Sunteți pe pagina 1din 123

Introduction to heterogeneous catalysis

Per Stoltze
Department of Chemistry and Applied Engineering Science Aalborg University

Contents
Introduction......................................................................................................................6 Definition of catalysis..................................................................................................6 Catalysis and process design........................................................................................6 Catalysis and kinetics...................................................................................................6 The basis for catalysis..................................................................................................7 Status of the study of catalysis.....................................................................................7 Chalenges................................................................................................................8 Caveats........................................................................................................................8 Reaction mechanism.......................................................................................................10 Complications ...........................................................................................................10 Kinetic equations are nonlinear...........................................................................10 Inerts.....................................................................................................................11 Nonconsecutive steps..........................................................................................11 Elusive intermediates............................................................................................11 Undetectable steps.................................................................................................12 Dead ends..............................................................................................................12 Linear dependence between reaction steps.............................................................12 Consistency................................................................................................................13 Guidelines..................................................................................................................13 Kinetics..........................................................................................................................16 Reversibility..............................................................................................................16 The reaction rate........................................................................................................17 Rate laws...................................................................................................................17 Forward and backward rate........................................................................................18 Stoichiometric matrix................................................................................................18 Example ...........................................................................................................18 Properties..........................................................................................................19 Rate..................................................................................................................19 Equilibrium equation........................................................................................20 Rate constant..............................................................................................................20 Rate limiting steps.....................................................................................................21 Most abundant reaction intermediate..........................................................................22 Reaction order............................................................................................................22 The activation energy.................................................................................................23 Activation energy vs reaction energy.....................................................................23 Graphical determination........................................................................................24 Analytical determination.......................................................................................24 Activation energy for composite reactions.............................................................24 Compensation effect..............................................................................................25 Two reactions in series .........................................................................................26 Rate and conversion...................................................................................................27 Pseudofirst order kinetics.........................................................................................27 2

Analogy between reactors..........................................................................................28 Solving the reaction scheme...........................................................................................30 The full solution........................................................................................................30 Steady state approximation........................................................................................30 The quasiequilibrium approximation.......................................................................31 Irreversible step approximation..................................................................................31 MARI approximation.................................................................................................32 Adsorption .....................................................................................................................34 Introduction...............................................................................................................34 The LennardJones picture........................................................................................34 The potential energy surface......................................................................................35 Motion of adsorbed molecules...................................................................................37 Example A2*....................................................................................................38 Catalysis.........................................................................................................................39 LangmuirHinshelwood mechanism..........................................................................40 Example................................................................................................................40 Example................................................................................................................40 Example................................................................................................................41 EleyRideal mechanism............................................................................................41 Example................................................................................................................41 Example................................................................................................................42 EleyRideal or LangmuirHinshelwood....................................................................42 The principle of Sabatier............................................................................................42 Structure sensitivity...................................................................................................44 Physisorption.............................................................................................................45 The BET isoterm.......................................................................................................45 Chemisorption...........................................................................................................46 The langmuir isoterm............................................................................................46 Nondissociative adsorption.............................................................................46 Dissociative adsorption.....................................................................................47 Competitive chemisorption...............................................................................48 Sticking.....................................................................................................................49 as a rate..............................................................................................................49 The activation energy for $\sigma$.......................................................................49 Flux and exposure.................................................................................................50 The exposure.........................................................................................................50 Temperature programmed desorption........................................................................51 Kinetics.................................................................................................................52 First order kinetics............................................................................................52 Redheads equation............................................................................................53 Second order kinetics........................................................................................55 Zero order desorption.......................................................................................56 Catalyst structure and texture.........................................................................................59 Catalyst structure.......................................................................................................59 Pore structure.............................................................................................................59 Catalyst models..........................................................................................................60 3

The terraceledgekink model...................................................................................60 Defects excluded in the TLKmodel.....................................................................61 The stereographic map...............................................................................................61 Bodycentered cubic lattice.......................................................................................63 Facecentered cubic lattice........................................................................................64 Hexagonal closepacked lattice.................................................................................66 Adsorbates and impurities..........................................................................................66 The Wulff construction..............................................................................................68 Homogeneous catalysts..............................................................................................69 Preparation of catalysts...................................................................................................70 Shape.........................................................................................................................70 Precipitation...............................................................................................................73 Pelletizing..................................................................................................................73 Fusion........................................................................................................................74 Catalyst supports........................................................................................................75 Impregnation..............................................................................................................76 Metal sponges and colloids .......................................................................................77 Poisoning and deactivation.............................................................................................78 Deactivation...............................................................................................................78 Sintering................................................................................................................78 Fouling..................................................................................................................79 Dynamic poisoning...............................................................................................80 Microkinetic modelling..................................................................................................81 Limitations.................................................................................................................82 Level of the model.....................................................................................................83 The full reaction dynamics................................................................................83 Mean field model..............................................................................................83 Types of models.........................................................................................................84 Phases in a simulation................................................................................................85 Input parameters........................................................................................................86 Test of the model.......................................................................................................87 Significant parameters...............................................................................................87 Applications of microkinetic modelling.....................................................................87 Ammonia synthesis....................................................................................................88 Solution for *......................................................................................................88 Calculation of the equlibrium constants.................................................................89 Calculation of reaction enthalpies..........................................................................90 Calculation of the rate...........................................................................................90 Practical calculation..............................................................................................91 Parameters.............................................................................................................91 Stability of intermediates.......................................................................................92 The NH3 concentration.........................................................................................92 Coverage by *...................................................................................................93 Coverage by N*................................................................................................94 Coverage by N2*..............................................................................................94 Variations with temperature..............................................................................95 4

The reaction rate....................................................................................................96 The turnover frequency.........................................................................................96 Lifetime of intermediates......................................................................................96 Activation enthalpy...............................................................................................97 Reaction orders.....................................................................................................98 Why does it work ?....................................................................................................99 Experimental methods..................................................................................................102 Titration of active sites.............................................................................................102 Rate measurements..................................................................................................103 Ideal reactors................................................................................................................105 Rate.........................................................................................................................105 The tank reactor.......................................................................................................105 The semibatch reactor..............................................................................................106 The plugflow reactor...........................................................................................108 The isothermal plugflow reactor.....................................................................108 Fit of kinetic data.........................................................................................................110 The simple approach................................................................................................110 A more general method............................................................................................112 Kinetic measurements.........................................................................................112 Thermodynamics.................................................................................................113 Choice of kinetic expression................................................................................114 Fitting the parameters .........................................................................................114 Checking the fit...................................................................................................115 Bibliography............................................................................................................116

1 Introduction
1.1 Definition of catalysis
A catalyst was defined by J. J. Berzelius in 1836 as a compound, which increases the rate of a chemical reaction, but which is not consumed by the reaction. This definition allows for the possibility that small amounts of the catalyst are lost in the reaction or that the catalytic activity is slowly lost. However, the catalyst affects only the rate of the reaction, it changes neither the thermodynamics of the reaction nor the equilibrium composition. Catalysis is of crucial importance for the chemical industry, the number of catalysts applied in industry is very large and catalysts come in many different forms, from heterogeneous catalysts in the form of porous solids over homogeneous catalysts dissolved in the liquid reaction mixture to biological catalysts in the form of enzymes.

1.2 Catalysis and process design


The thermodynamics frequently limits the concentration of a desired product. As the catalyst does not affect the thermodynamics of the reaction, it is futile to search for a catalyst to improve the situation. Instead the reaction conditions (temperature, pressure and reactant composition) must be optimized to maximize the equilibrium concentration of the desired product. Once suitable reaction conditions have been identified, the reaction rate is found to be too low, frequently by orders of magnitude. And the search for a suitable catalyst begins.

1.3 Catalysis and kinetics


The study of the kinetics of heterogeneous catalyzed reactions consists of at least three rather different aspects. Kinetics studies for design purposes. In this field, results of experimental studies are summarized in the form of an empirical kinetic expression. Empirical kinetic expressions are useful for design of chemical reactors, quality control in catalyst production, comparison of different brands of catalysts, studies of deactivation and of

poisoning of catalysts. Kinetics studies of mechanistic details. If a reasonable and not too detailed reaction mechanism is available, an experimental kinetic study may be used to determine details in the mechanism. Mechanistic considerations may be very valuable as a guidance for kinetic studies. Kinetics as a consequence of a reaction mechanism. The deduction of the kinetics from a proposed reaction mechanism generally consists in a reasonably straightforward transformation, where all the mechanistic details are eliminated until only the net gasphase reaction and its rate remains. This approach may be used to investigate if a proposed mechanism consistent, what the reaction rate is and if it is consistent with available experimental data.

For the three aspects of the study of kinetics, the optimal experimental and theoretical approach is quite different.

1.4 The basis for catalysis


The modern basis for the understanding of catalysis is spectroscopy of catalysts and catalyst models. kinetic data for catalytic reactions quantumchemical calculations for reactants, intermediates and products. calculation of the thermodynamics of reactants, intermediates and products from measured spectra and quantumchemical calculations. microkinetic modelling. Modern approaches to the study of reaction mechanisms consists of two approaches, experiments on well defined systems and detailed calculations for individual molecules and intermediates. The studies of well defined systems consists of spectroscopic studies of individual molecules and measurements of the rate of catalytic reactions on single crystal surfaces. as well as structure and reactivity of welldefined catalyst models. The computations consist of electron structure calculations including calculations for transition state as well as large Monte Carlo simulations.

1.5 Status of the study of catalysis


One of the most fascinating aspects of heterogeneous catalysis is that it is largely an empirical science. The application of catalysis has been a necessity for the chemical 7

industry for at least 150 years, while the experimental techniques for investigation of catalysis at the atomic level did not become routine until less than 25 years ago and the computational techniques are even younger and have hardly become routine yet. For this reason vast amounts of emipirical knowledge exists and awaits systematic investigation.

1.5.1 Chalenges
The challenges in the deduction of reaction mechanisms from spectroscopic studies are The pressure. Spectroscopic studies of molecules adsorbed on single crystal surfaces are made in ultrahigh vacuum and computations are made in the limit of zero pressure. The pressure must be extrapolated by at least 12 orders of magnitude. The temperature. Computations are made at zero temperature and a proper thermodynamics must be constructed. The structure. The catalyst consists of small particles stabilized by a structural promoter. This challenge may be overcome by studies of suitable catalyst models The conversion. The changes in gas phase concentrations which may be reached in a single crystal reactor is generally very low. This dictates that measurements are performed in the limit of zero product concentration. In this limit the kinetics may be entirely different from the kinetics at higher conversions.

