along
with hydrogen adsorbed on definite surface sites, ZH, , adsorption of hydro-
gen by dissolution in the subsurface layer at a depth of the order of the atomic
radius is assumed. This corresponds to the s-type (solution type) hydrogen
adsorption by Horiuti and Toya (79). Hydrogen adsorbed in this form is
denoted in the scheme as (H). Hydrogen in the (H) form does not occupy
surface sites and, therefore, it is of no consequence that its quantity corre-
sponds to the region of medium coverages (as follows from electrochemical
studies). The present formulation of the mechanism differs from that given
in Savodnik et al. (77) in that the intermediate (H) is introduced instead of
intermediate ZH. This does not alter the deduction of kinetic equations and
is done for the better agreement of the scheme with the data on hydrogen
adsorption on platinum, and in particular, with the results of the studies of
the behavior of platinum as a hydrogen electrode in the H,SO4 solution
242 M. I. TEMKIN

(80).The coverage of the surface with H, and OH particles is assumed to


be small; thus,
[ZH,NO] + [Z] = I (255)
where [ZH,NO] and [Z] are the fraction of the surface covered with H,NO
and the fraction of the free surface, respectively. The surface is assumed to
be uniform. It follows from electrochemical data that the surface manifests
induced nonuniformity (81) with respect to the form of adsorbed hydrogen
denoted as (H). However, this is of no consequence for the kinetics of the
reaction under discussion since in elementary reactions, 2, 3, and 5, which
determine the kinetics, (H) does not take part as a reactant. Stoichiometric
coefficients were chosen in such a way that the rates along the routes are
measured by the rate of expenditure of NO.
The system of equations obtained from the stage steady-state conditions
can be solved without simplifications (it is reduced to a quadratic equation).
However, a simple approximate solution is sufficient, based on the rate of
nitrous oxide formation, r(3),being much less than the sum (r(')+ r(')) of the
rates along the routes N")and N2).
From the steady-state conditions of stages 2, 3, and 5

r(2' =
r5 (258)
it follows that
r2 = r3 + r5. (259)
[The relationship (259) expresses the steady-state condition of the inter-
mediate ZH,NO.] Since

it follows from (259) that


[ZH2N0] = Ckdk3 + k5)]CN0. (263)
Stage 1 is at equilibrium; therefore
[ZHZI = K,[ZIC",.
Equations (257), (261), (263), and (264) give
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 243

and
= [k2/(k3 + k5)]aNO*
Using a similar method we find, from (258) and (262), that
r(') = k(2)P,,PNo(l - AP,,)
where
/v2'= k(1)(k5/k3).
The equations obtained are approximately correct also if the formation
+
of N,O is not neglected since r(3)< (r(') r(')). The steady-state condition
of stage 7
r ( j ) = $k7[ZH,NOIZ (270)
leads to the equation
+3) = k(3)p2
NO

where
k'3' = {3k7[k2/(k3
1
+ k5)laNO}2.
The value of (PNO)max quoted above corresponds to A = 4.6 atm-'. Experi-
ments gave the following values of other kinetic constants at 25C expressed
in units of mol NO/(cm2 Pt)(hr)(atm)z :k(')= 3.4; k(') = 0.65;
/d3)= 0.75. Solid curves in Fig. 3 were calculated from (268), (269), and
(271) with the use of these constants; the points show the experimental
results. It can be seen that the equations describe the general character of the
kinetic relationships. Experimental results that are not given here confirm
the proportionality of r'') and r(') to PH,.
Scheme (254) does not predict explicitly any dependence of reaction rates
r('), d2), and r'3)on the concentration of H,SO, and product salts. It is
24.4 M. I. TEMKIN

known, however, that the solubility of a gas in aqueous electrolyte solution


depends on concentration of the electrolyte; this is the so-called salting out
effect that can be described as the result of a change in the activity coefficient
of the dissolved gas. For our treatment of the kinetics involving the presenta-
tion of rates as functions of PHIand PN0 rather than CH2and CNo,only the
dependence of activity coefficientsof activated complexes on concentrations
of H,SO, and product salts (NH,OH),+SO:- and (NH4)z'SO:- is im-
portant for the rates of the reaction stages. Experiments demonstrated that
the concentration of H,S04 does not markedly affect the reaction rate
(provided that it is not less than 2 wt%) nor does it affect the increase in the
concentration of salts.

XIV. Reaction of Methane with Steam

Interaction of methane with steam on a nickel surface is the basis of the


natural gas reforming process. The process is used as a source of hydrogen
for ammonia production and methanol synthesis and, therefore, finds a
large-scale industrial application.
The reaction
CHL + H2O = CO + 3Hz (272)
is endothermic. In connection with this, two versions of the process are used
in the industry. In the case of steam-oxygen reforming, the gas feedstock
contains oxygen; heat evolved in the combustion of a fraction of methane
compensates the endothermic effect of the main reaction. In the case of
reforming in the tubes, heat is supplied from the outside of the tubes. Here
we shall consider only the methane reforming in the tubes (i.e., in the ab-
sence of oxygen).
At high temperatures, such as 9OO"C, the effect of entropy increase in
reaction (272) resulting from the increase in the number of gas molecules
outweighs the effect of enthalpy increase due to the reaction ;consequently,
the reaction can proceed in the forward direction to its end under atmospheric
pressure, or even at pressures of some tens of atmospheres. At lower temper-
atures (e.g., 300C) the reaction under atmospheric pressure can go com-
pletely in the reverse direction. In the intermediate temperature range and
at pressures near 1 atm, the substances entering (272) coexist at equilibrium
in comparable amounts.
The stoichiometry of the reaction of methane with steam on nickel is
described by (272) together with the equation of the water-gas shift reaction :
CO + H 2 0 = C02 + Hz. (273)
We studied methane reforming kinetics at atmospheric pressure. The uti-
lization of circulation flow systems made it possible to use nickel foil as
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 245

a catalyst (82-84); this was a guarantee that the reaction proceeded in the
kinetic region. The reaction on a commercial type catalyst (i.e., nickel
supported on porous alumina) proceeds in the internal diffusion region
(85). Here we shall discuss only the kinetics of the reaction not complicated
by diffusion.
At 900C the rate of the reaction as measured by the amount of CH,
consumed, r, is satisfactorily described by a simple first-order equation (82) :
r = k,Pc,,. (274)
Here k, is the rate constant of the irreversible first-stage high temperatures.
The stage occurs at impacts of a gas-phase CH, molecule on a catalyst sur-
face that is supposed to be practically free.
In the range of temperatures and pressures where the reaction is sub-
stantially reversible, the kinetics is much more complicated. There is no
grounds to consider chemical changes described by (272) and (273) as inde-
pendent, not interconnected, reactions. Conversely, if processes (272) and
(273) occur on the same surface sites, then free sites will act as intermediates
of both processes. Thus one must use the general approach, treating (272)
and (273) as overall equations of a certain single reaction mechanism. But
if a reaction is described by two overall equations, its mechanism should
include at least two basic routes; hence, the concept of reaction rate in the
forward and reverse directions can be inapplicable in this case. However,
experiments show that water-gas equilibrium (273) is maintained with suffi-
cient accuracy in the course of the reaction. Let us suppose that the number
of basic routes of the reaction is 2; then, as it has been explained in Section
VIII, since one of the routes is at equilibrium, the other route, viz., the route
with (272) as overall equation, can be described in terms of forward, r + , and
reverse, r - , reaction rates. The observed reaction rate is then the difference
of these
r = r+ - r- (275)

and the first problem in the kinetic studies is the determination of the
average stoichiometric number of the reaction, 5. According to (83),

where K is the equilibrium constant of (272), which is well known. The


knowledge of 5 makes it possible to pass to r+ values from experimentally
determined r values, and the dependence of r + on partial pressures of
reactants is easier to analyze.
In order to find J (83), we measured the rate of methane consumption
at P C O P A 2 / P C H 4 P H 2 06 K that is approximately equal to r+ and the rate
246 M. I. TEMKIN

of methane formation at PCOPi2/PCH4PH20 B K that is approximately equal


to r - . The experiments were done at 530C. Experimental results can be
summarized by the following equations :

where g(pH,, pH,,) and h(PH,,pHzo) are some, not yet disclosed, functions
Of pH, and PHI,. Equation (277) is valid at PCH4 < 0.2 atm and (278) holds
at pco, < 0.1 atm (total pressure is 1 atm).
It follows from (277) and (278) that r - / r + cc P& since r + a PCH4 and
r - does not depend on PCH4. According to (276), this means that ij = 1.
Equations (277) and (278) also show that r - / r + a P,, since r - a P,, and
r+ is independent of P,,. Again, it follows from this result and (276) that
ij = 1.
For a further verification of this conclusion, we determined equilibrium
constant, K , by combining the results of kinetic measurements in the region
of r + > r - and in the region of r+ < r - , having assumed that J = 1. The
agreement of the obtained value with the value known from thermodynamic
data ( K = 0.0364 atm at 530C) has confirmed that B = 1. This result
means that stoichiometric numbers of all nonequilibrium stages of the
reaction equal unity.
After this preliminary study, we measured the rates of methane reaction
with steam (i.e., r values in the r + > r - region), and the rates of carbon
monoxide hydrogenation [i.e., ( - r ) values in the r+ < r- region] at 470,
530, 600, and 700C (84). In these experiments the PCH4and Pc, values
did not exceed 0.2 and 0.1 atm, respectively ; therefore, (277) and (278) were
applicable. Variations in catalytic activity were taken into account on the
basis of control experiments.
It was found in all experiments that the ratio P ~ ~ , P ~ , / P in ~the~ P ~ , ~
exit gas mixture was near to the equilibrium constant of (273) at the reac-
tion temperature.
The data obtained can be interpreted with the following reaction mech-
anism :
jq) E J ( 2 )
1. +
CH4 Z @ ZCH, H, + 1 0
2. ZCH, +
H2O Ft ZCHOH H, + 1 0
3. ZCHOH Ft ZCO HZ + 1 0
(279)
4. zco Ft z co + 1 0
5. +
Z HzO ZO H, + 0 1
6. ZO + CO = Z + CO, 0 1

N; CH, + HZO = CO + 3H2


CO + H,O
M2); = CO, + H,
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 247

Stage 1, if assumed to be irreversible and taking place on a nearly free


surface, explains the reaction kinetics [Eq. (274)] in the high temperature
range (82). Stages 2 and 3 correspond to the notions of Eidus and Zelinskii
(86,87) on the mechanism of catalytic synthesis of hydrocarbons from CO
and H,. Stages 5 and 6 reproduce the mechanism of reaction (273) that
will be discussed in detail in Section XVII.
To correlate kinetic equations obtained from scheme (279) with experi-
mental (277) and (278), the degrees of surface coverage with CH,, CHOH,
and CO particles must be assumed small and only that for oxygen, [ZO],
to be comparable with the fraction of free surface, [Z]. For simplicity
catalyst surface will be regarded as uniform.
Since stage 5 is at equilibrium,
EZ0] = KS(PH20/PH2)[ZI (280)
and since
[ZI + [ZOI = 1, (28 1)
we have

Basic route N ( ' ) is a block (i.e., has with basic route N(') no common stages);
therefore, (71) may serve for the derivation of the kinetic equation, with
only the stages of the block included.
The rates of corresponding elementary reactions are expressed as follows :

rl = k1PCH4rZ1 r - l = k-lP~,[ZCH2]
r2 = k2PHzo[ZCH2] 1-, = k-,PHZ[ZCHOH]
r3 = k,[ZCHOH] r _ 3 = k-3PH,[ZCO] (283)

r4 = k4[ZCO] r - 4 = k-,Pc0[Z].

We substitute in (71) s, = 1, s, = 2, s3 = 3, and s4 = 4. Then all con-


centrations of intermediates cancel except [Z], which is known from (282).
We obtain

r+ =
+
k l k2k3k4PCH,PH20
(k2k3k4P,20 k-lk3k4PH2+ k- lk-2k4P& + k-lk-2k-3PA2)
. (284)

c1 + KS(PH20/PHz)I

Denoting, as before, equilibrium constant of reaction (272) as K, we have,


248 M. I. TEMKIN

in compliance with the requirements of thermodynamics,


K = k 1k2 k3 k4lk - 1 k - 2 k - 3 k - 4 . (285)
According to scheme (279), 5 = 1, so, it follows from (276), (284), and (285)
that
k - k - 2k - 3k -4PcoP;z
.
r- =
+ +
(k2k3k4PH,o k - k3k4PHz k - k - 2k4P& + k- k - 2 k - 3PA2)
(286)

L1 + KS(PH,dPH2)I

Let us introduce the following notations :


I, = k-,/k2

1, = k - , k _ 2 / k 2 k J
l3 = k-lk-2k-3/k2k3k4.
Then (284) and (286) are transformed into the following

Equation (288) becomes (274) if I , = l2 = I3 = K s = 0 (since at high tem-


peratures r = r + ) .
Handling the experimental data for the temperature range of 470-7WC,
we used theoretical expressions for preexponential factors from the theory
of absolute rates of surface reactions (Section IV), assuming the elementary
surface reactions under consideration to be adiabatic, and also the known
K values. Computer calculations were performed by the method of minimiz-
ation of the sum of the squares of relative deviations of calculated reaction
rates from experimental values.
It turned out that l1 is so small that the term 11pH2 may be neglected.
According to (287), this means that stage 2 of mechanism (279) is at equi-
librium. With this simplification we obtain from (275), (288), and (289)
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 249

The values in this equation are expressed as functions of temperature as


follows :
k - 2.38 x 1 0 2 1 T - 3 e - 3 3 7 2 0 / R T
1 -

At this the reaction rate, r, is the amount of CH, reacted per square meter
of the apparent surface of nickel per hour, measured in cubic centimeters
reduced to 0C and 760 mm Hg; R is the gas constant in cal K-' mol-'.
Kinetic equation (290) is in satisfactory agreement with experiments
both at r > 0 and r < 0. Equation (290) can be obtained not only from
mechanism (279). Some authors have developed a carbide theory of catalytic
syntheses of hydrocarbons from CO and H2 (88,89).In terms of this theory,
the following mechanism involving carbon chemisorbed on nickel as an
intermediate should be accepted instead of mechanism (279).
"1'

1. CH4 Z + G ZCH2 + H2 1 0
2. ZCH2 # ZC + H2 1 0
3. ZC + H,O F? ZCO + Hz 1 0
4. zco # z + co 1 0 (292)
5. Z H2O + ZO + H2 0 1
6. ZO CO + = Z + C02 0 1

MI'; CH4 + H20= CO + 3H2


"'); CO + H 2 0 = CO, + H2
Assuming as before that of all surface coverages only [ZO] is substantial,
we obtain the following equation for the rate along route N('):

Equations (287)are again used for the definition of 11, 12, and l3 in this equa-
tion via rate constants of elementary reactions of scheme (292). At l , = 0,
(293) becomes (290). This corresponds to stage 2 of scheme (292), being at
equilibrium. Thus the reaction kinetics does not provide a possibility for
giving preference to the Eidus-Zelinskii theory or the carbide theory.
The notions of the reaction mechanism presented here are in better agree-
ment with the results of kinetic measurements if the simplifying assumption
250 M. 1. TEMKIN

of the surface being uniform is not made and instead the surface is assumed
to be evenly nonuniform [see Chap. 10 of the contribution by Snagovskii
and Ostrovskii (2)].