1.6 Caveats
Before we proceed, it may be useful to list some of the key problems in the study of the kinetics of catalytic reactions. We cannot deduce the kinetics from the net reaction. For the reaction aA + bB cC + dD the kinetics is in general not p A a p B b k pC c p D d r k o p po K p o po although this kinetics predict the correct equilibrium pC c p D d

po pA

po pB

po po where po is the reference pressure.

In the absence of solid evidence it is dangerous argue by analogy. As an example, the reaction H2 + I2 2HI has a simple mechanism and a reaction rate of the form 2 pH pI k p HI r k o o K po p p while the reaction H2 + Br2 2HBr proceeds by a chain mechanism and has a complicated kinetics.
2 2

A reaction with a simple kinetics does not necessarily have a simple mechanism. As an example, the reaction 2N2O5 4NO2 + O2 has a simple kinetics of the form r = kpN2O5 but has a rather complex mechanism 1 2 3 N2O5 NO3 + NO2 NO3 + NO2 NO2 + O2 + NO NO + NO3 2NO2

A simple mechanism such as 1 2 3 A+* A* B2+2* 2B* *+B* AB+2*

may have a very complex kinetics. Very different reaction mechanisms may predict the same overall reaction rate. Even if we have reliable data for the overall reaction rate over a large range of reaction conditions we may be unable to distinguish between two different reaction mechanisms. Many mechanistic details cannot be deduced from an experimental determination of the form of the kinetic expression. As an example, the TemkinPyzhev rate expression for ammonia synthesis reproduces the experimentally observed kinetics quite well. However, this rate expression was originally derived from a proposed mechanism which had both the wrong key intermediates and the wrong ratelimiting step.

2 Reaction mechanism
A net reaction such as A2 + 2B 2AB often consists of a number of steps. Shortlived intermediates may be formed by some steps and consumed in other steps, e.g A2 + B A2B A2B + B 2AB Evidently, we can always subdivide the steps further and introduce hypothetical intermediates, e.g A2 + B A2B A2B + B A2B2 A2B2 2AB This leads to the introduction of the concept of an elementary step. A step in a reaction mechanism is elementary if it is the the most detailed, sensible description of the step. A step, which consists of a sequence of two or more elementary steps is a composite step. The question if a step in a reaction is an elementary step obviously depends on how detailed the available information is. The reaction mechanism deduced from a few, crude measurements of the reaction rate may consist of a small number of elementary steps. If we then decide to investigate the reaction through quantum chemical calculations, we will most likely find that many of these steps are in fact composite. The key features of a mechanistic kinetic model is that it is reasonable, consistent with known data and amenable to analysis. The description of a net reaction as a sequence of elementary steps is the mechanism for the reaction.

2.1 Complications
There are a number of features a reaction mechanism may have, which greatly complicates the situation.

2.1.1 Kinetic equations are nonlinear


For mechanisms where all steps consist of unimolecular reaction steps, the kinetics of the reaction is available analytically for arbitrarily large mechanisms. However, kinetic 10

expressions for elementary steps are not necessarily first order in the concentration of reactants. In a mechanism consisting of several steps, steps may even have different same order.

2.1.2 Inerts
An inert adsorbate does not have a well defined chemical potential and if inert surface species are present, the model is not soluble without additional assumptions on the behavior of the inert. Adsorbed inerts with constant coverage are better described by an adjustment of the number of adsorption sites. Adsorbed inerts with variable coverage are better described as reactants.

2.1.3 Nonconsecutive steps


The reaction mechanism does not necessarily consist of a sequence consecutive steps. Apart from the trivial case where consecutive steps are written in random order, some more interesting possibilities are One or more steps have been written "backwards" E.g step 2 in the mechanism: 1 A_2 + * A_2* 2 2A* A_2* 3 B + * B* 4 A* + B* AB + * The steps may not appear to be consequtive The mechanism has been written such that all steps except one has the form ni2A1 + ni2A2 + ... + * = mi1B1 + mi2B2 + ... C* where A1, A2, .. B1, B2 .. are all gases and C* is an adsorbed molecule. Parallel steps Parallel steps convert the same reactants into the same products through different routes.

11

2.1.4 Elusive intermediates


In a reaction mechanism, shortlived intermediates may be formed by some steps and consumed by other steps. The mechanism may contain intermediates, which have not been observed experimentally. The introduction of an hypothetical intermediate in the mechanism is in many cases a necessity to link observed the intermediates formed from the reactants with the observed intermediates formed form the products. If the calculated concentration of the hypothetical intermediate is too small and the lifetime too short to allow the experimental observation, the introduction of the hypothetical intermediate is of no consequence for the agreement between the model and experimental results. The introduction of hypothetical intermediates in excess of the absolutely necessasary is not sensible.

2.1.5 Undetectable steps


A mechanism may contain steps that are irrelevant as they are of no consequence for the consistency of the mechanism and of no consequence for the reaction rate. Some examples of undetectable steps are A slow reaction step that is short circuited by a sequence of equilibrium steps. The net rate of the slow step is then zero. A fast step in series with a slow step A slow step in parallel to a fast step.

2.1.6 Dead ends


One or more steps may form a dead end in the form of an intermediate formed through an elementary reaction and consumed exclusively by the reverse of this step. Although the deadend will not contribute to the overall reaction rate, the step may affect the kinetics if the intermediate is strongly adsorbed on the surface. The poisonous effect of H2O in ammonia synthesis is an example.

2.1.7 Linear dependence between reaction steps


A reaction mechanism may have linearly dependent reaction steps. This may happen for two reasons. First, the same reaction step occurs more than once with different kinetic parameters, e.g A* + B* AB* + * with A=109 and E#=6 kJ/mol. 12

A* + B* AB* + * with A=1013 and E#=52 kJ/mol where the first equation describes a low barrier, low temperature channel and the second describes a high barrier, high temperature channel. Second, steps may be combined, such as the following steps that occur in the watergas shift reaction 1 2 3 4 H2O + * H2O* + * OH* + * 2OH* H2O* OH* + H* H* + O* H2O* + O*

Although linear dependence among the reaction steps complicates the analysis of a mechanism, the mechanism is physically meaningful provided the mechanism is stochiometrically and thermodynamically consistent.

2.2 Consistency
For a proposed reaction mechanism, there must be a sequence of steps that leads from reactants to products. This requirement is implicit in the definition of a reaction mechanism. All intermediates occur as reactant for at least one step and as product for at least one step. This requirement is essentially the definition of an intermediate. Further all reaction step must have a thermodynamics and all slow steps must have a rate. If an reactant, intermediate or product participate in two or more steps, the stoichiometry of molecule must be independent of the way the intermediate is formed. This is the principle of stoichiometric consistency. If two or more different sequences of steps lead from reactants to products, these sequences must describe the same gas phase thermodynamics. This is the principle of thermodynamic consistency.

2.3 Guidelines
There are some rules of thumb which can guide the formulation of reaction mechanisms.

13

The reaction enthalpy of each step is moderate. If the reaction enthalpy is large and positive, the activation energy in the forward direction must be large and the reaction rate will be negligible. Either the step in question is in reality an irrelevant byway in the mechanism or the overall reaction will have negligible rate. If the reaction enthalpy is large and negative, the activation energy in the backward direction must be large and the reaction rate will be negligible. Either the step in question is in reality an irrelevant byway in the mechanism or the overall reaction may have problems establishing equilibrium. For any step, the number of reactants and product molecules If the number of molecules is large, the activation entropy will be large and we have the same complications as for large reaction enthalpies. Actually the problem is a little worse, because it will go away at high temperatures. For any step the number of broken or formed bonds are small.

14

15

3 Kinetics
The rate of chemical reactions can be described at two levels: dynamics and kinetics. Dynamics is the description of the rate of transformation for individual molecules. The molecule has a welldefined energy, it may even start in a welldefined quantum state. There is no temperature. Temperature is a property of a large number of molecules, not individual molecules. The detailed microscopic description of a chemical reaction in terms of the motion of the individual atoms taking part in the event is known as the reaction dynamics. The study of reaction dynamics at surfaces is progressing rapidly these years, to a large extent because more and more results from detailed molecular beam scattering experiments are becoming available. Kinetics is the description of the rate of reaction for a large number of molecules. The molecules have a temperature, although the temperature may change in the course of the reaction. The energy is welldefined, but the energy is a statistical average.

3.1 Reversibility
For a reaction, e.g. A2+ 2B 2AB the reverse reaction 2AB A2 + 2B will proceed through the same mechanism, although the sequence and the direction of each of the elementary reactions is reversed This is known as the principle of microscopic reversibility The cause of this principle is that in the kinetic description, we explicitly assume that the intermediates equilibrate at the reaction temperature. This implies that the intermediates have no memory how they are formed. A formed by dissociation of AB is identical to A formed by dissociation of A2. As the reaction proceeds through the same steps in the forward and in the backward reaction, while the rate of the individual steps may differ by many orders of magnitude it is convenient to consider two classes of steps. Fast steps have a high rate in both forward and backward direction, while slow steps have a low rate in the forward direction, in the backward direction or both. 16

3.2 The reaction rate

For net a reaction, say, A2 + 2B 2AB with mechanism A2 2A A + B AB each step in mechanism proceeds with some rate r+ in the forward direction and some rate r in the backward direction. The net rate of the step is obviously r= r+ r The rate is a function of temperature, pressure and concentrations r(T,p,xi). The rate generally decreases with time as the composition approaches quilibrium. While the forward reaction rate for the net reaction may depend on the concentration of both reactants and products, the forward rate of each elementary step can depend only on the concentration of reactants for this step. This leads to the expression of the rate of as the number of times the reaction proceeds per second, the turnover frequency. The rate is thus a rate for the reaction, not the rate for the reactants or for the products. If ni is the number of moles produced of product number i and i is the stochiometric coefficient for product number i, the turnover frequency is r

ni
i

This turnover frequency is obviously the same for all products. The reactants have negative stochiometric coefficients and are "produced" with negative rate, so the turnover frequency is actually the same for all reactants and products.

3.3 Rate laws


In the simplest case the reaction rates are proportional to the coverages e.g for the mechanism A2 + 2* 2A* B + * B* 17

A* + B* AB + 2* the rates are p pB po


B*

A*

The assumption that rates are proportional to coverages eliminates some, but not all, hysteresis phenomena.

3.4 Forward and backward rate


For each step the rate, r = r+ r, is obviously the difference between a forward rate, r+, and a backward rate, r. For each of the gases we have an formation rate. For many applications, it is sufficient to determine the rate of formation for each of the gases. However, if we want to determine the kinetic parameters for the net reaction, the form of the rate expression must be determined.