XV. Ammonia Synthesis-Simple Kinetics

Ammonia synthesis is one of the most important processes of chemical


industry; tens of millions of tons of this product are synthesized annually
in various countries of the world. On a commercial scale the reaction is
operated on promoted iron catalysts at temperatures near to 500C and
high pressures, mostly at 300 atm. At present K 2 0 , A1,0,, and CaO in
amounts of several parts by weight per 100 parts of catalyst are usually
employed as promoters. The application of high pressure is caused by the
reversibility of the reaction ; molar fraction of ammonia corresponding to
the equilibrium
N, + 3H, = 2NH; (294)
increases with increasing pressure. Modem ammonia synthesis reactors can
produce 1500 metric tons of ammonia per day. The manufacturing of am-
monia synthesis reactors requires large investments; so process engineers
tend to make the best use of the inner space of the reactors. Knowledge of
the kinetic features assists in the solving of this problem.
From the viewpoint of the experimenter, the ammonia synthesis reaction
is an advantageous subject for kinetic studies since it proceeds only in one
direction without any by-products; the activity of catalysts is usually suffi-
ciently stable, it being an important condition for the success of kinetic
investigations.
Many works were devoted to the mechanism and kinetics of ammonia
synthesis. In accordance with the outline of this article, only the main results
of studies of ammonia synthesis by the author with co-workers will be pre-
sented here; other works will be cited, but in connection with our studies.
The discussions of kinetics and the mechanism of ammonia synthesis in the
reviews published in this series (90)and in the contribution by Nielsen et al.
(91) can serve as supplementary sources of information. The contribution
by Malina (92) contains interesting historical data and an extensive list of
references.
The kinetics of reaction (294) at the conditions of its commercial realiza-
tion answers to the following mechanism :

2. ZN, + 3H2 = Z + 2NH3 1

N2 + 3H2 = 2NH3
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 25 1

Stage 1 (i.e., nitrogen chemisorption) determines the reaction rate. Stage


2 is a sum of several simple stages united into one complex stage since they
all are at equilibrium and since intermediates of nitrogen hydrogenation
such as ZNH, etc., occupy only a small part of the surface. The surface
concentration of nitrogen corresponds to the region of medium coverages ;
the surface is nonuniform. Mechanism (295) is a particular case of mechanism
(1 88). According to (148)
rn
r+ = k*PN2/PN2 (296)
and according to (157)
r- = k,pi2 (297)
where P,, is N, pressure in the gas phase andp,, is the fugacity of adsorbed
nitrogen (i.e., the pressure of N, in the state of ideal gas that would corre-
spond to adsorption equilibrium with the surface at the present coverage).
Equilibrium of stage 2 determines p N 2; therefore,
PN2 = (1/q(piH3/p&) (298)
where K is equilibrium constant of reaction (294). It follows from (296),
(297), and (298) that

where

and

Since m + n = 1, we obtain from (301) and (302)

We note also that according to (301) and (302)

The observable reaction rate


r = k+PN2(Pi2/PiH3)m - k-(P&a/P3~2)1-rn. (305)
At equilibrium r = 0, taking (303) into account, we obtain from (305)

PkH3/PN2Pi2 = (306)
252 M. I. TEMKIN

in accordance with the mass action law for equilibrium. Since k - and k ,
are bound by (303) and K values for different temperatures are well known,
only two constants in (305) are to be determined for a given catalyst from
experiments, viz., m and one of the values, k , or k - ; then other constants
can be calculated by means of (303) and (304).
In the original derivation of (305), it was supposed (40) that the nitrogen
adsorption equilibrium on the catalyst follows the logarithmic isotherms
(i.e., that the surface is evenly nonuniform). In this case y = 0 and, according
to (143) and (164), m = a, n = /?.Experiments with iron catalyst promoted
with A1,03 and KzOgave m = 0.5. This was interpreted as a = 0.5 (93).
It was later demonstrated that if the reaction mechanism corresponds to
scheme (295) and the linear relation between standard Gibbs energy of
adsorption and Gibbs activation energy of adsorption is obeyed [see (91)],
then the kinetic (305) corresponds in general to the exponential nonuni-
formity of the surface with even nonuniformity included as a particular
case (44). In the general case the exponent m is not equal to transfer coeffi-
cient a, but is connected with it according to (143).
A study of adsorption-chemical equilibrium
ZNZ + 3Hz = Z i-
2NH3, (307)
which corresponds to stage 2 of mechanism (295), done with iron catalyst
promoted with A1,0, and K,O (49) showed that the logarithmic isotherm
was followed. Thus the original approach was correct as applied to this
catalyst. In ammonia synthesis on cobalt (95, 96) and nickel (96) catalysts,
m equals 0.2 and 0.3, respectively. A study of equilibrium (307) showed that
on cobalt and nickel catalysts the Freundlich isotherm with y = 0.3 for
cobalt and 0.2 for nickel was valid. Thus, in both cases a = m + y = 0.5 as
for iron (94). This result is in agreement with the natural assumption that
the differences in m values for catalysts should result from differences in
surface nonuniformity, but not in the transfer coefficienta.
We note that the very possibility of observing the equilibrium (307) sup-
ports the notion that the reaction rate is determined by stage 1 of mechanism
(295). At temperatures somewhat lower than that of ammonia synthesis,
the reaction rate of this stage becomes negligible, but the equilibrium of
stage 2 still is established sufficiently rapidly.
The concepts underlying the derivation of (305) were also confirmed in the
extensive study by Scholten (97). This author used a balance of special design
that permitted following the weight changes of the ammonia synthesis
catalyst in the course of the reaction and in the process of nitrogen adsorp-
This proof has been quoted by the author in his paper for the 5th International Congress on
Catalysis. Unfortunately, during the preparation of the proceedings of the Congress, the
corresponding pages were lost and are absent from the published text (94).
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 253

tion so to establish the dependence of the rate of adsorption of nitrogen on


its amount on the surface and to compare the rate of adsorption of nitrogen
and the reaction rate at equal coverages.
At integrating (305) for the conditions of a flow system (93,98),it proved
to be convenient to introduce a constant k proportional to k - . The value
of k was also calculated from data obtained in circulation flow systems (4,
96, 99-203). If the volume of ammonia reduced to 0C and 1 atm, formed
in unit volume of catalyst bed per hour, is accepted as a measure of reaction
rate, then k = (4/3)3(1-m)k-(101). The constancy of k at different times of
contact of the gas mixture with the catalyst and different N,/H, ratios in
the gas mixture can serve as a criterion of applicability of (305). Such con-
stancy was obtained for an iron catalyst of a commercial type promoted
with Al,03 and K 2 0 at m = 0.5 (93) from our own measurements at at-
mospheric pressure in a flow system and literature data on ammonia syn-
thesis at elevated pressures up to 100 atm. A more thorough test of applic-
ability of (305) to the reaction on a commercial catalyst at high pressures
was done by means of circulation flow method (99),it confirmed (305) with
m = 0.5 for pressures up to 300 atm. Similar results were obtained in a large
number of investigations by different authors in the USSR and abroad.
These authors, however, have obtained for some promoted iron catalysts
m values differing from 0.5. Thus, Nielsen et al. (204) have found that
m % 0.7.
Since k is proportional to k - , the temperature dependence of k is deter-
mined by the apparent activation energy of ammonia decomposition, E - :
k a exp( -E - / R T ) (308)
where R is gas constant and T is temperature.6 In numerous studies of the
reaction on iron catalysts, E- values obtained were close to that found in
the original work (93,i.e., ca 40 kcal/mol. According to laboratory data
for a commercial type catalyst with m = 0.5, k = 6.6 x lo4 atm0.5hr-' at
300 atm and 500"C, reaction rate being expressed as stated above (98).
Equation (305) describes the ammonia synthesis rate not only on iron
catalysts, but also over molybdenum catalyst (105), tungsten (106), cobalt
( 9 9 , nickel (96),and other metals (107). Equation (300) describes ammonia
decomposition on various metals (provided that there is enough H, in the
gas phase).
Data on the rate of synthesis or decomposition of ammonia on a number
of metals give activation energies of ammonia decomposition, E - , close to
40 kcal/mol, as in the case of iron catalysts, and m = 0.5 (107).

In a number of works the dependence of the type k a T-' exp ( - A / R T ) was used. A and
E- do not differ greatly. E - = A - RT, where T, is the average temperature (101).
254 M. I. TEMKIN

According to (302) and (159),


k- = (n/sin nn)[y/(eyf -l)]~-~a~~K-", (309)
Since c 0a exp( -E_O/RT) and a. oc exp(qO/RT)where E-O and qo are
the activation energy of desorption and heat effect of adsorption on most
strongly adsorbing sites, respectively, and K cc exp( - AH/RT) where AH
is the enthalpy change in reaction (294), we obtain, neglecting the tempera-
ture dependence of the factor (n/sin nn)y/(eyf - l),
E - = E-O - nqo - nAH. (310)
It is natural to assume for metals with m = 0.5 that m = a ; i.e., the surface
is evenly nonuniform as it has been found for iron. Therefore, n = p, and
the Constance of E - in passing from one metal to another means that
E-O - pqo = const. (31 1)
or, since
E-O = E+O + qo (312)
where E , 0s the energy of adsorption on the most strongly adsorbing sites,
E,' + crqo = const. (313)
Equation (313) corresponds to the rule of transfer; viz., the higher the energy
of adsorption, the lower the activation energy of adsorption; the change in
the activation energy being the fraction, a, of the change in energy of adsorp-
tion. Thus, the rule of transfer holds true, at least approximately (as E - values
do not coincide exactly), in passing from one metal to another (not only for
different sites of the same nonuniform surface). This result substantiates the
application of the rule of transfer to the discussion of the problem of a catalyst
of the highest activity (35).
Ozaki et al. (108) showed that the assumptions underlying the derivation
of (305) can be verified by comparing the rate of ammonia synthesis (294)
to that of deuteroammonia synthesis
Nz + 3Dz = 2ND3. (314)
Indeed, the constant k, in (296) and (297) characterizes only the rates of
adsorption and desorption of nitrogen and is thus one and the same for both
reactions. The difference in the rates of these reactions results only from the
nonidentity of the fugacities of adsorbed nitrogen pN2and&, ,the last being
determined for reaction (314), instead of (298) by the equation
Ppl;, = (1/~*)(~~D3/~;2)9 (3 15)
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 255

where Ky is the equilibrium constant of reaction (314). The values of K


and K* are known (109);hence, if the concepts under discussion are valid,
the kinetic isotopic effect of the substitution of deuterium for hydrogen
(protium) can be predicted. Denoting rate constants for reaction (314) as
k+* and k - * , we obtain from (301) and (302)
k+*lk+ = (K*/Klrn (3 16)
k - * / k - = (K/K*)". (3 17)
K y is larger than K ; e.g., at 475"C, K = 2.54 x atm-' and K* =
16.08 x atm-2. Therefore, it follows from (316) and (317) that the
rate of reaction (314) should be higher than that of reaction (294) under the
same conditions, in contrast to the usually observed retardation of reactions
at the substitution of D for H.
Our experiments (110)conducted at atmospheric pressure with promoted
iron catalyst in the temperature range of 400-475C have completely con-
firmed the theoretical deductions. It was found that for the catalyst used in
the experiments m = n = 0.5, so that for the ratio of constants k (propor-
tional to k -) the equality k/k* = (K*/K)0.5should hold true. This was found
to be the case; e.g., at 475C the ratio k/k* from experimentation is equal
to 2.53 and 2.47, while (K*/K)0.5= 2.52.
The authors cited above (108)obtained only a qualitative agreement of
experimental results with theoretical prediction, viz., the rate of reaction
(314) was higher than that of reaction (294), but the increase in the reaction
rate at the substitution of D for H was less than expected. This can be ex-
plained by the temperatures of the experiments being 21 8-302"C, which is
considerably lower than the usual temperature of ammonia synthesis. In
this temperature range, on a level with adsorbed nitrogen, large amounts
of intermediate products of hydrogenation of nitrogen are present on the
surface of iron catalysts in the process of ammonia synthesis (IZZ,112).
Thus, one of the premises of (305), viz., the assumption that the surface
coverage is virtually determined solely by the amount of adsorbed nitrogen, is
not fulfilled. This assumption is correct, however, at usual temperatures of
ammonia synthesis.
The integration of (305) for a flow system (98)with the assumptions that
the laws of ideal gases are applicable, the catalyst layer is isothermal, and
the plug flow is realized, gives for m = 0.5 and the gas mixture of stoichio-
metric composition
k = +
P0.5V1(l z,)[J(z~)- J(z,)] (318)
where P is total pressure, Vl is space velocity at inlet, z1 and z2 are molar
256 M. I. TEMKIN

fractions of NH, in the gas mixture at the reactor inlet and outlet, respect-
ively, and
~ ( -i z p 5 d z
J(z) =
(1 + Z)Z[L2(1 - z)4 - Z Z ] .
(319)

L? denotes a value determined as follows:


L? = 0.25 x 0.753P2K (320)
whence

where z, is molar fraction of NH, at equilibrium. The product Vl(l + zl)


in (318) can be replaced by the equivalent product V2(1 + z 2 ) where V2 is
the space velocity at the outlet. To evaluate the integral in (319), the follow-
ing approximate formula may be used :