3.5 Stoichiometric matrix


For a systematic treatment of mechanisms, we need a suitable mathematical device. On possibility is to use a stoichiometric matrix to represent the mechanisms in symbolic form. We write the reaction mechanism using a the stoichiometric matrix, . For a mechanism consisting of G gases, S adsorbates including free sites, and R reactions, is a R by G+S matrix. We use the convention than rc < 0 if c is a reactant of step r, rc > 0 is a product of step r and rc = 0 if c does not participate in step r. We will frequently need products or sums running over subsets of the molecules. We will use the convention that the molecules are enumerated with gases number 1,...,G, free sites is number G+1 and adsorbates are number G+2,...,G+S. 18

r 3 k +3

k +3

p AB

r 2 k +2

r 1 k +1

p A2

2 *0

k 1 k 2

2 A*

B*

2 *

3.5.0.1 Example
As an example we will consider the mechanism for ammonia synthesis. We include Ar in the gas phase to illustrate the effect of inerts. This mechanism is rich enough to illustrate most of the features discussed below. For this mechanism R=7 (steps 1 to 7), G=4 (N2, H2, NH3, and Ar), S=7 (, N2*, N*, NH*, NH2, NH3, and H*). The stoichiometric matrix is c=1 c= 2 c=3 r=1 r=2 r=3 r=4 r=5 r=6 r=7 1 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 1 0

c=4 0 0 0 0 0 0 0

c= 5 1 1 1 1 1 1 2

c=6 1 1 0 0 0 0 0

c= 7 0 2 1 0 0 0 0

c=8 0 0 1 1 0 0 0

c=9 0 0 0 1 1 0 0

c=10 c=11 0 0 0 0 1 1 0 0 0 1 1 1 0 2

3.5.0.2 Properties
rc has a number of interesting properties: Surface sites are conserved R r 0 for c=G+1,...,G+S All elements in the stoichiometric matrix are integers. The use of noninteger stoichiometric coefficients is unnecessary and greatly complicates the treatment. Gas inerts have rc =0 for r=1,...,R.

r 1

rc

3.5.0.3 Rate
The rate is calculated as a turnover frequency i.e as a number of molecules produced1 per site per second.

1 The distinction between reactants and products depends on which gases are present in the initial mixture. From now on, we use the word product for all gases when information on the initial mixture is unknown or irrelevant 19

The reaction rate for step r is r r r +r r r


rc
 

c 1,

For each of the gases we have an formation rate


R

rc r r 1 rc r this rate is evidently negative for the reactants.

3.5.0.4 Equilibrium equation


As a consequence of this choice of sign for rc the equilibrium constants are G G S pc for r=1, ..., R Kr o c 1 c G 1 c p
 

rc

rc

3.6 Rate constant


At equilibrium the net rate is by definition zero. If we compare the rate equation for an elementary step, e.g the rate equation pA r 1 k +1 o 2 k 1 2 * A* p and the equilibrium equation pA K1 o 2 2 * A* p for the step A2 + 2* 2A* we find that the forward and backward rate constants are related k+ k K Usually both k+ and k have Arrhenius form H# k A exp RT at least for small or moderate variations of T. In this equation Ais the preexponential factor while H# is the activation energy

20

rc

Kr

c 1,

c G 1,

rc

r r

kr

rc

G S

p pc

c G 1,

rc

rc

r +r k r

pc

rc

G S

rc

rc

H # RT 2

d ln k dT

If k+ and k have Arrhenius form, the equilibrium constant, K, k+ K k will have Arrhenius form as will any product or quotient of rate and equilibrium constants. The activation energies and the energy of reaction will be related by # # H H H+

3.7 Rate limiting steps


It is often the situation that most reaction steps in a mechanism are fast, while a single step is much slower than the other. In this situation, the slow step is called the rate limiting step (RLS) or the rate controlling step as it determines the rate of the overall reaction. Let us return to the question why the slowest step controls the overall rate of the reaction. As an example consider the net reaction A2 + 2B 2AB with mechanism A2 + B A2B A2B + B 2AB and assume that the first step is much slower than the second. In this case we will have that r1+, r1 << r2+, r2 However, A2B is a common intermediate in these two steps. If A2B is an intermediate, rather than a product, the concentration of A2B must be essentially constant, i.e the net rate of formation of A2B must be essentially zero r1+ r1 r2+ + r2 = 0 A rearrangement yields r1+ r1 = r2+ r2 i.e the net rate of the fast step is the same as the net rate of the slow step. To summarize, the first step converts A2 + B. to A2B at some rate and dissociates a little A2B back into A2 + B, while the second step rapidly converts A2 + B into 2AB and 2AB back into A2 + B. Still the massbalance and the fact that A2B is an intermediate results in the net reaction for both steps. The following figure provides an illustation of this principle:

21

r1+ r1 r2+ r2 r

3.8 Most abundant reaction intermediate


For many reaction mechanisms one of the reaction steps is much slower than all other steps and this slow step determines the net rate of the reaction. There is an analogous situation with respect to the concentration of reaction intermediates. In many reaction mechanisms there are several intermediates, but frequently the concentration of one of the intermediates is much larger than the concentration of all other intermediates. This intermediate is then called the most abundant reaction intermediate (MARI). It may be tempting to speculate that the most abundant reaction intermediate should be either the product of the first reaction step or one of the reactants for the rate limiting step. This is not the case.

3.9 Reaction order


The forward rate typically has the form

22

r+ k

where A is the reaction order for A. The reaction orders for the forward and backward reactions are related through the equilibrium equation. If we write the backward rate as r k

pA

pB

po po and the equilibrium equation as K we have A = + = + and


A B

pA po

pB po

po po

po
A

po

The thermodynamics of the overall reaction thus provides a connection between the reaction orders for the forward and backward reaction orders. There is no simple connection between the reaction orders for the overall reaction and the stochiometry of the rate limiting step. For the same reason, the experimental reaction orders for the overall reaction provides essentially no information on the reaction mechanism.

3.10 The activation energy


The ratelimiting step can usually be described as an energy barrier the system must cross.

pA

pB

1 pA K po

pB po

23

r k

...

po
B

...

pA

pB

k pA K po

pB

pA

pB

...

...

...

...

The rate constant then has the temperature dependence: k A exp

E kBT

where A is the preexponential factor and E is the activation energy.

3.10.1 Activation energy vs reaction energy


The rate constants, k+ and k and the equilibrium constant, K, are thus related by K

k+ k

and the activation energies, E+# and E#, and the energy of reaction, E, are related by E = E+# E#

3.10.2 Graphical determination

24

The activation energy can be determined graphically from experimental determination of the rate constant through an Arrhenius plot.

3.10.3 Analytical determination


The activation energy can be determined analytically from an expression for the rate: d ln r + # E+ k BT 2 dT

3.10.4 Activation energy for composite reactions


While the Arrhenius form of the rate constant for an elementary step is easy to justify, the situation is more complicated for composite steps or even complete reaction mechanisms. The complexity occurs because even if all the rate constants for the elementary steps have Arrhenius form, the overall rate constant will in general not have Arrhenius form. We can force the net rate to have Arrhenius form only if we allow the apparent activation energy to depend on the reaction conditions. Only in the case that the activation energy is more or less independent of the reaction conditions can we determine the activation energy for the mechanism and only in this case can the activation energy be interpreted as an energy barrier. Fortunately, for most reactions of practical interest, the dependence of the activation energy on the reaction conditions is weak except at extreme reaction conditions. The question is then how to define the activation energy for a composite reaction. The obvious answer is that as the activation energy is determined experimentally through an Arrhenius plot, we should use the Arrhenius plot to define the activation energy for 25

composite reactions.

3.10.5 Compensation effect


The activation energy for a reaction is sometimes measured under different reaction conditions. An example might be a measurement using a very active catalyst at a moderate temperature and a measurement using a less active catalysts at higher temperatures. The increase in temperature partially compensates for the lower activity. We might expect to find the same activation energy, however, one usually finds different activation energies and different preexponential factors: A large value for the activation energy is correlated with a large prefactor and all lines in the Arrhenius plot intersect in a single point, the isokinetic point. The correlation between activation energy and preexponential factor is known as the compensation effect.

There are a number of possible explanations for the compensation effect: The most interesting case is when the compensation effect is caused by differences in the transition state. In this case the compensation effect is a real physical effect. A trivial explanation is when each of the measurements has been made over a narrow range of temperatures. The slopes of the lines are then very uncertain and the observed compensation effect may be statistical insignificant. If the kinetic expression used to convert the measured rate into a rate constant is not quite right, the conversion will be more or less in error. If error depends on temperature, the Arrehenius plot will display a fictitious compensation effect.

3.10.6 Two reactions in series


One of the simplest cases of two consequtive reation steps is the MichaelisMenten kinetics: A + * A* 26

A* B Assuming that the second step is irreversible and using the steadystate approximation, we can determine the coverage

K1

and the reaction rate

K1

From the rate we can calculate the activation enthalpy for the reaction k1

dT

K1

The activation enthalpy has two limits. When the first step is ratelimiting, k1 <<k2 and H# = H1# When the second step is ratelimiting, k1 >> k2, and H# = H2#+ H(1) Going back through the calculation, we see that the first term comes from k2 and is caused by the rate limiting step speeding up when the temperature is increased. The second term comes from the differentiation of . When the temperature is increased, the rate will change partially due to changes in . The second term describes this effect. For two consequtive reactions the reaction enthalpy for the overall reaction is the sum of the reaction enthalpies for each of the steps
  

H


H1

H2

For the activation enthalpy, there is no similar sumrule. The activation enthalpy for the overall reaction is not the sum of the activation enthalpies for each step, actually the activation enthalpy for the overall reaction is only slightly larger than the activation enthalpy for the slowest step.

27

kBT2

d ln r +

H1 H 2

K1 k1 p
o

k1

k2

r k2

k1 k 2 k1 p o p k1 k2

k1

p 1 po

k1

k2

p # H1 po

k1

H1

k2 H2

3.11 Rate and conversion


When we measure a reaction rate, we are usually not measuring the rate itself, instead we measure a conversion, i.e a change in concentration from the inlet to the outlet of a reactor or the change in concentration over some time interval. Generally the reaction rate depends on the concentration in a nonlinear fashion. The change in concentration divided by the contact time or some other rubbish is not a measure of the reaction rate. We need to know or to assume the dependence of the rate on the concentration in order to calculate the reaction rate from the observed change in concentration. In general this is difficult for realistic models:

In the next section we will look at the pseudo1order kinetics, which is just about the only realistic kinetic model that can be treated by hand. We will return to the treatment of realistic models later.

3.12 Pseudofirst order kinetics


The pseudo1order reaction has the kinetics: d x xe k x xe dt where xe is the equilibrium concentration. If the concentration is x1 at t=0 and x2 at t=t we can solve the differential equation x dx kt x x xe The solution is

28

For a catalyst bed with mass m and flow F, the contact time is proportional to m t F The pseudo1order reaction cannot easily be generalized to a pseudonthorder reaction. d x xe k x xe n dt If n is even the rate is negative for all concentrations and the composition after infinite time is zero, independent of the value of xe.

3.13 Analogy between reactors


There is an important equivalence between position in a reactor and the contact time. The upper instrument is placed at the exit of the catalyst bed. The middle instruments moves along the catalyst bed with the same linear velocity as the gas flowing in the catalyst bed. When the yellow instrument reaches the end of the reactor, the green and the yellow instruments show the same concentrations illustrating the equivalence of position and contact time.

The lower instruments monitors a closed container sealed off at time zero. The blue and the yellow instruments show the same concentrations from time zero until the yellow instruments reaches the end of the catalyst bed, illustrating the equivalence between flow and batch reactors. When we calculate the conversion for a given reaction, the important parameters are the time and the reaction rate. It does not matter if we let the reaction run in a closed container or if the gas flows while it reacts.

kt x1 x e and after rearrangement we get x 2 x e x 1 x e exp




ln

x2 xe

kt

29

30

4 Solving the reaction scheme


4.1 The full solution
In this case we dont make any approximations in the reaction scheme. We consider a Langmuir Hinshelwood mechanism with stoichiometric matrix rc. The kinetic model for this mechanism consists of the forward and backward rates.
G


rc

and the balance over sites


G S c G 1


The solution to this system of equations evidently gives a correct description of both equilibrium, steady state and transient behavior. If the reaction mechanism contains more than one or at most two steps, the full solution becomes very complicated and we will have to solve for the rates and coverages by numerical methods. Although the full solution contains the steady state behavior as a special case, it is not generally suitable for studies of the steady state as the transients may make the simulation of the steady state a numerical nightmare.

4.2 Steady state approximation


In the steady state approximation the net rate of formation for all intermediates is set to zero. In the steady state approximation the net rate of formation for all intermediates is explicitly set to zero, 31

rc

r


Kr

c 1,

c G 1,

rc

rc

kr

pc

G S

rc

r


c 1,

c G 1,

rc

kr

rc

pc

rc

G S

rc

rc

Assuming that the net rate of formation for intermediates is zero does not imply that the coverage by the intermediates is small. For a mechanism with S1 intermediates in addition to free sites the steady state equations is a system of S equations, at most (S1) are linearily independent. The steady state approximation eliminates transient behavior in the kinetics. However, it is only the transient behavior of the rates and coverages that has been eliminated. The expression for the rate obtained through the steady state approximation is perfectly suitable for the simulation of e.g the conversion through a catalyst bed or most aspects of the transient behavior of a reactor.

4.3 The quasiequilibrium approximation


If all steps except one are fast, we can use the quasiequilibrium approximation: For the fast steps we use the corresponding equilibrium equations instead of the kinetic equations.

... ...
rc

This approximation will in most cases provide a very significant simplification in particular for large reaction mechanisms. In the quasiequilibrium approximation the transient behavior is eliminated. Further, the description of changes in ratelimiting step has been lost.

c 1

c G 1

32

KR

pc

Rc

G S
Rc

rc

rc

rc

c 1,

p ... ...

c G 1,

rc

Kr

c 1,

c G 1,

rc

r r kr

pc

G S

rc

kr

pc

K1

c 1

c G 1

r 1

rc

rr

0 for c = G+1,...,G+S

pc

1c

G S
1c

rc

G S rc

rc

4.4 Irreversible step approximation


In the irreversible step approximation, we neglect the forward or backward rate for one of the steps. For small mechanisms the irreversible step approximation may be used alone, for larger mechanisms it is usually combined with the quasi equilibrium approximation

... ...


c 1

c G 1

This approximation is very crude as we have lost the description of the approach to equilibrium. The model will now continue to convert all reactants right across the equilibrium, often ending in a spectacular numerical instability when the concentration of a reactant becomes negative. Although this approximation is useless for the quantitative modeling of reactions, it has two important uses in the analysis of reaction mechanisms: If we want to determine the limiting behavior of a kinetic model very far from equilibrium, the irreversible step approximation is the appropriate limit. If we have difficulties making sense of a complicated reaction mechanism, the irreversible step approximation may provide a simplification, which allows us to understand the mechanism well enough to choose a better approximation.

4.5 MARI approximation


The mostabundant reaction intermediate (MARI) approximation is a further development of the quasiequilibrium approximation. Often one of the intermediates is much more abundant than all other intermediates, the coverages by the lessabundant intermediates may then be neglected in the balance over sites. If we assume that A* is the most stable intermediate and the balance over sites becomes * + A* = 1 33

KR

pc

Rc

G S
Rc

rc

c 1,

c G 1,

rc

r r kr

rc

pc

p ... ...

c 1

c G 1

rc

K1

pc

1c

G S
1c

G S

rc

In the MARI approximation, we have made one of the intermediates MARI at all reaction conditions. This has two consequences 1. We have lost the description of a change in MARI. 2. We have explicitly assumed that all intermediates except the MARI are much less abundant than free sites. In particular the latter consequence is troublesome as an estimate of the validity of this aspect of the MARI approximation amounts to solving the problem without using the MARI approximation. In other words, if the validity of the MARI approximation can be verified, only if it is superflous. The MARI approximation can be used for quantitative modeling, if we have verified that it is valid at the reaction conditions we are considering. The MARI approximation is very much used for the analysis of reaction mechanisms, both when we have difficulties in formulating a kinetic model for a complicated reaction mechanism and when we want to derive a limiting form of a kinetic model.

34

5 Adsorption
5.1 Introduction
When a gas reacts with a solid, the most common situation is that a new compound is formed and that the crystal structure of the solid is destroyed, Fe reacts with H2S to form FeS, Al reacts with O2 to form Al2O3 etc. However, there are reactions where the crystal lattice of the solid is only slightly odified: absorption, where the gas enters the bulk of the solid, and adsorption, where the gas rmains at the surface. There are two main classes of adsorption: physisorption and chemisorption.

Gas solid reactions

Generic

Absorption

Adsorption

Chemisorption

Physisorption

The generic case is of course the case, where the crystal structure of the solid is destroyed.

5.2 The LennardJones picture


In adsorption the gas molecules form some kind of bond with the surface, the strength of this bond depends on both the surface and on the gas molecule. The nature of the bond varies from very weak van der Waals interactions to very strong chemical bonds.

35

While the binding energy is very variable, the qualitative form of the potential energy surface was deduced by LennardJones in 1932. The abscissa is the reaction coordinate and may loosely be interpreted as the height over the surface. As the molecule approaches the surface, it first feels a weak, longranged attraction. Near the surface there is an energy minimum and closer to the surface there is repulsion in the form of a steep increase in energy.

In reality, the potential energy surface is multidimensional and depends on the height, the position and orientation of the molecule. The LennardJones picture is obviously an oversimplification, but is still an extremely useful way of analyzing a reaction, and it has been shown in a number of cases to work remarkably well.

5.3 The potential energy surface


If we consider the energy as a function of the height, z, above the surface and the interatomic distance, r, in the molecule, we can represent the energy as a contour plot. The existence of adsorption sites causes problems for the LennardJones picture, the interaction energy depends on many coordinates, not just the height above the surface. The transition state is at the arrow.

36

A vibrating molecule might follow the black curve. Only the start of this curve is shown. The continuation may be either simple or extremely complicated.

The outer part of the potential energy surface describes the motion of the molecule far from the surface. The motion along z is a smooth translation, while the motion along r is just the vibration of the free molecule.

The inner part of the potential energy surface describes the motion of the chemisorbed molecule. The motion along z is now a stiff vibration, while the motion along r has become softer due to the weakening of the intermolecular bond.

5.4 Motion of adsorbed molecules


For most adsorbates the motion along the surface is at reasonably low temperatures better approximated by 2 vibrations around an equilibrium position. 37

For a threedimensional gas 3 translational degrees of freedom exist. Upon adsorption these 3 dimensional degrees of freedom are transformed into one degree of freedom orthogonal to the surface and 2 degrees of freedom parallel to the surface. The degrees of freedom in the adsorbed state, which originates from the translation of the gasmolecule will be referred to as parallel and orthogonal frustrated translation in the following. The degree of freedom orthogonal to the surface is of vibrational nature. If the electronic structure of the surface does not affect the movement of the molecule, the 2 degrees of freedom parallel to the surface will be translational and the adsorbed molecule will behave as a two dimensional gas. However, as a considerable variation in binding energy exists over the unitcell. Assuming that the potential varies smoothly over the unit cell we can approximate it by 1 V x E 0 1 sin kx 2 The free translation has a very high entropy, while the entropy of a vibration is moderate. For this reason the adsorption entropy is usually negative. At a given pressure the equilibrium between gas and adsorbate will shift towards desorption when the temperature is increased. For surface species one or more of the rotational degrees of freedom may be restricted. Such frustrated rotations are better described as vibrations.


5.4.0.1 Example A2*


In the gasphase the A2 molecule has three translational, one vibrational and a rotational degree of freedom. In the adsorbed state these degrees of freedom are transformed into one rotational and 5 vibrational degrees of freedom.

38

A rotation of the molecule around the vertical axis. This is essentially one of the rotations in the gas phase.

A vibration, which is the remains of the other rotation in the gas phase.

A vibration parallel to the surface. This is the remains of the free translation along x

A vibration parallel to the surface. This is the remains of the free translation along y.

A vibration orthogonal to the surface. This is the remains of the free translation along z.

An intramolecular vibration. This is essentially the intramolecular vibration in the free molecule.

6 Catalysis
As a catalyst takes part in the chemical reaction, but is not consumed by the reaction, the catalyst must be a reactant in one of the first steps in the mechanism and a product in one of the last steps. The reaction proceeds in a cyclic fashion, where the catalyst, or more correctly the catalytic sites, are regenerated and used again and again.

39

The activity of a catalyst can be written as a product of two factors, the number of active sites and the turnover frequency. The turnover frequency is the subject of this section, we will return to the number of active sites later.

6.1 LangmuirHinshelwood mechanism


When we consider a catalytic reaction, we may imagine that the reaction mechanism consist of many different steps. Each of these steps may be of different types, we may imagine that adsorbed species react with each other, that surface species may migrate into the bulk, that reactive radicals desorp and then react in the gasphase etc. Fortunately, real reaction mechanisms appear to be rather simple. The LangmuirHinshelwood mechanisms form an important class of reactions. These mechanisms consist of the following types of steps: Adsorption from the gasphase Desorption to the gasphase Dissociation of molecules at the surface Reactions between adsorbed molecules

The questions if the reaction has a LangmuirHinshelwood mechanism and what is the precise nature of the reaction steps cannot be solved without either experimental or computational studies.