J ( z ) x -i(1 - z ) , . ~In

At z values small as compared to z,, (322) differs but little from the following:
J(z) x +(Z/L)2. (323)
Originally (93),because of an inaccuracy, an equation was obtained (for
z1 = 0) where the denominator of the integrand corresponding to that
+ +
given in (319) contained not (1 z), but (1 z). This inaccuracy was de-
tected by Emmett and Kummer (113), but these authors did not perform
the calculations rigorously either and obtained the factor (1 + z),, which
was used in subsequznt articles (107,114,115) until the correct value was
obtained; i.e., (1 + z ) ~(98). These differences are minor since z is small
compared to unity.
At high pressures the effects that are not taken into account in (305) are
manifested, viz. the deviations from the laws of ideal gases and the effect of
pressure on the reaction rate depending on the volume change at activation.
The equation that is strictly applicable at high pressure (114) contains, in
contradistinction to (305), not partial pressures, but fugacities of gaseous
participants of the reaction. The right-hand side of the equation includes,
moreover, a factor exp[ -( Vg - a V,,,)P/RT] where V, is the partial molar
volume of the activated complex of nitrogen adsorption, VZN2is the partial
molar volume of adsorbed nitrogen, P is total pressure (the equation was
derived for m = a). When k value is calculated from the simple (305) its
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 257

change with pressure observed in the experiment corresponds to the theo-


retical expectation (99, 114). The change is not large, e.g., at 300 atm k is
equal to 0.7%' where ko is k value at low pressures (98).
In industry large pellets of a catalyst were employed (e.g., 6-8 mm in
size), and the rate of the process was essentially affected by the slowness of
the diffusion of ammonia in the pores of the catalyst; the efficiency factor
at this size of pellets is about 0.5. The effect of diffusion retardation of the
ammonia synthesis was studied both at high pressures (99), when the free
path of molecules is much smaller than the radius of catalyst pores so that
the bulk diffusion is operative, and at pressures near to 1 atm (116), where
there is a transition from the bulk to the Knudsen diffusion.
Recently the industry turned to smaller catalyst pellets or to pellets con-
taining, along with narrow pores determining the surface area of the catalyst,
wide transport pores; in this way the efficiency factor approaches unity.
In practical calculations for designing of the commercial equipment at
not negligible diffusion retardation, satisfactory results are obtained from
(305) (which is rigorously applicable only in the kinetic region) with a
diminished value of rate constant.
As early as the original paper (93),the kinetic equation (305) was applied
to calculating the optimum temperature distribution in the catalyst bed of
the ammonia synthesis reactor. Later (305) was widely used in the designing
of commercial ammonia reactors in the USSR and abroad. Methods of
the calculations for ammonia synthesis reactors with heat exchangers of
various types, based on (305), have been elaborated (91,117). Computer
programs for converter calculations are now available (117). The calcula-
tions for the first ammonia synthesis reactor with utilization of the reaction
heat is an example of the application of (305) (118). It is the process of
ammonia synthesis that, owing to the existence of a simple and sufficiently
accurate kinetic equation, was used as a model in the development of the
methods of designing catalytic reactors. Later these methods were greatly
elaborated, owing to the progress in computers and applied to a variety of
processes.
Equation (305) is unique in chemical kinetics in that it holds true at pres-
sures varying by a factor of 2000; viz., from 1/4 atm (116) to 500 atm (115)
(with high pressure effects taken into account). As far as I know, no other
reaction has been studied in such a broad pressure range.

XVI. Ammonia Synthesis-Compl icated Kinetics

Although (305) is sufficient for the description of ammonia synthesis


kinetics under the conditions of commercial production, some essential
258 M. I. TEMKIN

questions pertaining to the reaction mechanism had remained unanswered


when this equation was established. In the derivation of (305) given above,
it was supposed that nitrogen is adsorbed in the molecular form (being bound
chemically with the surface at the expense of rupture of one or two out of
the three chemical bonds in the molecule). It can be also supposed that
adsorption is accompanied by dissociation of nitrogen into atoms ; both
assumptions equally lead (107) to the kinetic equation (305). Thus, the nature
of adsorption process remained undisclosed. Furthermore, since the stage
of hydrogenation of adsorbed nitrogen is at equilibrium and any equilibrium
is independent of the mechanism of its attainment, (305) supplies no possi-
bility to conjecture on the mechanism of the hydrogenation of adsorbed
nitrogen. Answers to these questions were obtained in the course of later,
more detailed, studies of kinetics of ammonia synthesis.
It is easy to see that (305) cannot be correct at PNH, = 0 since in this case
it gives r = co. The unapplicability of (305) at low ammonia concentrations
may result either from the diffusion retardation that limits the increase in
reaction rate or from a change in kinetics. The first case is actually observed,
but only the second case is of interest for the elucidation of the reaction
mechanism. Here two possibilities can be envisaged a priori. First, with the
decrease of PNHp, the equilibrium of stage 2 of mechanism (295) is shifted
to the right to such an extent that the reaction passes from the region of
medium coverages to that of small coverages (107). For a practically free
surface the reaction rate, evidently, must be

where k is a constant.
The second possibility is that at small PNHsthe equilibrium of stage 2 is
not established during the reaction. Therefore, the reaction rate is deter-
mined not only by the rate of the adsorption of nitrogen, but also by the
rate of hydrogenation of adsorbed nitrogen. In this case the reaction pro-
ceeds in the region of medium surface coverages with nitrogen, but the
degree of covering corresponds not to the equilibrium but to a steady state
with equality of the rates of adsorption and hydrogenation of nitrogen. Then
the reaction rate should depend not only on PN,, but also on PH2.
An experimental study of kinetics of ammonia synthesis on iron ( l o ] ) ,
cobalt, and nickel (96) catalysts, at ammonia concentrations much lower
than that at equilibrium, showed that at pressures of the order of 1 atm the
second of these possibilities is realized. When far from equilibrium, the

The first possibility can become realistic at high pressures since at high partial pressures of
hydrogen the rate of hydrogenation is large.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 259

reaction rate on any of the three metals is described by equations of the fol-
lowing form :
r z k'P;,P',;"', (325)
where n' is a constant satisfying the inequalities 0 < n' < 1, instead of (324)
that can actually be formally considered as a particular case of (325) at
n' = 1. Equation (325) results from (217) if A, is identified with N, and A,
is identified with H, . This means that at small concentrations of NH, the
reaction rate is determined by two slow irreversible stages, as in a two-stage
scheme (53); they are followed by equilibrium stages that may be united
into one complex stage:

1. Z + N2 + ZN, 1
2a. ZN, + H, -P ZN,H, 1 (326)
2b. ZN2H2 + 2H2 = Z + 2NH3 1

N2 + 3H2 = 2NH3
Stages 2a and 2b together comprise stage 2 of mechanism (295). To have
the change in the rate of stage 2a at passing from one surface site to another
dependent only on the energy of adsorption of N, and independent of the
energy of adsorption of the diimide radical N,H,, one must accept that the
latter energy is the same on all sites. It seems that in general the smaller the
energy of adsorption, the smaller its change at passing from one site to an-
other. The variations in energy of adsorption can be supposed equal only for
substances that are adsorbed with approximately equal strengths.
From this interpretation of (325), it follows that m' = m where m' = 1 - n'
and m is the exponent in (305) found for the same catalyst from the reaction
rate in the region where (305) applies.
This conclusion was verified by studies of kinetics on each of the above-
mentioned metals of the iron group at very low and near to equilibrium
concentrations of ammonia. It was completely confirmed that for an iron
catalyst m = 0.5 and m' = 0.5 (101), for cobalt m = 0.2 and m' = 0.2 (96),
for nickel m = 0.3 and m' = 0.3 (96).
Schemes (295) and (326) can be united into one general scheme of the
mechanism. It is sufficient to regard stages 1 and 2a as reversible:

1. Z + N, ZN, 1
2a. ZN, +
Hz # ZN2Hz
2b. ZNzH2 + 2H2 = Z + 2NH3
1
1
(327)

N, + 3H, = 2NH3
260 M. I. TEMKIN

Equation (21 1) is applicable to the reaction rate determined by stages 1 and


2a of mechanism (327). It is only necessary, as the adsorbed radical N,H,
is a product of stage 2a, to substitute for the partial pressure of this radical
in the kinetic equation, its fugacity, p N 2 H 2 , determined by the equilibrium
of stage 2b:
P N ~=
H K&H2(pkH3/p$2)
~
where K N z H 2 is the constant of the gas equilibrium
N,H, + 2H2= 2NH3. (329)
Thus, a substitution should be made in (211) in accordance with the
following:

After simple transformations the equation obtained is reduced to the


following :

The constant k, has the same meaning as in Section XV; i.e., it is related
to k+ by (301); the constant I is determined by the equation
1 = u-o/u;, (33 1)
so that the term l / P H 2 is the ratio of rates of desorption and hydrogenation
of adsorbed nitrogen.
We note that although, according to scheme (327), the reaction includes
two slow stages, 1 and 2a, so that the concept of the rate-determining stage
is inapplicable to this case, (330) still corresponds to
r - / r + = (1/K)(piH3/pN2pi2)-
In compliance with the treatment given in Section VIII, this is explained
by the stoichiometric numbers of both slow stages being equal to 1 so that
the average stoichiometric number 5 = 1.
Let us make, instead of a supposition used in deriving (305) that stage 1
is rate-determining, a more general assumption that almost all adsorbed
nitrogen molecules react with hydrogen and only a small portion of these
molecules are desorbed. This means that
l/PH2 4 1. (332)
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 26 1

If PNHlin the reacting system is near to equilibrium value, then (l/K)(PiH3/


PN2P;,) is close to unity and the term l/PH2can be neglected as compared
not only with 1, but also with (1/K)(p&/pN2p;2). Then from (330) we ob-
tain (305). Alternatively, if PNH,is so small that, regardless of inequality (332),
(1/K)(p~H~/pN2p~~) 4 1/pH2 9 (333)
then (330) is transformed into

i.e., we obtain (325) and see that


k' = k,l-*. (335)
With (335) it was found from experimental data that 1 = 1.77 x l o v 2atm
at 450C. The constant I increases with temperature according to the
Arrhenius equation with effective activation energy E, = 26.0 kcal/mol(102).
Measurements at pressures from 0.5 to 1 atm have shown (201)that (330)
describes well the kinetics in the region that is transitional from (334) to
(305).
A later study by ICI researchers (129) at pressures up to 351 atm led to
the conclusion that (330) describes the reaction rate in the kinetic region
much better than (305), and at the same time, at the conditions of the com-
mercial process it is very close to (305). This work gave rn = 0.465; i.e., near
to 0.5.
These results, among them the equality of rn and m' found for different
metals, support the concept of the reaction mechanism represented by
scheme (327); i.e., they show, first, that nitrogen is adsorbed in a molecular
form and, second, that the addition of the first hydrogen molecule is the
slowest stage of hydrogenation of adsorbed nitrogen, the addition occurring
at the impact of the hydrogen molecule from gas phase upon an adsorbed
nitrogen molecule, without preliminary adsorption of hydrogen (102).This
was confirmed by Tamaru (120) who found that the rate of hydrogenation
of adsorbed nitrogen is proportional to the partial pressure of hydrogen in
gas phase rather than to the amount of adsorbed hydrogen.
The character of the chemisorption of nitrogen can be also judged from
the results of studies of ammonia synthesis kinetics at the reversible poison-
ing of the catalyst with water vapor (102,103).If a gas mixture contains water
vapor, an adsorption-chemical equilibrium of adsorbed oxygen, hydrogen
gas, and water vapor sets in on the iron catalyst.
We shall consider the reaction in the region where for the unpoisoned
catalyst (305) is valid. In the case of the poisoning with water vapor the reac-
262 M. I. TEMKIN

tion mechanism is described as follows :

1. 2 + NZeZN2 1
2. ZNZ + 3H2 =2NH, +Z 1 (336)
3. ZO t H, = Z t H,O 0

We need to find the rate of adsorption of nitrogen by the surface that, apart
from nitrogen, contains oxygen in comparable amounts. We shall use
Assumption 3 formulated in Section IX; i.e., we shall assume that the change
in AG,' at passing from one surface site to another is the same for N2 and
0. Then we may apply (180) substituting for P the pressure of nitrogen,
P,, , and for p and p' the fugacities of adsorbed nitrogen, pN2,and oxygen,
p o , as determined by the equilibria of stages 2 and 3; i.e.,
PN2 = (1/Q(piH3/pAz) (337)
and
Po = (1/K0)(pH20/p"2) (338)
where KO is the equilibrium constant of the reaction
0 + HZ = H 2 0 . (339)
Thus we find that

where C = c K / K , . Since, as before, the average stoichiometric number


v = 1,

Equation (341) becomes (305) if PHzo= 0. Equation (341) describes very


well the results of experiments with iron catalysts promoted with A1203 +
K,O, A1203 + K,O + CaO, or A1203 alone (102,103). In deriving (341)
the supposition that nitrogen is adsorbed in a molecular form is substantial
since, with respect to the surface coverage, atom 0 plays the same role as
the molecule N2. Therefore, the agreement of (341) with experiments shows
that in ammonia synthesis nitrogen is adsorbed in a molecular form. This
conclusion was confirmed by the data of J. H. Block et a!. described in the
communication made by Sastri (121).
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 263

Equation (341) solves the task of quantitatively describing the effect of


water vapor, and also oxygen gas (the last being rapidly converted to water
vapor at the conditions of the reaction), on the activity of commercial am-
monia synthesis catalysts. This result is of practical importance for ascer-
taining the necessary degree of purity of the inlet gas mixture with respect
to poisons containing oxygen (122).
In contrast to poisoning with water vapor, the poisoning of ammonia
synthesis catalysts by hydrogen sulfide is irreversible. It was studied using
H,S labeled with Hi5S (123) that made the radiochemical technique applic-
able to the determination of sulfur content in catalysts.