6.1.1 Example
The reaction A2 + 2B 2AB may have the following mechanism A2 + * A2* A2* + * 2A* B + * B* A* + B* AB* + * AB* AB + *

40

6.1.2 Example
If AB* is formed through the following steps A2* + B* A2B* + * A2B* + B* 2AB* the mechanism remains a LangmuirHinshelwood mechanism.

6.1.3 Example
If we have evidence that the reaction appears to proceed as A2 + 2* 2A* B + * B* A* + B* AB* + * AB* AB + * then we cannot immediately conclude that the mechanism is not a Langmuir Hinshelwood mechanism. Actually, to rule out that this mechanism is not a Langmuir Hinshelwood mechanism, we need to rule out that the first step may be a composite reaction A2 + * A2* A2* + * 2A* with a vanishing small equilibrium concentration of A2.

6.2 EleyRideal mechanism


The EleyRideal mechanisms form an important class of reactions. These mechanisms consist of the following types of steps: Adsorption from the gasphase Desorption to the gasphase Dissociation of molecules at the surface Reactions between adsorbed molecules Reactions between gas and adsorbed molecules.

The last type of steps cannot occur in a LangmuirHinshelwood mechanism.

41

6.2.1 Example
The reaction A2 + 2B 2AB may have the following EleyRideal mechanism A2 + * A2* A2* + * 2A* A* + B AB + * where the last step is the direct reaction between the adsorbed molecule A* and the gas molecule B.

6.2.2 Example
However, without further evidence we cannot conclude that the above mechanism is an EleyRideal mechanism. The last step may be composite and consist of the following steps B + * B* A* + B* AB* + * AB* AB + * with a vanishing small equilibrium concentration of B*. The mechanism is then LangmuirHinshelwood and not EleyRideal.

6.3 EleyRideal or LangmuirHinshelwood


If we can vary the coverage of A* and monitor this variation or if we can vary the ratio between A2 and B we can discover if the mechanism is EleyRideal or LangmuirHinshelwood. The trick is that the step B + * B* requires a free site. If we measure the reaction rate as a function of the coverage by A*, the rate will initially increase for both mechanisms.

42

For the EleyRideal mechanism, the rate will increase with increasing coverage until the surface is completely covered by A*. However, for the LangmuirHinshelwood mechanism the rate will go through a maximum and end up at zero, when the surface is completely covered by A*. This happens because the step B + * B* cannot proceed when A* blocks all sites.

6.4 The principle of Sabatier


When different metals are used to catalyse the same reaction, it is generelly observed that the reaction rate can be correlated with the position of the metal in the periodic table:

This plot is, for obvious reasons, called a volcano curve and the principle that the points will fall on a smooth curve is called the principle of Sabatier It is generally found that the activation energy for dissociation of simple diatomic molecules decrease when going left from the noble metals in the periodic table. This can be described most simply in terms of an increased interaction between the antibonding adsorbate states and the metal dstates. During the adsorption of simple diatomic molecules the antibonding molecular orbitals are gradually filled. For CO, for instance, the antibonding 2 states are partly filled for the chemisorbed molecule and fills even more during the dissociation process, and a similar picture holds for the other simple gas molecules. The transfer of electrons to the antibonding molecular orbitals is in general most facile when the metal work function is small. Low work functions are found for the most open surfaces. The work function can also be lowered locally by adsorbed electropositive species like alkali atoms. During the electron transfer the antibonding molecular states must be close to the 43

Fermilevel of the metal, and when the metal has dstates around the Fermi level, as in the transition metals there will be a strong covalent interaction between the antibonding states and the metal dstates. Trends in dissociative energies and activation energies for dissociation as a function of the number of delectrons. The results are calculated in the NewnsAnderson model including the coupling between an adsorbate level $\epsilon_a$ and the metal dband. The interaction tends to stabilize the adsorbing molecule more for metals towards the left in the transition metal series which have fewer d electrons than ten, the largest effect being around the middle of the series where the number of d electrons is around five. We can use the model to begin understanding which metals are active and which are not. We know roughly how the stability of the intermediates and the barrier for N2 dissociation varies through the periodic system. If we include these variations in the kinetic model, we get the following variations in the ammonia activity: The calculated ammonia production for a fixed set of reaction conditions as a function of the number of delectrons. For the elements left of Fe, the kinetics is similar to the kinetics over Fe, but the rate is low due to the strong bonding of N*. For the elements right of Fe, the rate is low due to the low sticking coefficient for N2, the coverage by N* is low and the predicted kinetics is somewhat different from the kinetics over Fe.

Fe (and the other elements with approximately 7 delectrons) are the optimum choice not because the sticking probability is high or because there is much free surface, but because it is the best compromise between the two effects. These model calculations have treated all the transition metals as equivalent except for the dband occupancy. This is of cause a gross oversimplification and much more detailed calculations are needed for a detailed picture. The simple description does, 44

however, give a physical picture of the main trend. Going to the finer details the interaction energy does, for instance, depend on the dband width, even in the simple NewnsAnderson model. The main effect is that the narrower the band the stronger the interaction. This is an additional reason why, in the calculations described in the previous section, the open surfaces have lower activation energies than the more close packed ones. The surface atoms in an open surface have a lower metal coordination number and since the band width is roughly proportional to the square root of the coordination number, the band width is smaller.

6.5 Structure sensitivity


The reactivity of different facets of a metal may be different. It depends on both the metal and on the reaction, if the reaction rate will be different on different facets and if the difference is small or large. Reactions where the rate varies significantly from one facet to another are called structure sensitive. It is rather surprising that most reactions appear to be more or less structure insensitive. Some possible explanations are Sites consist of only a few atoms and the local geometry of the site is more or less the same independent of the structure of the facet. The surface reconstructs and the the pure surface has more or less the same structure independent of the orientation of the bulk orientation. The adsorbate layer reconstructs and has a structure which is more or less independent of the structure of the pure surface. The coverage under reaction conditions is high and differences in adsorbate adsorbate interactions balance the differences in adsorption energy.

6.6 Physisorption
In physisorption the bond is a van der Waals interaction and the adsorption energy is typically 510 kJ/mol. This is much weaker than a typical chemical bond and the chemical bonds in the adsorbing molecules remain intact. However, the van der Waals interactions between adsorbed molecules is not much different from the van der Waals interaction between the molecules and the surface. For this reason many layers of adsorbed molecules may be formed.

45

6.7 The BET isoterm


The BrunauerEmmettTeller (BET) isoterm follows from the following assumptions 1. The adsorption takes place on a lattice. 2. The first adsorbate layer is adsorbed on the solid surface, the second adsorbate layer is adsorbed on the first etc. Except, of course, for the first layer, a molecule can only be adsorbed on a given site in layer number n, if the same site is occupied in layer n1. 3. At the saturation pressure p0 the number of adsorbed layers is infinite. 4. The adsorption enthalpy is H1 for molecules in the first layer and HL for molecules in the following layers. The BET isoterm is


p N p0 p

1 n mC

C 1 p n MC p0

where n is the amount of gas adsorbed at pressure p nm is the amount of gas correponding to one monolayer. p0 is the saturation pressure, n is infinite at p=pm. C is a constant, C exp

H1 H L kBT

where H1 is the adsorption enthalpy for the first layer and HL is the adsorption enthalpy for the following layers.

6.8 Chemisorption
For chemisorption the adsorption energy is comparable to the energy of a chemical bond. The molecule may chemisorp intact or it may dissociate. The chemisorption energy is 3070 kJ/mol for molecules and 100400 kJ/mol for atoms. The number of sites is constant and the competition for the adsorption sites has important consequences for the macroscopic kinetics. This is the reason for treating the surface sites as if they were a reactant in the reaction equations. The competition for adsorption sites is very important for the kinetics of a heterogeneous catalytic reaction. For this reason sites, *, are included as a reactant in the kinetic model. As a site must be either free or occupied by one of the surface intermediates, there is a conservation law for the coverages 46

where X is the coverage by the intermediate X. In writing this equation we have implicitely defined X=1 to be saturation. With this convention, coverages may be interpreted as probabilities. An isoterm is the coverage, , considered as a function of temperature and pressurte, (T,p). In the next two sections we give a simple and intuitive derivation of the Langmuir isoterm before we present a much more general derivation.

6.8.1 The langmuir isoterm


An adsorption isoterm describes the coverage, function of temperature and pressure, (T,p). The simplest, useful adsorption isoterm is the Langmuir isoterm. This isoterm was first derived by Irving Langmuir in 1916 and is one of the oldest concepts in surface science and catalysis. The Langmuir isoterm occurs when The adsorbate forms a monolayer. There is zero or one adsorbed molecule at each site. All sites have the same adsorption energy. There is no interaction between sites.

6.8.1.1 Nondissociative adsorption


If we consider the reaction A + * A* and for a moment assume that the mechanism for the reaction is equal to the net reaction. The adsorption rate is proportional to the pressure, p, and to the coverage by free sites, 1, while the he desorption rate is proportional to the coverage by A*. The proportionality constants are actually the rate constants. At equilibrium the adsorption rate balances the desorption rate

47

p 1 kb po Solving this equation with respect to leads to the Langmuir isoterm p K o p p 1 K o p where kf K kb

6.8.1.2 Dissociative adsorption


If we consider the reaction A_2 + 2* 2A* and for a moment assume that the mechanism is identical to the net reaction, the adsorption rate will be proportional to the pressure, p, and to the probability, (1)2 of finding a vacant pair of sites. The desorption rate is proportional to the probability, 2, of finding an occupied pair of sites. At equilibrium the rate of adsorption equals the rate of desorption:

kf

p 1 po

kb

Solving this equation with respect to the coverage, , gives the Langmuir isoterm K

p po p po

6.8.1.3 Competitive chemisorption


If instead of n molecules of A* adsorbing on S sites, we have n1 molecules of A*, n2 molecules of B* etc we obtain the following equations

kf

48

po ... ...

6.9 Sticking
Sticking is a process where a gas molecule collides with a surface and ends in an adsorbed state. There are two interesting aspects of this process, the probability that the molecule is adsorbed and the state of the adsorbed molecule. The probability that an atom or molecule hitting the surface will adsorp is known as the sticking coefficient. If the sticking coefficient is $\sigma$, the adsorption rate is .

6.9.1 as a rate
In the following we will usually calculate turnover frequencies, i.e the number of reactions per site per second. The adsorption rate, , is the rate per m2 per second. One may obtain an expression for the sticking coefficient, , by equating the rate of adsorption in the kinetic model to the expression used to define . If the rate constant is k, the turnover frequency in the forward direction is

r k

p 1 po

If we consider a surface with area A (m2) and density of sites d (mol m2), we can equate the rate of adsorption (mol/s) in the two models

r k

dA

mk B T 49

p 1 po

K2

p pB

K1

pA

A*

B*

pA

which after simplification gives us the desired transformation between rate constant, sticking coefficient, rate and turnover frequency. From this equation it is obvious that while the rate constant, k, generally does not depend on the coverage, the sticking coefficient must depend on the coverage. The value usually reported for the sticking coefficient is the initial sticking coefficient 0 corresponding to 0.

6.9.2 The activation energy for $\sigma$


There is a further complication in the interpretation of the sticking coefficient. If k has Arrhenius form, the activation energy for the sticking coefficient is larger than the activation energy for k by kBT. This difference is significant in most situations.

6.9.3 Flux and exposure


A partial pressure, p, of a gas with molecular weight M, corresponds to a flux, f, p 2 mk B T

where T is the temperature. In words, the flux is the number of collisions per second pr m2. Even at atmospheric pressure the flux is huge. Each atom in a surface exposed to the air is hit by close to 109 gas molecules per second. For all catalytic reactions the flux is in very round numbers a million fold larger than the reaction rate. Except for laboratory experiments under vacuum, the reaction rate is saturated with respect to the flux and an increase in the flux will have a marginal effect on the rate for reactions at 1 atm or above. However, many reactions are performed at high pressures. The reason is purely thermodynamic. High pressures are applied to shift the equilibrium in favor of the desired product.

50

6.9.4 The exposure


Integrating the flux over time gives the exposure, i.e the total number of collisions per m2 . Informally exposures are often reported in Langmuir or as a number of monolayers. One Langmuir is 1 torr second. Within a factor of 2 or so, 1 L corresponds to 1 collision per surface atom. The exact number of collisions can of course be worked out when if we know the details. This is precisely what has been done if an exposure is reported as a number of monolayers. 1 L is one collision per surface atom.

6.10 Temperature programmed desorption


The principle in TPD is that the surface is cleaned and then exposed to a gas at a relatively low temperature. When the desired coverage has been reached, the pressure is reduced, the temperature is increased linearily and the flux of desorbing molecules is measured.

The principle in TPD is that the surface with adsorbed molecules or atoms is placed in front of a mass spectrometer. The temperature of the surface is then increased linearly with time. The final temperaure 51

is high enough to desorp all the adsorbed atoms and molecules but should not be so high that the sample is damaged. The desorption is performed in vacuum. Once a molecule has desorbed it is rapidly eliminated by the pumps. The events in TPD are easy to visualize. At low temperatures the coverage is high, but the rate constant is low and the desorption rate is low. As the temperature increases, the rate constant and thus the desorption rate increases. At intermediate temperatures the desorption rate is high as both the coverage and the rate constants are high. At high tempererature the molecules have desorbed, the coverage is virtually zero and although the rate constant is enourmous, the desorption rate is virtually zero.

6.10.1 Kinetics
When the surface is heated in a vacuum the rate equation for the scheme A2 + * A2* A2* + * 2A* becomes d K1 k2 2 dt If K1 and k2 have Arrhenius form, the kinetic equation becomes

where is the heating rate. This equation is usually integrated using e.g a RungeKutta algorithm. The parameters are T0, A1A2 and E1+E2 addition to the initial value A. Note that only the product of the preexponential factors and the sum of the energies enter into the equation. It is a very general problem in the interpretation of TPD spectra that the As and Es appear so that the A or E for a single step cannot be determined. As the As and Es often are lumped, the generic rate expression for TPD has the form

52

dt

A1 A 2 exp

kB T0

E1 E2 t

dt

where the order, n, is 1 or 2, or, in rare cases, 0.

6.10.1.1 First order kinetics


The kinetic equation is

dt

This is the kinetic equation one would expect for e.g for the mechanism A + * A* The generic shape of the desorption peak for TPD, first order kinetics:

The peak is not symmetric around the peak temperature (Parameters: A=$10^9$~$s^{1}$, E = 100 kJ/mol, initial coverage 0.95):

53

The peak becomes larger, but the peak temperature remains constant when the initial coverage is increased. This is characteristic for first order kinetics and is important for diagnostic purposes (Parameters: A = 109 s1, E = 100 kJ/mol, initial coverage 0.1 (lower peak), 0.3, 0.5, 0.7, 0.9 (upper peak).)

6.10.1.2 Redheads equation


The rate equation for TPD is

dt

where k has Arrhenius form k A exp


E# kBT

TPD is one of the simpler experimental techniques and the rate equation makes it obvious that it should be possible to determine the activation energy for desorption from a TPD spectrum. However, a general problem in TPD is that we cannot measure the coverage during the experiment. Also determination of the absolute reaction rate is rather difficult. One possibility is to determine the activation energy from a detailed simulation of the experiment. While this is doable, we will here rewrite the rate equation for TPD to eliminate the rate and the coverage.

54

The derivative of the rate constant with respect to temperature is

dk dT

E k kBT2

and the derivative of the temperature, T, with respect to time, t, is the heating rate

If we combine these equations, we can determine the derivative of k with respect to t. From the rate equation we can determine the second derivative of the coverage:

dt

nk

This second derivative of the coverage is the first derivative of the rate. The peak in the TPD spectrum is in reality the maximum in rate and thus a zero for the second derivative

d2

dt 2

By straightforward algebra we find




where T is the temperature of the peak maximum. If and only if n=2n1, i.e n=1, we obtain Redheads equation E# kBT

AT
!

where T is the peak temperature. If we know that the desorption follows first order kinetics (n=1) and we have an educated guess on the value of A (e.g from transition state theory), Redheads equation makes it easy to determine E from a measurement of the peak temperature, T. In a TPD experiment, the difference between the smallest and the largest applicable 55

E# k kBT2

nk 2

2n 1

exp

E# kBT

dk dT dT dt

n 1

dt

d2

dT dt

heating rate is often less than a factor of 10. If we dont know A, we could measure the TPD spectrum using a range of heating rates and then use Redheads equation to determing both E and A. However, a variation by a factor of 10 is usually too little for a reliable determination. Finally, a caution about the Redhead equation. The requirement n=1 is not explicit in the equation. For this reason nothing prevents that the equation is applied in situations where the order of the desorption is unknown or, even worse, when the order is known to be different from 1. To describe E obtained in this fashion as an approximation is an understatement, completely wrong is a better description.

6.10.1.3 Second order kinetics


The kinetic equation is d dt

This is the kinetic equation one would expect for e.g for the mechanism A2 + * A2* A2* + * 2A*

The generic shape of the desorption peak for TPD, second order kinetics. The peak is not symmetric around the peak temperature. (Parameters: A=109 s1, E = 100 kJ/mol, initial 56

coverage 0.95. When the preexponential factor is increased, the peak moves towards lower temperatures. When the binding energy increases, the peak becomes broader and moves towards higher temperatures.

6.10.1.4 Zero order desorption


The kinetic equation is

dt

This is the kinetic equation one would expect e.g when the adsorbate rapidly moves around on the surface, but desorption takes place from a small number of defects. This makes the desorption rate independent of the coverage.

The generic shape of the desorption peak for TPD, zeroth order kinetics. (Parameters: A=109 s1, E = 100 kJ/mol, initial coverage 0.95.)

57

58

7 Catalyst structure and texture


7.1 Catalyst structure
Many, but not all heterogeneous catalysts used in industry, consist of small metal particles on a support. As the catalytic reaction proceeds at the surface of the metal particles, the catalysts have been prepared to expose a large metal area, typically 10 100 m2 per gram of catalyst. The function of the support is primarily to increase and stabilize the area of the metal particles. The size of the metal particles may be reported in three closely related ways The area of the metal surface The diameter of the metal particles The dispersion, defined as the fraction of all metal atoms that are present at the surface.

7.2 Pore structure


A catalyst is porous and contains pores. The pores are of course of a very irregular shape.

When the pore volume is measured, one usually find pores of two distinct sizes: Macro pores have a diameter of 100 nm or more. Macropores are formed as cracks between between crystallites. Micropores have a diameter of 1 nm or less. Micropores are formed by the roughness of the surface. 59

7.3 Catalyst models


In the study of reaction mechanisms, industrial catalysts are complicated to work with. For this reason, mechanistic studies are often performed on model systems, such as single crystals. For this purpose, single crystals are available of most metals in sizes of e.g 10 by 10 mm. Experience has shown that it is often much easier to study a number of single crystal surfaces, deduce the reaction mechanism and then verify this for the real catalyst, than to attempt to deduce the reaction mechanism directly from studies of the catalyst.

7.4 The terraceledgekink model


The surface of real crystals are not perfect. At high temperatures thermal disorder will lead to a nonzero concentration of defects. At low temperatures, defects created at high temperatures may be frosen at the surface, defects in the bulk may extend to the surface. Also the orientation of the surface with respect to the crystal lattice may be oriented so that an atomically flat arrangement of the atoms is impossible. The simplest model of a surface with defects is the terraceledgekink model. In this model only 4 types of defects are assumed present in the surface:

60

A kink is a defect formed at the end of the ledges.

If we tilt the surface to make the ledge horizontal, it should be obvious that the only difference between ledges and terraces is the width. Adatoms are a limiting case of a terrace made up of a single atom.

7.4.1 Defects excluded in the TLKmodel


Some of the defects excluded from the TLK model are cavities of missing atoms, surface dislocations and bulk defects extending to the surface. The success of the TLKmodel demonstrates that although such defects do exist and are interesting, they are in many contexts not terribly important.

61

7.5 The stereographic map


If we look at a sphere, all possible surface orientations will be present:

Only a few of the more important orientations have been shown. However, depending on the symmetry of the lattice, we have already shown some orientations where the corresponding surface planes are equivalent. As a example, for the face centered cubic lattice the (100), (010) and (001) planes look identical and the (110), (101) and (011) planes look identical. But the sphere also has a backside with a (100) plane, a (010) plane etc what about them ? The answer comes in the form of the stereograpic map. The stereographic map is an area on the unit sphere, which contains one representative for each equivalent surface orientation. For the simple cubic, the body centered cubic and for the face centered cubic lattices the corners of the stereographic map are (100) (110) and (111). For BCC and FCC (100) (010) and (001) are equivalent. For the hexagonal closepacked lattice, (100) and (010) are equivalent, but (100) and (001) are not eqivalent and the stereographic map is considerably larger for HCP than for SC, BCC and FCC.

62

7.6 Bodycentered cubic lattice


Besides the conventionel, cubic cell, the BCC lattice can be build from a primitive cell. The primitive cell is akward for many purposes. First it is a parallelipiped and not cubic. Secondly, the crystallograhic directions are defined with respect to the conventional cell.