XVII. Carbon Monoxide Conversion

The reaction
CO + H 2 0 = CO, + H, (342)
called carbon monoxide conversion or water-gas shift reaction is used on
a large scale for production of hydrogen in ammonia plants. In this pro-
cess carbon monoxide obtained by natural gas reforming is used (Section
XIV). The process comprises two steps. The main amount of CO is con-
verted at 400-500C on a so-called high-temperature catalyst. But at these
temperatures, as follows from the conditions of equilibrium of reaction
(342), a sufficiently complete CO conversion cannot be achieved. A supple-
mentary conversion is effected at 200-250C on a so-called low-temperature
catalyst.
The high-temperature catalyst is prepared in the form of ferric oxide,
Fe,O3, to which some chromium oxide, Cr,03, is added to prevent the
thermal sintering of the catalyst in operation. Prior to the reaction ferric
oxide is reduced to Fe304 with hydrogen or carbon monoxide (in the pres-
ence of steam in excess). When reaction (342) occurs on the high-temperature
catalyst, the catalyst contains iron in the form of magnetite, Fe,O,, as an
x-ray phase analysis showed (124).
Our first study of reaction (342) kinetics (124) was done in a flow system
under atmospheric pressure at 400, 450, and 500C on a Fe,O, catalyst
without any additions and also on a mixed oxide catalyst of the nitrogen
type, the active component of which is Fe,O, .
In this work the following reaction mechanism was proposed :
1. Z+H,O#ZO+ H2 1
2. zo + co P Z + co, 1 (343)

CO + H 2 0 = C02 + H,
264 M. I. TEMKIN

The two-stage mechanism of CO conversion became a prototype of a


general two-stage mechanism discussed in Sections VII and X that later
proved applicable to numerous heterogeneous catalytic reactions as is evi-
denced by the review by Boudart (125) and the present article.
For the understanding of the mechanism of reaction (342), the results are
essential of studies adsorption-chemical equilibrium (Section XI) of oxygen
on Fe,04 surface with H, and H 2 0 in the gas phase (50); i.e., of the equi-
librium
2 + HZO E ZO + Hz (344)
that corresponds to stage 1 of mechanism (343).
In these experiments a sample of Fe304obtained by reduction of Fe(OH),
was placed in a perforated ampule suspended on spring made of copper-
beryllium alloy. The change in weight of the sample was followed by change
in the extension of the spring. The experiments were carried out at 400C.
The higher the ratio P,,/PHz0in a mixture of H, and H,O vapor, the
greater the decrease in weight of the sample; it agrees with the diminution
of the amount of H2 in the gas phase owing to the formation of H,O and
at a large excess of H, corresponds, in the order of magnitude, to a mono-
atomic layer.
The results of the measurements were represented as Po,values that would
correspond to equilibrium with the surface at a given amount of oxygen
taken off the surface; these values were calculated from the equation
= [KH,0(pHzO/pHz)12 (345)
where K H z o is equilibrium constant of the reaction'
H2OCp.r) = H, + ioz * (346)
The points representing the data plotted in the coordinates log Po,vs. mass
of 0 in the surface layer lie on a straight line. Thus the logarithmic adsorption
isotherm is obeyed. A particular feature of this system is that the state of
free surface, 8 = 0, corresponds to a complete removal of oxygen from the
monoatomic surface layer of the Fe,04 lattice and the state of occupied
surface, 8 = 1, corresponds to the oxide with the stoichiometric composi-
tion in the surface layer. Thus Z in (344) is an oxygen vacancy on the surface
of a magnetite crystal.
Later the kinetics of the process on a high-temperature catalyst of the
type commercially employed at present was studied (126). The catalyst
contained, prior to reduction, 93% Fe,O, and 7% Cr20, ; specific surface
area measured after kinetic experiments was near to 20 m2/g, bed density

* In the article referred to (50) KHlowas by an oversight defined as K &


INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 265

was 1.37 g/cm3, the size of catalyst grains in the experiments was from
0.5 to 1 mm. The determination of the effective diffusion coefficient made
it possible to ascertain that at this size of the grains the reaction occurred
in the kinetic region. The measurements were done in a flow circulation
system under atmospheric pressure and at 390,420,450, and 483C. It was
assumed in compliance with scheme (343) that the average stoichiometric
number, 7, equals unity, and on the basis of (85) and the observed reaction
rate, r, the forward reaction rate, r+ , was calculated from the equation

(347)

The data obtained agree with the equation


kPH20PC0
r+ = (348)
APH20 + pCOl
where k and A are temperature-dependent constants. Therefore,

r= k(PH20PC0 - K- 1pC02pH2)
(349)
APH20 + pC02
Such an equation was proposed earlier than our work (129, but later a
more complex one was substituted (128). Later other authors advanced
an equation practically equivalent to (349) (Z29).
A comparison between scheme (343) with a general two-stage scheme
(53) shows the following correspondence :

Scheme(53) A, B1 A2 B, I
Scheme(343) H20 H2 CO CO, 0

l o g A = -(8800/4.57T) -t 2.31. (350)


266 M. I. TEMKIN

Let the reaction rate be expressed as the volume of CO reduced to 0C


760 mm Hg that reacts per second in a unit volume of the catalyst bed and
let partial pressures be in atmospheres. The results of the measurements give
log(k/sec-' atm-') = -(34000/4.57T) + 10.3. (351)
Equations (350) and (351) pertain to the reaction on a catalyst 93% Fe,O, +
7% Cr,O, in the kinetic region. On large catalyst grains used commercially
the reaction proceeds with a considerable diffusional retardation.
The first study of reaction (342) (124) led to a kinetic equation that corre-
sponds to the region of medium coverages. This equation results from (213)
'
if it is assumed that A K - % 1 and hence, A % 1 (since in this temperature
range K > 1, e.g., at 450C K = 7.7). After simplifications we obtain
* = k+pco(p"201pH2Y- k-pco2(p"*/p"20)1-n (352)
where
n=l-m
k+ = kKn/A (353)
k- = kK'-"/A.

Experimental results of this work (124) were found to agree with (352) at
n = 0.5. Since according to the data on adsorption-chemical equilibrium
(3441, the logarithmic isotherm is valid; i.e., y = 0, (158) gives n = p. There-
fore, as usual, a = p = 1/2.
The difference between the results obtained in the first (124) and the
subsequent (126) works consists not only in the region of surface coverages
to which the kinetic equation corresponds, but also in the order of magnitude
of the constant A . It has been explained in Section XI that A is the ratio of
rate constants of adsorption directions of the first and the second stages.
According to (350), A -= 1 in the temperature range of 400-500C; whereas
for the deduction of (352) it must be assumed that A % 1.
The source of this discrepancy is unknown to us. Equation (349) is, un-
doubtedly, adequate for the description of the reaction kinetics on an iron-
chromium oxide catalyst. The fact that in one of the works (124) magnetite
without the addition of chromium oxide served as a catalyst can hardly be
of consequence since a study of adsorption-chemical equilibrium (344) on
an iron-chromium oxide catalyst (7% Cr,O,) (52) led to the value of the
average energy of liberation of a surface oxygen atom that practically coin-
cides with that found earlier (50) for an iron oxide catalyst with no chromium
oxide. It may be suspected that in the first work (124) the catalyst was poi-
soned with sulfur of H,S that possibly was contained in unpurified COz
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 267

from a gas cylinder. Carbon dioxide was employed for preparation of CO


according to the reaction
c + co2= 2 c o
and also was added in some experiments to the inlet gas mixture.
A study of reaction kinetics on a low-temperature catalyst (130) was
performed under atmospheric pressure at 150, 175, 200, and 225C. The
composition of the catalyst was (moles) :
ZnO + 0.24 Cr,O, + 0.24 CuO
and additives in parts by weight per 100 parts of the main components:
MnO - 2, MgO - 2, A1,0, - 5.
The pellets of the commercial catalyst were crushed to grain size from 0.5
to 1 mm. A calculation on the basis of the measurements of the effective
diffusion coefficient showed that the reaction proceeded in the kinetic region.
Bed density of the catalyst was 1.23 g/cm3, specific surface after kinetic
experiments was 36 m2/g. In the temperature range of 150-225C reaction
(342) is practically irreversible. The experiments proved (348) to be valid;
thus, the kinetics on low- and high-temperature catalysts is the same.
Constants A and k for the low-temperature catalyst are described by the
equations
A = 2.5 1 0 9 e - 2 1 , 5 0 0 / R T (354)
k = 6.0 x IOlle--26,800/RT atm-1 sec- (355)
the reaction rate being expressed as for the iron-chromium oxide catalyst
(see above).
The form of the kinetic equation suggests that the reaction mechanism
on the low-temperature catalyst does not differ from the mechanism on the
high-temperature catalyst; i.e., is described by scheme (343).

XVIII. Isotopic Exchange between Steam and Hydrogen

The reaction
HDO + H, = HD + H20 (356)
or the reverse finds its application in the production of heavy water.
The commercial catalyst is nickel on a support, but reaction (356) can
also occur on surfaces of a number of other catalysts. Its kinetics was studied
268 M. I. TEMKIN

on Ni, Co, Pd, Cu, Ag, Au, Pt, and Fe,04 (131, 132). Experiments were
done in a flow (131)and a circulation-flow (132) systems under atmospheric
pressure. Hydrogen was saturated at various temperatures with water vapor
containing from 2.2 to 4.5 at.% D. Depending on the activity of a catalyst,
the temperature range of experiments was chosen within the limits of 200-
500C. Measurements were conducted in the kinetic region.
In the theoretical discussion of the kinetics we shall suppose that contents
of D in water vapor and hydrogen are small (as it was in our experiments).
Then the gas mixture contains practically no D 2 0 and D2molecules and
only reaction (356) occurs.
The results of kinetic studies conform with the notion of a reaction pro-
ceeding on a evenly or exponentially nonuniform surface in the region of
medium coverages as described by the following mechanism :
1. Z + H D O # Z O + H D 1
2. Z O + H, e Z +HzO 1
(357)

HDO + Hz = H D + H z 0

Scheme(l88) A, B1 uz A, b, B, I
Scheme(357) HDO HD 1 Hz 1 HzO 0

Equations (196) and (197) give for the reaction rate, r = rA - rB :


r = k+PHDO(PH2iPH20)m - k-PdPH,oiPH,)l-m (359)
Rate constants, k, and k-, are related to the equilibrium constant of reac-
tion (356) by
k,lk- = K. (360)
If D content is not supposed to be small, then to obtain an equation for
the rate of reaction (356), (213) must be employed. This gives
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTION6 269

It follows from the physical meaning of the constant A (as explained in


Section XI) that for the reaction under consideration A is about unity; the
value of K is also of the same order.
For this reason, if inequalities (358) are followed, (361) becomes (359) with
k
k+ = (/lK-l)l-m

Equations (362) and (363) agree with (360).


Equation (359) with m = 0.5 was obtained empirically by M. G. Slin'ko
from experiments with a nickel catalyst. Starting from this result the general
equation (359) was obtained theoretically for reaction (356) with exponent
m not necessarily equal to 0.5, but of some value between 0 and 1, depending
on the nature of the catalyst. In this form (359) was confirmed for all studied
catalysts; obtained values of m did not depend much on temperature. The
theoretical K values (133) were employed in the calculations after they were
checked experimentally. The values of m and absolute (i.e., calculated for
unit area) k+ values for the same catalyst obtained in flow and circulation
flow systems coincided within the accuracy of kinetic measurements. The
table below gives approximated m values for some catalysts.

Catalyst Ni Co Pd Cu Fe,O, Ag
m 0.5 0.5 0.3 0.3 1.0 0.8

At PH2 = PH20 the reaction rate is independent of m so that k, characterizes


directly the activity of catalysts in the mixture of equal volumes of hydrogen
and steam. Catalysts form the following series according to the absolute
values of k + at 180 and 300C :
Pt > Ni > Co > Pd > Cu > Fe,O, z Ag > Au.

For nickel, activation energy determining the temperature dependence of


k,, E ,= 12 kcal/mol; at 150C k, x 2 x lo6 mol/atm m2 sec.
The kinetics of tritium exchange between steam and hydrogen
HTO + H2 = H2O + HT (364)
was investigated on porous nickel at 1.65-200C and nickel supported on
chromium oxide at 100- 132C (Z35). An equation corresponding to (359),
i.e.,
- k -Pn~(Pn201Pn2 m,
r = k + PnTo(Pn2/PH2oIrn (365)
was verified with m = 0.5.
270 M. 1. TEMKIN

In handling experimental results the constant of equilibrium (364) was


calculated on the basis of the literature data (136) substantiated by the meas-
urements made by the authors (135).
The stages of isotopic exchange mechanism (357), if we do not distinguish
between isotopes H and D, coincide with the forward and reverse directions
of stage 1 of mechanism (343) of carbon monoxide conversion. The reactions
of isotopic exchange corresponding to stage 2 or this mechanism

co, + COi8 = co + coo'8 (366)


and
c1402+ co = c140 + co, (367)
can also be observed. The kinetics of reaction (366) was investigated on
Fe,O, (137) and that of reaction (367) on other catalysts as well (138).
On Fe,O, the rates of isotopic exchange reactions (356), (366), and (367)
are close to the. rate of carbon monoxide conversion, as should be expected
from mechanism (343).

XIX. Phosgene Synthesis

Phosgene is used in the production of diisocyanates, starting materials


for polyurethanes, and in other organic syntheses. Commercially, phosgene
is obtained from carbon monoxide and chlorine

co + c1, = COCI, (368)

with activated charcoal as a catalyst. The reaction is conducted at 80- 150C


(temperature in various parts of the commercial reactor is different) and at
pressures near to atmospheric.
An experimental study of kinetics of reaction (368) on activated charcoal
(139) preceding our investigation was carried out at 30-65C (i.e., at tem-
peratures below those used in industrial reactors). This promoted us to
reinvestigate the kinetics of this reaction (140).
The utilization of a flow circulation system for laboratory experiments
was made difficult by the toxicity of all three gases, CO, Clz, and COCl,.
For this reason, in contrast to other reactions, we studied reaction (368) in
a closed system. The experimental apparatus was equipped with a circula-
tion pump so that the gas mixture continuously passed through the catalyst
bed in the reactor. The reaction rate was determined from the change in total
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 27 1

pressure, P,with time, t. The following formula was used in the calculations :

- - - -2P(t - 2h) - P(t - h) P(t


dP(t) + + h) + 2P(t + 2h) (369)
dt 1Oh
where h is a small time interval between successive measurements of P. This
formula corresponds to the least-squares polynomial of the second degree
interpolation of five experimental points (141).
The experiments were done at 70, 100, and 130C and at pressures some-
what lower than atmospheric. Under these conditions reaction (368) is
practically irreversible. Activated charcoal of the trademark "Bayer AKT-4"
ground to grain size 0.25-0.5 mm served as a catalyst. Estimation of the
efficiency factor on the basis of the determination of the effective difusion
coefficient of hydrogen in nitrogen or helium has shown that for this grain
size the results of reaction rate measurements refer to the kinetic region.
Estimation of relaxation time of the reaction rate from (67) showed the
reaction to be quasi-steady at the condition of our experiments in the closed
system.
Activated charcoal is a hydrocarbon rather than an elemental carbon
because it contains hydrogen chemically bound to carbon. We observed
that when exposed to chlorine, activated charcoal evolves HC1; the sub-
stitution of hydrogen by chlorine, i.e., the reaction of the type
RCH + CI, = RCCl + HC1 (370)
probably occurs. Analysis after experiments in which charcoal served as a
catalyst of reaction (368) demonstrated that this material contained 23
wt. % C1 corresponding approximately to the composition CloC1. Specific
surface of the charcoal measured by low temperature adsorption of nitrogen
was 1000 m2/g before the kinetic experiments and only 200 m2/g after the
experiments. It may be conceived that reaction (370) causes the closing of a
part of the micropores.
The decrease in the catalytic activity of fresh charcoal that is observed
during reaction (368) may also be related to the process (370). After 15
hours' exposure of charcoal in chlorine at 13O"C, the catalytic activity be-
comes constant. Kinetic measurements were carried out with a catalyst
prestabilized in this way.
Experimental data on the reaction kinetics agree with the kinetic equation
for a reaction on a nonuniform surface in the region of medium coverages if
the mechanism is as follows:
1. z+coF?zco 1
2. zco + c1, F? z + coc1, I (37 1)

co + CI, = COCI,
272 M. I. TEMKIN

Scheme(53) A, A, B, Bz I
Scheme(371) CO C1, - co co
Since B, is absent, I coincides with A,.
At first sight, scheme (371) does not agree with the results of our adsorp-
tion experiments; these experiments showed that activated charcoal does
not chemisorb CO at 100C. It should, however, be taken into consideration
that the surface of charcoal subjected to activation or even simply after
storage in contact with air is covered with chemisorbed oxygen. The studies
of the reactions of carbon with C 0 2 and steam (see Section XX) have demon-
strated that oxygen chemisorbed on carbon is indistinguishable from chemi-
sorbed carbon monoxide. So it may be reckoned that activated charcoal is
already covered with carbon monoxide before the contact with this gas.
Desorption of CO from the surface of such charcoal occurs only at tem-
peratures much higher than those employed in phosgene catalytic synthesis;
therefore, stage 1 of scheme (371) is irreversible under the usual conditions
of the reaction. Kinetic measurements show that the reaction is retarded
by its product, phosgene. To account for this in terms of scheme (371),
stage 2 must be considered as reversible. In such a case (214) is directly
applicable. After the substitution specified above we obtain
r = k P c o [ ~ c , , / ( A P c+
o ~coc~,)lm. (372)
Experimental data gave m = 1/4 at the three temperatures (i.e., 70, 100, and
130C).
According to (99) and (143),
m = u - (T/O) (373)
where a and 0 are temperature-independent; therefore, if m # u, m depends
on T. However, this dependence is hardly noticeable. Let us, for instance,
accept for the transfer coefficient a the most usual value, 0.5, and assume that
m = 0.25 at 100C. Then we find that 0 = 1490K; hence, at 70"C, m = 0.27
and at 130"C, m = 0.23. The deviation of m from the average value 0.25,
by k0.02 is difficult to detect.
Taking m = 0.25, we found numerical values of constants in (372) from
experimental data for activated charcoal Bayer. These constants are de-
scribed by the equations
logA = -1900/T + 3.40 (374)
log k = - 1875/T + 3.80. (375)
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 273

Values of k from (375) give the phosgene formation rate, in moles per second
per gram (200 m2) of charcoal at pressures expressed in atmospheres.
According to (379, the apparent activation energy, E = 8.6 kcal/mol.
Equation (372) with rn = 0.25 and A found from (374) agrees well with
experiments not only for the charcoal Bayer, but also for the Soviet activated
charcoal AR-3.
In the temperature range covered by the experiments by Potter and Baron
(139)one may have A = 0 as follows from (374). This means that stage 2 of
scheme (37 1) is virtually at equilibrium under these conditions. Then,
instead of (372), we have the following equation corresponding to (21 5) :
r = kPc,(Pc,,/Pc,,,,)". (376)
Experimental results of the referenced work agree with (376) ;namely, points
fall into straight lines being plotted in the log(r/Pco) vs. log(PcI2/Pc~~,)
coordinates. Slopes of the lines for most of experimental series give m values
near to 1/4.

XX. Reactions of Carbon with Carbon Dioxide and Steam

The reactions of coal gasification by carbon dioxide


c + co2= 2co (377)
and by steam
C + (1 + s ) H ~ O= (1 - s)CO + s C O ~+ (1 + s ) H ~ ; 0 < s < 1 (378)
in combination with carbon monoxide conversion (the reaction discussed
in Section XVII) were widely used as a source of hydrogen in the production
of synthetic ammonia.
At present this method is excluded by a more economical process of
methane reforming (considered in Section XIV). The exhaustion of natural
gas resources may, however, restore the importance of coal gasification.
In addition, reaction (377) is a constituent part of the blast furnace process.
Reaction (378) is employed in the production of activated charcoal.
Reactions (377) and (378) are usually not ranked among catalytic pro-
cesses, although they are, as a matter of fact, autocatalytic since they occur
on the surface of carbon, one of the participants of the reactions. True, it is
more usual to speak of autocatalysis for reactions catalyzed by products
rather than by reactants, but it is not essential in our case. Equations (377)
and (378) may be read from right to left as the reactions are reversible.
Coal gasification is accelerated by the components of ash such as potash
and iron oxide; sometimes, to promote gasification, catalytically active
274 M. I. TEMKIN

substances are added. Here we will consider only the reactions of pure carbon
with carbon dioxide and steam. The kinetics of these reactions is closely
related to the kinetics of some of the reactions discussed above.
As a result of their investigations of reaction (377), Frank-Kamenetskii
and Semechkova (142, 143) advanced a mechanism that can be presented
as follows :
1. z c + c o 2 # z c o +co 1
2. zco # z + co 1 (379)

c + co, = 2CO

In scheme (379) ZC as well as Z denotes a site on the surface of carbon; in


the symbol ZC we only accentuate one of the C atoms; viz., the atom that
binds chemically to the 0 atom. Surface formation ZCO, according to
Frank-Kamenetskii, can be equally well considered as a chemisorbed 0
atom (on ZC) or as a chemisorbed CO molecule (on Z ) .
The equivalency of ZC and Z being taken into account, scheme (379)
becomes a particular case of a simple two-stage scheme (53). The corre-
spondence is specified by the following table:
~~ ~

Scheme(53) A, B, A, B, I
Scheme(379) C02 CO - co co
In deriving a kinetic equation for mechanism (379), Frank-Kamenetskii
assumed the surface to be uniform. The kinetic equation answering to this
assumption is obtained immediately from (61):

If the reaction is conducted under such conditions that it is practically


irreversible, but is not too far from equilibrium, then in accordance with
considerations in Section XI, one of its stages must be irreversible; namely
the slowest one. Therefore, one of the rate constants of backward reactions,
k - , or k-2, should be taken as equal zero. Assuming k-2 = 0, we obtain
from (380)

and assuming k - l = 0, we obtain


INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 275

Equations (381) and (382) are kinetically indistinguishable. Dividing the


numerator and the denominator by k,, we reduce them both to the equation

where I,, I,, and I, are constants. An equation of the form of (383) was
proposed after the works by Frank-Kamenetskii also by Gadsby et al.
( 1 4 4 , who derived it from notions about the mechanism of the reaction
somewhat different from those of Frank-Kamenetskii, but still assuming
the surface to be uniform.
Equation (383) is in qualitative agreement with the experiment. Reaction
(377) is indeed retarded by its product, carbon monoxide, even when the
concentration of the product is smaller than equilibrium concentration to
such an extent that the reverse reaction is virtually absent. This was verified
also by our experiments described below. According to the meaning of
(381), the retardation of the reaction by its product results from the reversi-
bility of stage 1 ; (382) implies that the reaction is retarded by adsorption of
the product. The correct choice can be made by the comparison of the reac-
tion rate and the rate of exchange of Oi8 between CO, and CO, since of the
two reaction stages in scheme (379), only the stage 1 (when it proceeds for-
ward and backward) results in this isotopic exchange.
Our investigation of the kinetics of reaction (377) (145) was done on a
practically ash-free carbon sample. The carbon was obtained by kilning
Bakelite prepared from chemically pure phenol and formaldehyde with
subsequent activation with carbon dioxide. Specific surface area of the car-
bon measured by adsorption of methylene blue was 2.2 mz/g for the initial
sample and 3.7 m2/g for the sample that reacted with COz for 155 hours.
Surface area measurements by low-temperature adsorption of nitrogen
showed a much greater increase in specific surface with the burning off of
carbon; namely, from 24 m2/g for fresh carbon to 300 m2/g for carbon
after 155 hours of reaction with CO, ,presumably, as a result of the opening
of the orifices of micropores. Kinetic experiments demonstrated that the
rate of the reaction (377) is proportional to the specific surface measured
by adsorption of methylene blue. This means that the reaction occurs mainly
on the surface of wide pores accessible to the molecules of methylene blue
and that the surface of micropores takes virtually no part in the reaction.
The experiments were conducted in the temperature range of 700-780C.
The degree of CO, conversion in these experiments did not exceed 3%; CO
concentration at the outlet comprised only 5-6% of its equilibrium content
under the conditions of the experiment. Therefore, in discussing the results,
the reaction could be considered irreversible.
276 M. 1. TEMKIN

The size of carbon grains in the kinetic experiments was 1-2 mm. Pre-
liminary tests with grains of different sizes had shown that this size range
ensured the reaction course in the kinetic region.
The same carbon that was used for reaction (377) was employed in the
measurements of the rates of the reactions
COO'S + co = coo + C0'8 (384)
and
coo + CO'8 = coo'8+ co (385)
in the temperature range of 550-570C. The concentrations of carbon mon-
oxide and carbon dioxide corresponded in this case to equilibrium with
respect to reaction (377). The measurements were conducted in a closed
circulation system.
The rate of isotopic exchange is proportional to the difference of contents
of labeled atoms in the exchanging molecules. The factor of proportionality
was proposed (145) to be named the "total rate of exchange" since it is equal
to the rate of the continuous exchange of atoms that cannot be detected
unless the atoms are labeled. The rate of isotopic exchange would be equal
to the total rate of exchange if all atoms of an element in one of the species
were labeled and those in the other species were normal.
Total rates of oxygen exchange between CO, and CO calculated from the
rates of reactions (384) and (385) were compared to the forward rate of
reaction (377) in the equilibrium state found by extrapolation of measured
rates by means of an empirical kinetic'equation. It turned out that at 750C
the total rate of exchange exceeds the rate of reaction by a factor of 50 and
at 700C more than of 100. The difference in rates is so large that it cannot
depend on the inexactness of measurement of the rates of isotopic exchange
or on the uncertainty associated with the extrapolation of the reaction rate
to equilibrium. It follows that stage 1 of scheme (379) is much faster than
stage 2. Therefore, if one of the reaction stages becomes irreversible as a
result of going away from equilibrium, it is stage 2 and not stage 1. This
means that (381), but not (382), is qualitatively correct.
The comparison of our measurements of the rate of reaction (377) with
(383) showed that this equation is not obeyed quantitatively. The disagree-
ment of (383) with experimentation already noted in preceding works can
be explained by the nonuniformity of the surface not being taken into
account. To describe the rate on a nonuniform surface under the conditions
of practical irreversibility of the reaction, (214) can be applied. It should,
however, be remembered that this equation corresponds to the mechanism
with the irreversible first stage, and in our case it is the second stage that is
irreversible. Therefore, in order to make use of (214), the numbering of
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 277

stages should be altered; stage 2 of scheme (379) should be regarded as the


first stage and stage 1 as the second stage. Since A = ic1/x2,A - should be
substituted for A . We obtain
r = "[Pco,/(A + PC0)I" (386)
or

where k = [ k ] A m .
Equation (387) at rn = 0.5 agrees quantitatively with experimentation
(145). Constants in (387) are described as functions of temperature by the
equations
k = 3.25 x 105e-48*000/RT (388)
A = 5.3 10-11eS8.S00/RT
(389)
Here the rate of reaction, r, is expressed in units (mole CO,)/mZ hr (specific
surface area of carbon being measured by adsorption of methylene blue);
pressures are in atmospheres.
A number of authors (246-150) who studied the reaction (378) concluded
that the kinetics of reaction (378) follows the equation
r(l, = 'lPH20
1 + i2PH2+ i3PHzO (390)

where r(') is the rate of entering of carbon into the reaction; i , , i, , and i,
are constants. This equation is analogous to (383). As in (383), it corresponds
to the assumption that the surface is uniform.
It is natural to assume on the basis of mechanism (379) of the reaction of
carbon with carbon dioxide and the mechanism of carbon monoxide con-
version, (343), that the reaction of carbon with steam occurs as follows:
Arc') "2'

1. ZC + H 2 0 + H2
P ZCO 1 1
2. zco ?3z + co 1. 0 (391)
3. z c o + c o * c o 2 +zc 0 1

MI); C + H 2 0 = CO + H2
N'"; CO + H 2 0 = C 0 2 + H,
Stage 1 of this mechanism is analogous to stage 1 of mechanism (379),
stage 2 of both mechanisms are identical, and stage 3 coincides with stage
1 of mechanism (379) written in the reverse direction. On the other hand,
stages 1 and 3 conform with the mechanism of carbon monoxide conversion
(343). Such a mechanism of reaction (378) was proposed by Key (15Z).
278 M. I. TEMKIN