The figure shows the conventionel unit cell for BCC, this cell is cubic and contains 2 atoms. Each atom has 8 neighbors.

63

The BCC(100) surface. Each surface atom has 4 neighbors in the surface plane and 1 neighbor in the plane below.

The BCC(110) surface. Each surface atom has 4 neighbors in the surface plane and 2 neighbors in the plane below. The (110) plane is the most open of the three basal planes for BCC.

The BCC(111) surface. This surface is very open, both the atoms in the first and in the second layer have lost neighbors. The atoms in the first layer have 3 atoms in the second layer and 1 neighbor in the third layer. The atoms in the second layer have 3 neighbors in the first layer, 3 in the second layer and 1 in the third layer.

7.7 Facecentered cubic lattice


Besides the conventionel, cubic cell, the FCC lattice can be build from a primitive cell. The primitive cell is awkward for many purposes. First it is a parallelipiped and not cubic. Secondly, the crystallograhic directions are defined with respect to the conventional cell. The figure shows the conventional cell for FCC. This cell is cubic and contains 4 atoms.

64

The FCC(100) surface. Atoms in the first layer have 8 neighbors.

The FCC(110) surface. This surface is the most open of the three basal planes for FCC. Atoms in the first layer have 7 neighbors and atoms in the second layer have 11 neighbors.

The FCC(111) surface. Atoms in the first layer have 9 neighbors. (111) is the most closepacked of the three basal planes for FCC.

65

7.8 Hexagonal closepacked lattice


Both FCC and HCP are closepacked structures, FCC has an ABC stacking, HCP has ABA stacking. The HCP(100) surface.

The HCP(001) surface.

The HCP(110) surface.

7.9 Adsorbates and impurities


Adsorbates and other foreign atoms may form patterns on the surface of a crystal. Fortunately, the most common patterns are rather simple and a simple notation is used to describe these lattices.

66

p(2x2) on FCC(111). The "P" (primitive) indicates one foreign atom and (2x2) indicates the size of the unit cell relative to the original cell.

p(2x3) on FCC(111). Again "P" indicates one foreign atom and (2x3) indicates the size of the new unit cell. The difference between p(2x2) (above) and p(2x3) (left) is in the size of the unit cell. Note that p(2x2) has a {\em higher} concentration of foreign atoms than p(2x3).

c(2x2) on FCC(100). The "C" (centered) indicates two foreign atoms in the cell and (2x2) indicates the size of the unit cell relative to the original cell. Note that c(2x2) corresponds to a concentration of foreign atoms, which is exactly twice that of p(2x2).

p(1x1) on FCC(100). One foreign atom ("P") for each (1x1) atom in the surface.

67

7.10 The Wulff construction


The surface energy is different for different facets. For a crystallite consisting of a given number of atoms, the equilibrium shape is the shape which minimizes the surface (Gibbs) energy. If we know the surface energies, the equilibrium shape can easily be determined from the Wulff construction: First, in 2 dimesions: In a polar coordinate system, draw a vector parallel to the normal of the surface and with length proportional to the energy of the surface. At the endpoint of the vector, draw a tangent line. Repeat for all surfaces. The equilibrium shape is the area limited by the tangent lines. Dont forget to use the symmetry of the crystal. For a cubic crystal the surface energy for (100) will give you 6 vectors and the surface energy for (111) will give you 8. It can easily happen that a particular surface has such a high energy, that is will not be present at all in the final construction.

68

The construction in 3 dimensions is in principle the same. You use a sperical (not polar) coordinate system, you draw the vectors and then the tangent plane (not line). The figure below shows the Wulff construction for a typical FCC metal. The object is a cubo octahedron. The larger hexagonal facets have a (111) orientation and the smaller, square facets have a (100) orientation.

7.11 Homogeneous catalysts


Heterogeneous catalysts have a number of advantages and disadvantages compared to homogeneous catalysts. In heterogeneous catalysis, the catalyst and the product can easily be separated. The solubility of some homogeneous catalysts is rather small, Some heterogeneous catalysist have problems with diffusion in the pore system. Homogeneous catalysis can usually be performed at mild reaction conditions. The high temperature in heterogeneous catalysis reduces the selectivity. Heterogeneous catalysis can be performed over much larger ranges of reaction conditions, this is important when the equilibrium constrains the reaction to extreme conditions. Homogeneous catalysis is possible in liquid phase and can be performed for reactants that are only stable in solution. Heterogeneous catalysis in solution is difficult.

69

8 Preparation of catalysts
The details of the manufacturing process for a given catalyst is a buisness secret of the catalyst manufacturer and only the broad principles are available in the open literature. The secrecy of the manufacturing processes is maintained partially by discouraging outside investigators through patent protection. However, the most important aspect of the protection is probably that examination of the finished catalyst offers little information on how it was prepared and no information on why a particular preparation procedure was used. For a catalyst the desired properties are high and stable activity high and stable selectivity controlled surface area and porosity good resistance to poisons good resistance to high temperatures and temperature fluctuations. high mechanical strength no uncontrollable hazards Once a catalyst system has been identified, the parameters in the manufacture of the catalyst are If the catalyst should be supported or unsupported. The shape of the catalyst pellets. The shape (cylinders, rings, spheres, monoliths) influence the void fraction, the flow and diffusion phenomena and the mechanical strength. The size of the catalyst pellets. For a given shape the size influences only the flow and diffusion phenomena, but small pellets are often much easier to prepare. Catalyst based on oxides are usually activated by reduction in H2 in the reactor. Some catalysts are prereduced at the factory and transported to the plant covered by a thin layer of oxide.

8.1 Shape

70

Pellet. Catalyst pellets for fixed bed reactors are typically 1.5 10 mm in diameter.

Ring. Typical sizes are 620 mm. The central chanel significantly improves the transport of the gas both in the pellet and in the catalyst bed.

Extrudate. Typical sizes are 15 mm in diameter and 1030 mm in length. Extrudates are only used for impregnated catalysts.

71

Pellet with several chanels. Typical sizes are 2040 mm in diameter and 1020 mm in height. Compared to the ring, the use of several, smaller chanels changes the tradeoff between mechanical strength of the pellet and mass and heattransport phenomena.

Fragment. Typical sizes are 112 mm. \par Fragments are usually prepared by crushing of catalysts manufactured by fusion.

Sphere. Typical sizes are 312 mm. Mainly used for impregnates catalysts.

72

Monolith. Typical sizes are from 50x100x200 mm up to 250x250x250 mm or more. The gas flows along the large chanels, which constitutes almost all of the volume. Monoliths are used when the flow rate of the gas is high, or when the gas contains dust or soot.

8.2 Precipitation
Catalysts are prepared by precipitation typically by mixing one solution with another solution or a suspension. The precipitate is washed and filtered (or filtered and washed), dried, calcined and crushed to a fine powder. A binder, often graphite or stearic acid is added and the powder is tabletted. Alternatively, the catalyst may be shaped before calcination by extruding the filtercake. The filtering and washing are particularily troublesome. A clogged filter will disrupt the production line and may be difficult to fix in a safe way. The wash liquid may contain small concentrations of transition metals or organic amines in addition to larger concentrations of less troublesome compounds. Disposal of this liquid may add significantly to the total manufacturing cost for the catalyst.

8.3 Pelletizing
Many catalyst are manufactured as powders which are pelletized as one of the last steps in the manufacturing process. The pellets are prepared by compressing the powder. The powder flows through a feeder mechanism, and is filled into a rotating die, where it is compressed by pistons.

73

At the start of the cycle, the lower piston is in the upper position. The pellet prepared in the previous cycle is pushed out as the die passes under the feeder mechanism. Catalyst powder fills the boring as the die moves under the feeder mechanism while the lower piston moves down. The feeder mechanism scrapes the upper end of the boring clean. At this moment, the position of the lower piston defines the amount of catalyst powder in the finished pellet. The upper piston moves down by a fixed length. At this moment, the position of the upper piston defines the dimension of the finished pellet. The pressure applied to the pellet and the density of the finished pellet are implicitely defined by the mechanical properties of the powder and by the movement of the pistons. The cycle is completed when the upper piston moves up and the lower piston pushes the finished pellet out of the die. Before the catalyst powder is pelletized, "binders" such as graphite or stearic acid are added. While stearic acid might act as a binder at room temperature, graphite is better known as a lubricant than as a glue. One might suspect that the "binders" work not as glue but by improvement of the flow properties of the powder.

8.4 Fusion
Catalysts based on metal oxides may be prepared by fusion, crushing and screening. Large pieces are returned to the crusher, small pieces are returned to the furnace. The finished catalyst consists of irregular pieces of a narrow range of sizes. Fusion is only possible for catalysts that are conductors at high temperatures. Most oxides are insulators at room temperature and become conductors at higher temperatures.

74

The catalyst mass is placed as a powder in a electro furnace. The mass is heated by passing a large current through 3 graphite electrodes. Each electrode is equipped with a mechanism for raising and lowering the electrode. At the start of the cycle, crushed catalyst mass is filled into the furnace and the electrodes are lowered. The voltage is fixed and the current is regulated through the height of the electrodes. The molten mass is stirred by the electromagnetic fields generated by the huge current flowing between the electrodes. At the end of the process, the electrodes are raised and the molten catalyst is poured out of the furnace. The furnace is operated at a rigid schedule. The thermal insulation is provided by a thick layer of catalyst mass. If the furnace spends too long time at high temperatures, too much of the catalyst mass will melt. The catalyst has no porosity before reduction. The poresystem is created when the metal oxide is reduced during the activation of the catalyst. The lack of porosity before reduction makes the reduction tricky and the catalyst may be supplied in prereduced form. The prereduced catalyst is first completely reduced then passivated by gentle oxidation to allow it to be transported and loaded safely.

8.5 Catalyst supports


Al2O3 The most widely used supports are gammaAl2O3 and etaAl2O3 both have a high area and a good stability. The structure of both gammaAl2O3 and etaAl2O3 is a more or less skew FCC packing of oxygen ions with Alions distributed somewhat disordered in the holes between the Oions. The disorder is the reason why Al2O3 can disolve other ions. Al2O3 is made by precipitation of Alhydroxide followed by calcination. SiO2 Kieselguhr has area 2040 m2/g. Silica gel has areas up to 700 m2/g.

75

Silica gel is prepared by adding acid to a solution of Na4SiO4 and Na2SiO3. Activated carbon Activated carbon has areas up to 1200 m2/g. Activated carbon is prepared by oxidation of porous carbon and contains significant amounts of oxygen. TiO2 TiO2 is more expenside and less used than Al2O3 and SiO2 ZrO2 ZrO2 is rather expensive and is only used as support for catalysts operating at extremely high temperatures. MgO The mechanical strength is poor. Not used in industry. ZnO ZnO has some tendency to reduction. Not used in industry. Cr2O3 Cr2O3 has some tendency to reaction with H2O. High surface area facilitates the oxidation of Cr(III) to Cr(VII). Proper destruction of the used catalyst is rather expensive.