The reaction is described by two basic routes; in (391) they are chosen
so that carbon enters the reaction only in route Mi), and route N)results
in the conversion of a part of CO formed in route N()into CO, . With such
a choice of the basis of routes, r(l), the rate of the reaction along route N()is
the rate of carbon gasification; i.e., corresponds to symbol r(l) in (390).
Our study of the kinetics of the reaction of carbon with steam (152) was
conducted by the circulation flow method under atmospheric pressure in
the temperature range of 900- 1000C. Dilution with helium was employed
to vary the sum of partial pressures of the reaction participants. The experi-
ments were carried out with nonporous graphite of high purity (the content
of admixtures did not exceed lo-%). The roughness factor of graphite was
found to be 2-2.5 (from electrochemical measurements). Equation (390)
proved not to be obeyed quantitatively; the results of the variation of PH2
in a broad range by addition of H, to the gas mixture at the inlet do not form
a straight line in the plot l/r( vs. P,, .
Experimental data answer the equation
r( = k(1)(PH20/PH2)m (392)
with rn = 0.5.
In order to obtain (392) from the scheme of mechanism (391), the surface
of carbon should be regarded as nonuniform, as in the derivation of (387)
for reaction (377). Besides that, it should I& assumed that stage 1 of mech-
anism (391) is at equilibrium (ie., occurs much faster than stages 2 and 3).
In this case stage 3 virtually does not change the surface coverage with
oxygen, the coverage being determined by the equilibrium of stage 1. Since,
on the other hand, carbon does not participate in stage 3, this stage may be
ignored in determining r(. Thus, we have a simple two-stage reaction con-.
sisting of stages 1 and 2, stage 1 being at equilibrium. In scheme (188) stage
2 is at equilibrium. Hence, to apply the kinetic equation of this scheme, the
numbering of stages should be altered. Then we obtain (392) from (196).
According to the results of experiments at 950, 1000, 1050, and 1110C
k(1) = 4.2 1014e-76.800/RT (393)
Reaction rate is expressed here in units cm3 (CO + CO,)NTP/hr m2.
The rate of formation of C 0 2 , r), is found in a similar manner. Since
stage 1 is at equilibrium, stage 2 does not change the surface coverage and
it is sufficient to take into account only stages 1 and 3 in determining r(2).
Thus we obtain mechanism (343), only stage 2 of this mechanism is now
designated as stage 3. Since stage 1 is at equilibrium, we have the mechanism
of type (188), stages 3 and 1 of mechanism (391) corresponding to stages 1
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 279

and 2 of mechanism (188), respectively. Therefore, in accordance with (196)


and (197)
r ( 2 )= - ~ 1 / K ' 2 ~ ~ ~ p ~ O ~ p H ~ / p ~ (394)
k~2~pCO(pH~O/pHz)m[1 OpHzO~~

where k(') is a constant and K")is the equilibrium constant of route N").
It follows from (392) and (394) that (153)
P)/r(') = k(z)/k(l)[l- (I/@ z, )(pco2pH2/pcopH20)I. (395)
Equation (395) is confirmed by experimental data (154).
Constants k(') and k(') are related to constants k and A in (386); it can be
shown that A = k(2)/k(') and [k]= k(')[fi2)]-". Therefore the measurements
of the rates of formation of CO and COz in steam gasification of carbon
give the possibility of calculating theoretically the rate of gasification of
the same carbon with COz . The comparison of the results of such a calcula-
tion with the direct measurement of the rate of reaction (377) on pyrographite
at 1050C demonstrated a good agreement (154). It shows that the homo-
geneous reaction
CO + H20= COz + Hz (396)
did not play any essential role under the conditions of these experiments.
It can be explained by the homogeneous reaction (396) being a chain one
with large chain length (155). On the surface of carbon the chains terminate
and reaction (396) is strongly retarded.

XXI. Ammonia Oxidation

Catalytic oxidation of ammonia to nitric oxide is the basis of production


of nitric acid. It is also used in other processes, of which may be mentioned
hydroxylamine synthesis (Section XIII) and the chamber process for the
production of sulfuric acid. The products of the reaction are nitric oxide,
water, and nitrogen, so that the reaction can be described by the equation
4NH3 + (3 + 2~)Oz= 4sNO + 2(1 - s)N2 + 6HZO; 0 < s < 1. (397)
In this equation s is the fraction of the whole amount of ammonia reacted
that is converted into nitric oxide; i.e., the selectivity. If the process is carried
out properly, selectivity is near to unity.
In the production of nitric acid the reaction is accomplished with mixtures
of ammonia and air or sometimes with mixtures of ammonia, air, and
oxygen, containing about 10 vol % NH, at pressures of 1-10 atm (below
280 M. I. TEMKIN

the concentrations of ammonia in gas mixtures will also be stated in volume


percent). Elevated pressure and the enrichment of gas mixture with oxygen
accelerate the subsequent oxidation of NO into NOz. Gauzes made of plati-
num with several percent rhodium are used as a catalyst. Rhodium reduces
the wear of platinum during the process and improves selectivity; under
favorable conditions s x 0.97. Together with rhodium, several percent of
palladium (which is cheaper than platinum) are introduced without any
noticeable adverse effect. The temperature of the reaction on platinum
gauzes is near 850C.
Besides platinum, certain metal oxides and their mixtures may serve as
catalysts of reaction (397). Cobaltous-cobaltic oxide, Co304, is the best
among nonplatinum catalysts. The reaction on this catalyst is carried out
at 7O0-75O0C, selectivity is near 0.95.
Until now we limited ourselves to the discussion of reaction rates in the
kinetic region even though in the commercial realization of the reactions
the effect of transfer processes could have been essential. This was justified
since the allowance for transfer effects would mostly be of the character of
corrections.
The situation is different in the case of ammonia oxidation. Both on plati-
num (156) and nonplatinum (157) catalysts under the conditions of a com-
mercial process, the reaction occurs in the external diffusion region. Diffusion
of ammonia rather than of oxygen is determining the rate since the reaction
is conducted with oxygen in excess with respect to stoichiometry, as given
by (397). Concentration of ammonia at the surface of the catalyst is so small
as compared to its concentration in the gas flow that the difference of con-
centrations that determines the rate of diffusion virtually coincides with the
ammonia content in the flow.
It may be supposed that it is the low concentration of NH3 as compared
to concentration of Oz (at the surface of the catalyst) that results in high
selectivity. Taking this into consideration, it is appropriate in this case to
first consider the course of the reaction in the external diffusion region.
A characteristic feature of reaction (397) arising from the external dif-
fusion character of its kinetics is a large temperature difference between
the surface of the catalyst and the gas stream. Ammonia-air mixture fed to
the catalyst is preheated only to approximately 150Cin the case of platinum
gauzes, and is not preheated at all in the case of Co304. Converters for
ammonia oxidation contain no arrangements for heat exchange. The reac-
tion is brought virtually to the end ;thus neglecting heat losses the tempera-
ture of the gas mixture leaving the catalyst layer may be taken as equal to
the isoenthalpic temperature of the reaction, TH.This is the h a 1 tempera-
ture of the gas mixture (calculated from the reaction heat) after the reaction
is completed without heat exchange with the surrounding medium. Often
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 28 1

THis called adiabatic temperature of the reaction; this is not rigorous.


Indeed, since the reaction is carried out at constant pressure, and not at
constant volume, the enthalpy H (as in the Joule-Thomson experiment) is
constant but not the energy. However, the difference is not large. At 10%
NH, in the mixture isoenthalpic temperature increment is about 700K. It
is not far from the factual temperature increment since, owing to the excep-
tionally high reaction rate and therefore large intensity of evolution of heat,
heat losses with radiation comprise only a small fraction of heat transferred
to the gas flow by convection. Whereas the temperature of the gas mixture
increases as the mixture passes through the catalyst bed, the temperature
of the whole catalyst surface is almost uniform and is near to the isoenthalpic
temperature. Thus, at the beginning of the catalyst layer the temperature
difference between the catalyst and the gas mixture is near to 700K and at
the end of the layer it nearly vanishes. Heat conductivity of the catalyst is of
no consequence in this. The description given is equally applicable to plati-
num gauzes with high heat conductivity and to Co,O, with low heat con-
ductivity.
Such temperature regime results from the coincidence of the form of
equations determining the distribution of concentrations and of tempera-
ture in a moving fluid, the only difference being that in the first case the
equation cantains diffusion coefficientD and in the second case heat dausiv-
ity a, defined by a = I / c p p where A is heat conductivity, cp is specific heat
capacity at constant pressure, and p is density of the fluid. There is a corollary
of the kinetic theory of gases verified by experiments that states in gas mix-
tures D and a are numerically close. Thus, at 0C and 1 atm heat diffusivity
of air, a = 0.172 cm2/sec and diffusion coefficient of NH, in air, D = 0.198
cm2/sec. Therefore, the thickness of the temperature layer is approximately
equal to that of the diffusion layer.g Let us assume for simplicity that the heat
capacity of the gas mixture is constant, and neglect heat losses of the catalyst
by radiation and heat conduction. We shall denote the isoenthalpic tempera-
ture increment at the completion of the reaction as ATH.Let the temperature
of the catzlyst be higher than the initial temperature of the gas mixture by
ATH.When a certain fraction, x,of the initial amount of ammonia undergoes
oxidation, the initial concentration of ammonia in the gas stream, c, decreases
by xc. At this the amount of heat is liberated equal to the fraction x of the
complete heat of reaction, i.e, the amount capable of increasing the tempera-
ture of the gas mixture by XAT,. The initial concentration difference that
determines the rate of diffusion, c, is diminished by xc. Since the rates of the
The thickness of the diffusion layer, 6,, is defined by j = (D/G,)hc wherej is the flow of a
substance to the unit surface, D is its diffusion coefficient, and Ac is the difference of concentra-
tions in the main gas stream and at the surface. The thickness of the temperature layer is defined
in a similar manner.
282 M. I. TEMKIN

equalization of concentrations and temperatures coincide, the gas stream


removes from the surface such an amount of heat that the initial temperature
difference, ATH,is diminished by xAT,. But the amount of heat liberated
on the surface is just this. Thus, heat liberation is equal to heat removal,
and the temperature regime is steady. The steady state is stable because if
the surface temperature is varied the rate of diffusion and with it the rate of
liberation of heat will change but slightly while the removal of heat will in-
crease if the surface temperature is increased or decreased in proportion to
the temperature difference. In both cases surface temperature will return to
the initial value.
Since the thickness of the diffusion layer decreases with increase in the
linear velocity of the gas stream, it does make a difference whether the gas
mixture passes through the catalyst bed of thickness h at a velocity w or
through the bed of thickness 2h at a velocity 2w. In the second case the rate
of the reaction is higher. Therefore, in contradistinction to the reactions in
the kinetic or internal diffusion region, in ammonia oxidation time of con-
tact (or space velocity) does not determine the fraction of ammonia reacted.
The characteristics of the process conditions must include the thickness of the
catalyst bed and the linear velocity of the gas stream or some other equivalent
values. For the process on gauzes it is convenient to state the number of
gauzes with definite characteristics (thickness of wires and density of weav-
ing) instead of the thickness of the catalyst bed. Instead of linear velocity
of the stream, it is convenient to employ a value proportional to it that will
be called the load. The load is defined as the ratio of the amount of
ammonia-air mixture fed into the reactor per unit time, to the cross-section
of the catalyst bed. Since the content of ammonia in the feedstock is almost
constant, it is allowable to state not the amount of ammonia-air mixture
but the amount of ammonia. This is more convenient from the viewpoint
of process engineering, for the load so defined determines more directly the
capacity of the reactor.
With increased load the thickness of the diffusion layer and that of the
temperature layer decrease equally and the balance is maintained as long as
the reaction rate is determined by external diffusion. But at a certain, high
enough load, owing to the limitation imposed by the rate of the reaction on
the surface of the catalyst, the concentration of ammonia at the surface
ceases to be small and the reaction starts to pass from external diffusion
region to the kinetic region. Beginning with this load, the growth of heat
liberation rate falls behind the growth of heat removal. The catalyst gets
cooler and this further slows down the reaction; the cooling then grows,
and the reaction stops. One may now say that the catalyst is extinguished.
The load that extinguishes the catalyst will be called the limiting load.
It is a function of the reaction rate on the surface in the kinetic region, and
hence depends on the nature of the catalyst.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 283

A more complete discussion of the heat regime in the external diffusion


region which takes into account the fact that a and D values are not exactly
equal can be found in the literature (9).Our presentation aimed to graphically
demonstrate the main features of the phenomenon.
Because of the external diffusional character of kinetics, the degree of
ammonia conversion can be predicted theoretically as a function of geo-
metric parameters of the catalyst bed and hydrodynamic characteristics of
the gas stream.
For ammonia oxidation on platinum gauzes under atmospheric pressure,
using the analogy between diffusion and heat transfer, an equation was
obtained (156) equivalent to the following:
+
lg(co/cl) = (Sn/d~)[0.43 0.274(d~)O.~~]. (398)
Here co is NH, concentration in the initial mixture, c1 is NH, concentration
in the mixture behind the gauzes, S is geometric surface of a gauze per its
unit area, n is the number of gauzes, d is the diameter of wire in centimeters,
and w is the load of the ammonia-air mixture in liters of mixture (NTP)/hr
cm2. It is to be noted that in the mixture at 10% NH,, the value of w in
1 mixture (NTP)/hr cm2 coincides numerically with the value in m3 NH,
(NTP)/hr m2.
Equation (398) was confirmed experimentally (156). The gauzes used in
these experiments were of d = 9 x lo-, cm, S = 1.87. The number of
gauzes varied between 1 and 8, the load was changed in the range of from
55 to 4000 1 mixture (NTP)/hr cm2.
In the commercial atmospheric pressure reactors with three gauzes, the
load is about 200 m3 NH, (NTP)/hr m2. It follows from (398) that under
these conditions the fraction of unreacted ammonia is practically nil.
c0304 pellets used in practice are of 4-5 mm in size. Thus, they are much
larger than the diameter of wires in platinum gauzes. For this reason, in
contrast to the reaction on gauzes, the reaction on c0304 pellets under
atmospheric pressure is characterized by the Reynolds number much larger
than 1, the Reynolds number being defined by Re = ul/v where u is the
linear velocity of the stream, 1 is the characteristic dimension, v is the kinetic
viscosity coefficient. The thickness of the diffusion layer for such pellets is
6, 11~~0.5 (399)
(since in gases 6, is close to the thickness of the Prandtl layer, 6). Hence it
can be deduced that
Ig(c,/c,) = A(h/PW0.5) (400)
where h is the thickness of the catalyst bed and A is a constant depending on
the shape of pellets. If h and I are expressed in centimeters and w is measured
in 1 mixture (NTP)/hr cm2, then, according to laboratory measurements
284 M. I. TEMKIN

under atmospheric pressure (150, A = 4.2 for cylindrical pellets with the
diameter of base, 1, equal to the height, and A = 2.5 for spherical pellets
with the diameter 1. These A values refer to the downward stream of the gas
mixture usually employed in the reactors with the nonplatinum catalyst.
Taking a co/cl ratio such that there would be practically no diffusion of
ammonia, e.g., co/cl = lo4, we obtain from (400) the required thickness of
the bed of pellets.
Let us consider now the effect on the process of passage from atmospheric
pressure to a higher pressure P, with load, w ,and the size of catalyst pellets,
I, unchanged. The Sherwood number, Sh = l / h D , is determined by the
Reynolds, Re = ul/v, and the Schmidt, Sc = v/D, numbers. The value of
w being given, v, D, and u values are inversely proportional to pressure, P.
Therefore, 6, is independent of P at constant w and 1. Since u and D are
inversely proportional to P, the increase in the time of contact of the gas
mixture with the catalyst by a factor of P as a result of passage from 1 atm
to P atm is compensated by the P times diminition of the diffusion coefficient,
and the fraction of ammonia reacted remains unchanged. We conclude that
(398) and (400) may also be used without any modification in the case of
reaction under pressures differing from atmospheric.
In ammonia oxidation, under elevated pressures higher loads are used
than under atmospheric pressure, possibly because there is no difficulty
in overcoming the resistance to the flow. For instance, at 7 atm the load is
5000 m3NH, (NTP)/hr m2 and the number of gauzes is 16. Although the
load is 25 times higher than the above-mentioned load for the atmospheric
pressure reactors, the number of gauzes is only 5.3 times larger. Therefore,
the productivity of gauzes in the reactors under elevated pressure is much
higher. It would be a mistake to think that this increase in productivity re-
sults from the increase in pressure; it results from the increase of load making
the diffusion layer thinner.
The possibility of such increase in productivity is restricted by the value
of limiting load, but for platinum gauzes this value is very high. Experiments
with gauzes characterized above (wire thickness 0.09 mm) done without
preheating of the ammonia-air mixture under atmospheric pressure and
at 10% NH, gave a limiting load about 30000 1 mixture (NTP)/hrcm2
(157). The limiting loads for nonplatinum catalysts are smaller than that for
platinum. The largest value was found for Co,O, (158). Pertinent data will
be given and discussed below.
To observe ammonia oxidation on platinum in the kinetic region, the reac-
tion must be carried out at low pressures since the decrease in pressure in-
creases the diffusion coefficient and slows down the reaction, both factors
favoring the transition to the kinetic region.
We studied the reaction under pressures about mm Hg in a static
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 285

system on an electrically heated platinum wire (159). The temperature of the


wire was 900C (160)."
Since the rate of the homogeneous reaction
2N0 + O2 = 2N02
obeys an equation of third order, the extent of this reaction under these low
pressures is negligible for the duration of the experiment. This allows one to
judge whether NO is the primary reaction product. Earlier Bodenstein
(161, 162) and Krauss (263, Z64 oxidized NH3 on Pt at low pressures with
a freezing off of the products on a surface cooled with liquid air. They
observed the formation of nitrous acid H N 0 2 and hydroxylamine N H 2 0 H .
The authors concluded that under usual conditions (i.e., at pressures of the
order of 1 atm) NO is formed as a result of the decomposition of primary
products such as HNO, in the hot gas stream.
In order to determine NO we applied the previously discovered (165)fast
heterogeneous reaction of NO with oxygen on a glass surface cooled with
liquid air. The product of this reaction has a composition intermediate
between N 2 0 3and N204.The reaction is characterized by negative activa-
tion energy; namely, its rate increases twofold at lowering the temperature
by 2 K. At pressures about lo-' mm Hg when a small glass appendix,
jointed to the system of 0.5 1 volume, is cooled with liquid air, the reaction
proceeds virtually to completion in several minutes. With the help of this
reaction it has been definitely ascertained that NO is the main product of
the reaction of NH, with O2 on Pt ; besides NO, N, is the only nitrogen con-
taining species formed. Evidently, HNO, in Bodenstein's experiments re-
sulted from the interaction of water with the product of the heterogeneous
reaction of NO with O2 that occurred on the wall cooled with liquid air.
The formation of NH,OH is more difficult to explain. In Bodenstein's
experiments NHzOH was observed at platinum temperatures above 1100C.
It may be supposed that platinum underwent dispersion. On its particles in
the condensate containing dissolved NO and H, (the product of decomposi-
tion of NH, on Pt) NH,OH could be formed by the reaction discussed in
Section XIII.
Experiments have shown that the reaction rate is approximately propor-
tional to PNH3 and is independent of Po,. A comparison of the reaction rate
with the number of impacts of NH, molecules upon the surface shows that
approximately one of 20 impacts results in the reaction. The simplest explana-
tion is that the rate-determining elementary reaction occurs at the impact
of the NH, molecule upon the surface nearly completely covered with
oxygen.
l o The previously given figure (820C) (159) was calculated without correction for the cooling
effect of the leads.
286 M. I. TEMKIN

Ammonia oxidation under low pressures, as a method of transferring


this reaction into the kinetic region, is inapplicable in the case of the c0304
catalyst since, at such temperatures as 700C and low O2 pressures, Co,O,
decomposes with the formation of COO.In order to obtain information on
the kinetic of NH, oxidation on Co304,we studied limiting loads at which
the catalyst is extinguished (166, 167). The experiments were performed at
pressures from 1 to 9 atm. A catalyst pellet was placed in a vertical tube of a
diameter such that the cross-section occupied by the pellet comprised one-
half of the cross-section of the tube. This was an imitation of conditions in
the bed of pellets. The gas mixture at the inlet was at room temperature; the
stream of the mixture was directed downward.
Cylindrical pellets of the catalyst were used with the diameter 4 mm, equal
to the height. In the experiments for determination of the dependence of
limiting load on the size of pellets, spheres with a diameter of from 2 to 5
mm were employed.
The limiting load for cylindrical pellets 4 mm in size, which worked for a
short time under atmospheric pressure and at 10% NH, in the mixture with
air, was near to 10,000 1 mixture (NTP)/cm2 hr. The measurements of the
temperature of pellets with an optical pyrometer showed that it is nearly
independent of load up to the limiting load.
The limiting load was found to be proportional to the total pressure of
the ammonia-air mixture and to the size of the pellets. The limiting load is
increased twofold with the increase in NH, content in the mixture with air
from 9 to 10%. It is also increased twofold at constant NH, content and the
substitution of oxygen for air.
From these results and applying the theory of critical ignition and extinc-
tion phenomena of strongly exothermic heterogeneous reactions developed
by Frank-Kamenetskii and Buben (9),it was found that the following form
of kinetic equation is the most probable:
r = kP$zp Pz;'4 (401)
where k is a constant, PNH3 and Po, are partial pressures of NH, and O2 at
the surface of the catalyst.
The industrial application of the Co,O, catalyst for ammonia oxidation
is complicated by its sensitivity to poisoning action of small amounts of
sulfur compounds in the ammonia-air mixture. This phenomenon was
studied with the use of radioactive sulfur containing S3' that made it possible
to measure very low concentrations of sulfur (168). Poisoning results in
decrease both of selectivity and limiting load. A noticeable decrease in selec-
tivity starts at sulfur concentrations in the gas mixture from 0.05 mg/m3.
This concentration is many times lower than minimum HzS and SOz con-
centrations in air detected by smell.
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 287

REFZRENCE.9

1. Kiperman, S . L., Vvedenie v kinetiku geterogennykh kataliticheskikh reaktsii. Nauka,


Moscow, 1964.
2. Snagovskii, Yu. S., and Ostrovskii, G. M., Modelirovanie kinetiki geterogennykh
kataliticheskikh protsessov. Khimiya, Moscow, 1976.
3. McGlashan, M. L., Annu. Rev. Phys. Chem. 24,51 (1973).
4. Temkin, M. I., Kiperman, S. L., and Lukyanova, L. I., Dokl. Akad. Nauk SSSR 74,763
(1950).
5. Temkin, M. I., Kinet. Katal. 3, 509 (1962).
6 . Sidorov, I. P., Shishkova, V. V., and Temkin, M. I., Tr. GIAP 6,323 (1956).
7 Korneichuk, G. P., and Streltsov, 0. A., Probl. Teor. Prakt. Issled. Obl. Katal. Chapter 3
(1973).
8. Ionov, Yu. V., Rastaturin, V. A., and Kurochkin, Yu. Yu., Zh. Prikl. Khim. 64,606 (1971).
9 . Frank-Kamenetskii, D. A., Diffuziya i teploperedacha v khimicheskoi kinetike, 2nd ed.
Nauka, Moscow, 1967.
10. Satterfield, C . N., Mass Transfer in Heterogeneous Catalysis. 1970.
11. Temkin, M. I., Kinet. Katal. 16, 504 (1975).
12. Thiele, E. W. 2nd. Eng. Chem. 31,916 (1939).
13. Smith, N. L., and Amundsen, N. R., Ind. Eng. Chem. 43,2156 (1951).
14. Graham, T., London Edinburgh Philos. Mag. J. Sci. [3] 2, 175, 269, and 351 (1833).
15. Herzfeld, K. F., in Lerbuch der Physik (Miiller and Pouilets, eds.), 11th ed., Vol. 3,
2nd half. Vieweg, Braunshweig, 1925.
16. Hoogschagen, H.,Ind. Eng. Chem. 47,906 (1955).
17. Scott, D. S., and Dullien, F. A., A2ChE J. 8, 113 (1962).
18. Evans, R. B., Watson, G. M., and Mason, E. A,, J. Chem. Phys. 35,2076 (1961).
19. Maxwell, J. C., London Edinburgh Philos. Mag. J. Sci. [4] u),21 (1860).
20. Langmuir, I., J. Am. Chem. Soc. 40, 1361 (1918).
21. Temkin, M., Zh. Fiz. Khim. 11, 169 (1938); Acta Physicochim. URSS8,141 (1938).
22. Langmuir, I. J. Am. Chem. SOC.17,621 (1921-1922).
23. Ternkin, M. I., Bulg. Acad. Sci., Commun. Dep. Chem. 1, No 3, 65 (1968) (in Russian).
24. Temkin, M. I., Zh. Fiz. Khim. 4, 573 (1933); Acta Physicochim. URSS 1,36 (1934).
25. Kuchaev, V. L., and Temkin, M. I., Kinet. Katul. 13, 719 and 1024 (1972).
26. Horiuti, J., and Nakamura, T., Z . Phys. Chem. (Frankfurt am Main) [N.S.] 11,358 (1957).
27. Temkin, M. I., in Mekhanizm i kinetika slozhnych kataliticheskich reaktsii (S. Z.
Roginskii, G. V. Isagulyants, and I. I. Tretyakov, eds.), p. 57. Nauka, Moscow, 1970; Int.
Chem. Eng. 11,709 (1971).
28. Slinko, M. G., Bykov, V. I., Yablonskii, G. S., and Akramov, T. A., Dokl. Akad. Nauk
SSSR 226,876 (1976).
29. Noyes, R. M., and Field, R. J., Apnu. Rev. Phys. Chem. 25,95 (1974).
30. Semenov, N. N., Tsepnye reaktsii. Goskhimtechizdat, Leningrad, 1934.
31. Temkin, M. I., Dokl. Akad. Nauk SSSR 152, 156 (1963).
32. Temkin, M. I., Nauchn. Om.Podbora Proizvod. Katal., Dokl. Vses. Soveshch., 1962 p. 46
(1964).
33. Temkin, M. I., J. Res. Inst. Catal., Hokkaido Uniu. 16, 355 (1968) (in Russian).
34: Harary, F., Graph Theory. Addison-Wesley, Reading, Massachusetts, 1969.
35. Temkin, M. I., Zh. Fiz. Khim. 31,3 (1957).
36. Temkin, M. I., Kinet. Katal. 17, 1095 (1976).
37. Temkin, M. I., Ann. N . Y . Acad. Sci. 213,79 (1973).
38. Khomenko, A. A., Apelbaum, L. O., Shub, F. S., and Temkin, M. I., Kinet. Katal. 13,251
(1972).
288 M. I. TEMKIN

39. Temkin, M. I., Dokl. Akad. Nauk S S S R 165, 615 (1965).


40. Temkin, M. I., Zh. Fir. Khim. 15,296 (1941).
41. Temkin, M. I., Kinet. Katal. 13, 555 (1972).
42. Evans, M. G . , and Polanyi, M., Trans. Faraday SOC.32,1333 (1936).
43. Zeldovich, J., Acta Physicochim. URSS 1,961 (1935).
44. Temkin, M. I., Probl. Kiriet. Katal. 6, 54 (1949).
45. Zeldovich, J., Acta Physicochim. URSS 1, 449 (1934); see also Roginskii, S. Z., Probl.
Kinet. Katal. 3, 356 (1937).
46. Bangham, D. H., and Burt, F. P., Proc. R. SOC.London Ser. A 105,481 (1924).
47. Temkin, M. I., Kinet. Karal. 16, 1461 (1975).
48. Temkin, M. I., Kinet. Katal. 5, 1005 (1967).
49. Romanushkina, A. E., Kiperman, S . L., and Temkin, M. I., Zh. Fiz. Khim. 27,1181 (1953).
50. Kulkova, N. V., and Temkin, M. I., Zh. Fiz. Khim. 21,2017 (1957).
51. Kilkova, N. V.,and Temkin, M. I., Zh. Fiz. Khim. 36, 1731 (1962).
52. Kulkova, N. V . ,Levchenko, L. P., and Temkin, M. I., Zh. Fiz. Khim. 42,2688 (1968).
53. Temkin, M. I., Dokl. Akad. Nauk S S S R 161, 160 (1965).
54. Kurilenko, A. I . , Kulkova, N. V., Ostrovskii, V. E., and Temkin, M. I., Dokl. Akad. Nauk
S S S R 123,878 (1958).
55. Ostrovskii, V. E., Kulkova, N. V., Lopatin, V. L., and Temkin, M. I., Kinet. Katul. 3,189
(1962).
56. Kayumov, R. P.,Kulkova, N. V., and Temkin, M. I., React. Kinet. Catul. Lett. 1,29 (1 974)
(in Russian).
57. Kayumov, R. P., Kulkova, N. V., and Temkin, M. I., Kinet. Katal. 15, 157 (1974).
58. Kayumov, R. P., Kulkova, N. V., and Temkin, M. I., Kinet: Katul. 15, 1349 (1974).
59. Kurilenko, A. I., Kulkova, N. V., Rybakova, N. A., and Temkin, M. I., Zh. Fiz. Khim. 32,
797 and 1043 (1958).
60. Kurilenko, A. I., Kulkova, N. V., Baranova, L. P., and Temkin, M. I., Kinet. Katal. 3,208
(1962).
61. Ostrovskii, V. E., Kulkova, N. V., Kharson, M. S.,and Temkin, M. I., Kinet. Katal. 5,469
(1964).
62. Ionov, Yu.V . ,Zhidkova, L. K., Kulkova, N. V., and Temkin, M. I., Kinet. Katal. 14,642
(1973).
63. Temkin, M. I., and Kulkova, N. V., Dokl. Akad. Nuuk S S S R 105, 1021 (1955).
64. Rudnitskii, L. A., Kulkova, N. V., and Temkin, M. I., Metody Issled. Katal. Katal. Reakt.
3, 145 (1965).
65. Rudnitskii, L. A,, Kulkova, N. V., andTemkin, M. I., Probl. Kinet. Katul. 12, 113 (1967).
66. Rudnitskii, L. A,, Shakhovskaya, L. I., Kulkova, N. V., and Temkin, M. I., Doki. Akad.
Nuuk S S S R 182, 1358 (1968).
67. Shakovskaya, L. I . , Rudnitskii, L. A,, Kulkova, N. V., and Temkin, M. I., Kine,. Katal. 11,
467 (1970).
68. Ostrovskii, V. E., and Temkin, M. I., Kinet. Katal. 7, 529 (1966).
69. Twigg, G . H . , Trans. Faraduy SOC.42,284 (1946).
70. Orzechowski, A., and MacCormack, K. E., Can. J. Chem. 32,388, 415, and 433 (1954).
71. Temkin, M. I., Zh. Vses. Khim. 0-ua. u),7 (1975).
72. Temkin, M. I., and Kulkova, N. V., Kinet. Katal. 16, 1211 (1975).
73. Margolis, L. Ya., Geterogennoe kataliticheskoe okislenie uglevodorodov (sintez mono-
merov). Gostoptekhizdat, Moscow, 1962.
74. Lyubarskii, G. D., Dokl. Akad. Nauk S S S R 110, 112 (1956).
75. Kilty, P. A,, and Sachtler, W. M. H., Catal. Re#.-Sci. Eng. 10, 1 (1974).
76. Kenson, R. E., and Lapkin, M., 1. Phys. Chem. 74, 1943 (1970).
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 289

77. Savodnik, N. V., Kulkova, N. V., Dokholov, D. M., Lopatin, V. L., and Temkin, M. I.,
Kinet. Katal. 13, 1520 (1972).
78. Brian, P. L. T., Hales, H. B., and Sherwood, T. K., AIChE J. 15, 727 (1969).
79. Horiuti, J., and Toya, T., Solid State Surf Sci. 1 (1969).
80. Shlygin, A. I., and Frumkin, A. N., Acta Physicochim. URSS 3,791 (1935); 4,911 (1936).
81. Temkin, M. I., Vopr. Khim. Kinet., Katal. Reakt. Sposobn., Dokl. Vses. Soveshch., 1955
p. 484 (1955).
82. Bodrov, I. M., Apelbaum, L. O., and Temkin, M. I., Kinet. Katal. 5, 696 (1964).
83. Khomenko, A. A,, Apelbaum, L. O., Shub, F. S., andTemkin, M. I., Kinet. Katal. 11,1480
(1970).
84. Khomenko, A. A., Apelbaum, L. O., Shub, F. S., Snagovskii, Yu. S., and Temkin, M. I.,
Kinet. Katal. 12,423 (1971); Temkin, M. I., Shub, F. S., Khomenko, A. A., and Apelbaum,
L. O., Nauchn. O m . Katal. Konvers. Uglevodorodov, 1977. p. 3 (1977).
85. Bodrov, I. M., Apelbaum, L. O., and Temkin, M. I., Kinet. Katal. 8,821 (1967).
86. Eidus, Ya. T., and Zelinskii, N. D., Izu. Akad. Nauk SSSR, Ser. Khim p. 289 (1940).
87. Eidus, Ya. T., Izv. Akad. Nauk SSSR, Ser. Khim. p 65 (1943).
88. Fisher, F., and Tropsch, H., Brennst.-Chem. 7,97 (1926).
89. Craxford, S. R., and Rideal, E. K., J . Chem. Sor. p. 1604 (1939).
90. Emmett, P. H., ed., Catalysis, 3, pp. 171 and 265. New York, 1955.
91. Nielsen, A,, An Investigation on Promoted Iron Catalysts for the Synthesis of Ammonia,
3rd ed. Giellerups, Copenhagen, 1968.
92. Malina, I. K., Razvitie issledovanii v.oblasti sinteza ammiaka. Nauka, Moscow, 1973.
93. Temkin, M., and Pyzhev, V., Zh. Fir. Khim. 13,851 (1939); Acta Physicochim. URSS 12,
327 (1940).
94. Temkin, M. I., in Catalysis, Proc. Int. Congr. 5th, 1972, (J. W. Hightower, ed.), Vol. 1,
p. F-I 13. North-Holland Publ., Amsterdam, 1973.
95. Morozov, N. M., Lukyanova, L. I., and Temkin, M. I., Kinet. Katal. 6, 82 (1965).
96. Morozov, N. M., Shapatina, E. N., Lukyanova, L. I., and Temkin, M. I., Kinet. Katal. 7,
688 (1966).
97. Scholten, J. J. F., Chemosorption of Nitrogen on Iron Catalists in Connection with
Ammonia Synthesis. Croniger, Amsterdam, 1959.
98. Temkin, M. I., Khim. Nauka Promst. 2, 219 (1957).
99. Shishkova, V. V., Sidorov, I. P., and Temkin, M. I., Tr. GIAP 7, 62 (1957).
100. Sokolova, D. F., Morozov, N. M., and Temkin, M. I., Zh. Fiz. Khim. 33,471 (1959).
101. Temkin, M. I., Morozov, N. M., and Shapatina, E. N., Kinet. Katal. 4,260 and 565 (1963).
102. Smirnov, I. A., Morozov, N. M., and Temkin, M. I., Dokl. Akad. Nauk SSSR 153, 386
(1963).
103. Smirnov, I. A,, Morozov, N. M., and Temkin, M. I., Kinet. Katal. 6, 351 (1965).
104. Nielsen, A,, Kjer, J., and Hansen, B., J. Catal. 3, 68 (19642.
105. Kiperman, S., and Temkin, M., Zh. Fiz. Khim. 20,369 (1946).
106. Kiperman, S., and Temkin, M., Zh. Fiz. Khim. 20,623 (1946).
107. Kiperman, S. L., and Temkin, M. I., Zh. Fiz. Khim. 21,927 (1947).
108. Ozaki, A., Taylor, H., and Boudart, M., Proc. R. Soc. London, Ser. A 258,47 (1960).
109. Schulz, G., and Schaefer, H., Ber, Bunsenges. Phys. Chem. 70,21 (1966).
110. Shapatina, E. N., Kuchaev, V. L., and Temkin, M. I., Kinet. Katal. 12, 1476 (1971).
I l l . Boreskova, E. G., Kuchaev, V. L., Penkovoi, B. E., and Temkin, M. I., Kinet. Katal. 13,
358 (1972).
112. Igranova, E. G.,Ostrovskii, V. E., and Temkin, M. I., Kinet.-2, Vtor. Vses. Konf. Kinet.
Katal. Reakts., Vol. 1, p. 135 (1975).
113. Emmett, P. H., and Kummer, J., Ind. Eng. Chem. 35, 677 (1943).
290 M. I. TEMKIN

114. Temkin, M. I., Zh. Fiz. Khim. 24, 1312 (1950).


115. Sidorov, I. P., and Livshits, V. D., Zh. Fiz. Khim. 21, 1177 (1947).
116. Sokolova, D.F., Morozov, N. M., and Temkin, M. I., Zh. Fiz. Khim. 33,471 (1959).
117. Kjm, J., Measurement and Calculation of Temperature and Conversion in Fixed-Bed
Catalytic Reactors. Copenhagen, 1958.
118. Fauser, G., Acta Technol. Chim. Reaz. Reattori, 6th, 1962 p. 1 (1962).
119. Bridger, G. W., and Snowdon, C. B., in Catalyst Handbook, pp. 142 and 145. Wolfe,
London, 1970.
120. Tamaru, K., Bull. Chem. SOC.Jpn. 37,771 (1964); Proc. Int. Congr. Catal., 3rd, 1%4(1965).
121. Sastri, M. V. S., Catal., Proc. Int. Congr. 5th, 1972 Vol. 2, p. 1273 (1973).
122. Smirnov, I. A., Kinet. Katal. I, 107 (1966).
123. Bulatnikova, Yu. I., Apelbaum, L. O., and Temkin, M. I., Zh. Fiz. Khim. 32,2717 (1958).
124. Kulkova, N. V., and Temkin, M. I., Zh. Fiz. Khim. 23, 695 (1939).
125. Boudart, M., AIChE J. 18,465 (1972).
126. Shchibrya, G. G., Morozov, N. M., and Temkin, M. I., Kinet. Katal. 6, 1057 and 1115
(1956).
127. Kodama, S., Fukui, K., Tame, T., and Kinoshita, M., Shokubai 8,50 (1952).
128. Kodama, S., Mazume, A., Fukuba, K., and Fukui, K., Bull. Chem. SOC.Jpn. 28,318 (1955).
129. Stelling, O., and Krusenstierna, O., Acta Chem. Scmd. 12, 1095 (1 958).
130. Cherednik, E. M., Morozov, N. M., and Temkin, M. I., Kinet. Katal. 10, 603 (1969).
131. Temkin, M. I., Nakhmanovich, M. L., and Morozov, N. M., Kinet. Katal. 2, 722 (1961).
132. Nakhmanovich, M. L., Morozov, N. M., Buadze, L. G., and Temkin, M. I., Dokl. Akad.
Nauk SSSR 148,1346 (1963).
133. Weston, R. E., Tetrahedron 6, 37 (1959).
134. Nakhmanovich, M. L., Thesis, Moscow (1963).
135. Taikhert, A. M., Morozov, N. M., and Temkin, M. I., Kinet. Katal. 4,904 (1963).
136. Black, J. F., and Taylor, H. S., J . Chem. Phys. 11, 395 (1943).
137. Kulkova, N. V., Kuznets, E. D., and Temkin, M. I., Dokl. Akad. Nauk SSSR 90, 1067
(1953).
138. Stroeva, S.S.,Kulkova, N. V., andTemkin, M. I. Dokl. Akad. Nauk SSSR 124,628 (1959).
139. Potter, C., and Baron, S., Chem. Eng. Prog. 47,473 (1951).
140. Shapatina, E. N., Kuchaev, V. L., Penkovoi, B. E., and Temkin, M. I., Kinet. Katal. 17,
644 (1976); Shapatina, E. N., Kuchaev, V. L., and Temkin, M. I., ibid. 18,968 (1977).
141. Lanczos, C., Applied Analysis. Prentice-Hall, Englewood Cliffs, New Jersey, 1956.
142. Frank-Kamenetskii, D. A., Dokl. Akad. Nauk SSSR 23,662 (1939).
143. Semechkova, A. F., and Frank-Kamenetski!, D. A., Zh. Fiz. Khim. 19,291 (1940).
I44. Gadsby, J., Long, F. J., Sleightholm, P., and Sykes, K. W., Proc. R. SOC.London, Ser. A
193, 357 (1948).
145. Evropin, V. A,, Kulkova, N. V., and Temkin, M. I., Zh. Fiz. Khim. 30,348 (1956).
146. Gadsby, J., Hinshelwood, C. N., and Sykes, K. W., Proc. R. SOC.London, Ser. A 187,128
(1946).
147. Long, F. J., and Sykes, K. W., Proc. R . Soc. London, Ser. A 193, 377 (1948).
148. Long, F. J., and Sykes, K. W., J . Chim. Phys. 47, 361 (1950).
149. Johnstone, H. F., Chen, C. J., and Scott, D. S., Ind. Eng. Chem. 44, 1564 (1952).
150. Jolley, L. J., and Poll, A., J. Inst. Fuel 26, 33 (1953).
151. Key, A., Gas Res. Board, Commun. 40,(1948); Strickland-Constable, R. F., J . Chim. Phys.
47, 356 (1950).
152. Cherednik, E. M., Apelbaum, L. O., and Temkin, M. I., Dokl. Akad. Nauk SSSR 174,891
(1 967).
153. Temkin, M. I., Cherednik, E. M., and Apelbaum, L. O., Kinet. Katal. 9,95 (1968).
INDUSTRIAL HETEROGENEOUS CATALYTIC REACTIONS 29 1

154. Cherednik, E. M., Temkin, M. I., and Apelbaum, L. O., Kinet. Katal. 9,824 (1968).
155. Shub, F. S., Apelbaum, L. O., and Temkin, M. I., Kinet. Katal. 11,566 and 11 19 (1970).
156. Apelbaum, L. O., and Temkin, M., Zh. Fiz. Khim. 22, 179 and 195 (1948).
157. Temkin, M. I., Morozov, N. M., Pyshev, V. M., Apelbaum, L. O., Lukyanova, L. I., and
Demidkin, V. A,, Probl. Fiz. Khim. 2, 14 (1959).
158. Morozov, N. M., Lukyanova, L. I., and Temkin, M. I., Kinet. Katal. 7, 172 (1966).
159. Apelbaum, L. O., and Temkin, M., Dokl. Akad. Nauk SSSR 74,963 (1950).
160. Apelbaum, L. O., and Temkin, M. I., Zh. Fiz. Khim. 33,2697 (1959).
161. Bodenstein, M., and Biittner, G., Z. Angew. Chem. 47, 364 (1934).
162. Bodenstein, M., Z . Angew. Chem. 48,327 (1935); Trans. Electrochem. SOC. 71,353 (1937);
Z.,Elektrochem. 47, 501 (1941); Z . Phys. Chem., Abt. B 50, 333 (1941).
163. Krauss, W . , Z . Phys. Chem., Abt. B 39,83 (1938).
164. Krauss, W . , and Schuleit, H., Z. Phys. Chem., Abt. B 45, 1. (1940).
165. Temkin, M., and Pyzhev, V., Acta Physicochim. URSS 1, 177 (1934); 2,473 (1935).
166. Shub, F. S., Khomenko, A. A., Apelbaum, L. O., and Temkin, M. I., Kinet.-2, Vtor.
Vses. Konf. Kinet. Katal. Reakts., Vol. 2, p. 13 (1975).
167. Shub, F. S., Khomenko, A. A., Apelbaum, L. O., andTemkin, M. I., Kinet. Katal. 17,1586
(1976).
168. Bulatnikova, Yu. I., Apelbaum, L. O., and Temkin, M. I., Zh. Fiz. Khim. 32,2717 (1958).

S-ar putea să vă placă și