8.6 Impregnation
Impregnated catalysts are prepared by impregnating a metal salt on a porous support. The metal loading in the finished catalyst is typically 15 %. When liquid is slowly added to a porous solid powder, the liquid is first absorbed in the pores and the powder will flow as if it is dry. When the pores have been filled the outside of the grains rather suddenly become wet, the grains will tend to stick together and the powder will form lumps instead of flowing freely. The situation when the pores have been filled but the outside of the grains is dry is callled incipient wetness and can easily be detected by shaking or stirring the powder. If you want a handson illustration of this effect, put a handful of dry sand into a cup and add water one spoonfull at a time while shaking the cup. At first the surface of the sand is dull. At some point the surface of the sand suddenly becomes glistening, and the sand is quite sticky. This is the incipient wetness. Add one more spoonful, and the sand becomes wet and starts to flow like mud.

76

Catalysts are prepared by impregnation by spraying a solution of a metal salt onto pellets of a porous support until incipient wetness. The pellets are then dried and calcined to transform the metal into insoluble form. The metal salt can be deposited homogeneously through the pellet or most of the metal may be deposited near the outside of the pellet. The distribution of the metal is controlled through the pH or through addition of chelating agents to the impregnation liquid. Compared to precipitation, impregnation offers a number of advantages The pellets are shaped before the metal is added. The filtering and the wash of the catalyst are eliminated. Small metal loadings are easily prepared. Impregnation offers some control over the distribution of the metal in pellets. and disadvantages High metal loadings are not possible. A good impregnation solution may be impossible to find.

8.7 Metal sponges and colloids


Highly porous metal sponges can be made in aqueous solutions by dissolution of aluminium alloys by a strong base. The metal sponges are typically used for hydrogenation in laboratory syntheses. Raney nickel is made by dissolving an (Al,Ni)alloy containing 50 % Ni in 20 % NaOH in water. The nickel sponge is black, the area is 80100 m2/g. The nickel contains some Al and Al2O3. The most used Raney metals are Ni, Co and Fe. Metal colloids can be prepared by reduction of metal salts in solution using CO, hydroxylamine, oxalic acid, formaldehyde, citric acid or sodium citrate.

77

9 Poisoning and deactivation


A model for the changes in reactivity for a reaction on a catalytic surface in the presence of adsorbed inactive atoms. The model is based on a mean field description of the formation of partly disordered structures for the adsorbed atoms. If other gases than the reactants and products of the catalytic reaction are present and if these foreign atoms are chemisorbed much more weakly than the intermediates of the catalytic reaction, the presence of the foreign gas will be of no consequence for the kinetic of the catalytic reaction. In this case the foreign gas behaves as an inert. If the foreign gas is chemisorbed much more strongly than the intermediates, the surface will to a large extent become covered by adsorbed foreign atoms. In this case the foreign gas behaves as a poison and the consequences for the kinetics will mainly be a large decrease in the reaction rate. The most complex situation occur when the binding energy for the foreign atoms is about equal to the binding energy for the reaction intermediates. The coverage of foreign atoms on the surface will then depend on the reaction conditions. The kinetics of the reaction under dynamic poisoning may be quite complicated. In the present chapter we formulate a simple model for the changes in reactivity of the catalytic reaction under the conditions of a partial poisoning by adsorption of foreign atoms. For the understanding of the kinetic phenomena an approximate but transparent model is preferable.

9.1 Deactivation
A number of different processes contribute to the loss of catalytic activity.

9.1.1 Sintering
The catalytic reaction takaes place at the surface of the catalyst. For this reason the catalyst must have a large area and the active component is present in the catalyst in the form of small particles, which are much less stable than large particles. A catalyst will slowly loose catalytic activity due to growth of the particle and loss of surface area. 78

There are two distinct mechanisms for sintering: The crystallites may continously emit and collect atoms. Since the atoms will be more stable at the surface of a large particle than at the surface of a small particle, the atoms will on average migrate from the small particles to the large.

The particles may move and when two particles come into contact, they may merge into a single, larger particle.

The loss of catalytic activity due to particle growth is irreversible.

9.1.2 Fouling
Catalytic activity may be lost due to the formation of carbon or due to the deposition of impurities or of dust in the catalyst. The formation of carbon is in some cases reversible. The catalyst can be taken of stream and the carbon removed by oxidation. The reactant gas may contain impurities, a few examples are metalorganic compounds in hydrodesulphurization or $H_2S$ in steam reforming. These impurities react with the catalyst and reduces the catalysts activity. Fortunately, the reaction between catalyst and impurity is often very strong and the impurity is completely absorbed in the first few percent of the catalyst bed. The catalyst bed is then designed to be a little larger than initially necessary, and the catalyst temperature is initially slightly lower than the nominal operating temperature. Over the lifetime of the catalyst, the temperature of the catalyst is slowly increased to compensate for the loss of catalytic activity due to fouling. Fouling by dust carried into the reactor can occur, a few examples are removal of $NO_x$ from flue gases or removal of CO from exhaust from car exhaust. In this case the catalyst must be manufactured in a shape, which will allow the dust to pass through the reactor. For most catalysts a much more serious problem with dust may occur, if the catalyst bed is improperly loaded or if there is a problem with excessive vibration in the reactor. In 79

this case dust may be formed by attrition. In this case the resulting loss of catalyst activity is overshadowed by two much more serious problems. The dust may accumulate somewhere in the reactor or immediately downstream of the reactor and rather suddenly block the gas flow. The catalyst dust is highly reactive and may cause a violent explosion when pipes or heat exchangers are opened for maintenance.

9.1.3 Dynamic poisoning


The more interesting situation occurs when the catalyst is partially and reversibly poisoned by impurities in the reactant gas. The degree of loss of catalyst activity then depends on the operating conditions.

80

10 Microkinetic modelling
The link between the microscopic description of the reaction dynamics and the macroscopic kinetics that can be measured in a catalytic reactor is a microkinetic model. Such a model will start from binding energies and reaction rate constants deduced from surface science experiments on well defined single crystal surfaces and relate this to the macroscopic kinetics of the reaction. If one can understand what the basic parameters of the reactants and the surface are that determine the reaction dynamics (activation barriers etc.) then given a microkinetic model one has a knowledge of the factors determining the catalytic activity of the catalyst. The kinetics of a catalytic reaction is usually measured in a reactor under conditions relevant to the industrial process. The measured overall rates can then be fitted to a mathematical model, the macroscopic kinetics. This is extremely convenient for process design purposes. If the aim is to explore the mechanism of the reaction and understand which are the important parameters of the catalyst determining the activity, then a microkinetic model is needed. A microkinetic model is based on a detailed mechanism and independent information about the rates of the elementary steps involved and the stability of the intermediates. The microkinetic model is the synthesis of all the basic knowledge about a reaction over a given catalyst. A kinetic model consists of a description of the elementary steps at the atomic level is presented. Input data for elementary steps are taken from available single crystal studies. The model is successfully tested against kinetic data for a working catalyst. The purpose of our studies is not to present a kinetic model, which will reproduce one or a few of the aspects of watergas shift reaction very accurately. The point we make is that a physically reasonable treatment of the proposed reaction mechanism with kinetic and thermodynamic data measured for Cu single crystals leads to a reasonably accurate description of most aspects of the observed kinetics. The number of parameters in the model is large. One could suspect that only a few parameters are critical, but we cannot know {\em a priori} which. Rather than trying to determine this at an early state we concentrate on the determination of reasonable values for all parameters. When this has been done, we can test the model, against independent experimental data and we can perform a proper sensitivity analysis for the input parameters. We then backtrack and concentrate on the determination of accurate values 81

for the critical parameters. The advantage of this iterative scheme is that a much more complete sensitivity analysis can be made on the full model compared to an analysis for a model where some parameters have been eliminated at an early state. The starting point for microkinetic modeling is the detailed reaction mechanism. Thus, while a conventional kinetic model is formulated as the rate for an apparent gas phase reaction, the surface species are explicitly included in a microkinetic model. Kinetic data Electron structure Single crystals

Important aspects

Model

Evaluation

Visualization

Interpretations

In a microkinetic model, the simplest feasible model at the molecular level is formulated based on available data. The model is evaluated through simulation of kinetic data for the catalytic reaction at high pressures and the results of the simulation is compared to existing kinetic datasets. Successful models are useful for visualization of the reaction as well as for detailed investigation of kinetic and mechanistic features.

10.1 Limitations
1. We will limit ourself to catalytic reactions. Surface sites consumed in the adsorption 82

2. 3. 4. 5.

of reactants must be regenerated in the desorption of products. The net production of any surface intermediate must be zero. We describe the reactions at the molecular level. The mass balance for the atoms is only implicitly described through the mass balances for the molecules. We dont describe the structure of the molecules. The gas phase is assumed ideal. Diffusion limitations and temperature gradients are neglected. The reaction mechanism is a LangmuirHinshelwood mechanisms comprising G gases, S surface species and R elementary steps, where G, S and R are arbitrary.

10.2 Level of the model


The modelling of a surface chemical reaction can take place at many different levels of complexity. At the most fundamental level the time development of the reaction system is followed in detail and at the most approximate level only gross averages are considered.

10.2.0.1 The full reaction dynamics


At the most fundamental level one follows the time development of the system in detail. The reactants are started in a specific initial (quantum) state and the equation of motion are propagated to give the final state. The equation of motion of the system is the time dependent Schringer equation, or, if the atoms involved are heavy enough (not H or Li) Newtons equation. The starting point is the adiabatic potential energy surface on which the process takes place. For some reactions electronic excitations during the reaction are important and must be included in addition to the electronically adiabatic dynamics. The adiabatic potential energy surface is the ground state electronic energy of the system as a function of all the degrees of freedom of the system. Such a detailed description could in principle be made of every elementary step in a given reaction, while a dynamical simulation of a whole chemical process including several elementary steps is usually impossible. Typically intermediates in a reaction can have lifetimes which are many orders of magnitude larger than typical times for a dynamical simulation (a few picoseconds). Another point is that if the aim is to get an elementary reaction rate for a system at a given temperature, the full dynamical approach may be too detailed.

10.2.0.2 Mean field model


In this description only average properties are considered. The rate of a given elementary step involving adsorbates A and B are assumed given by 83

r = k AB where A and B re the coverages of A and B, respectively, and the rate constant k usually has an Arrhenius form. The activation energy and prefactor may still depend on the surroundings, but only in an average way through the coverages. In the case where there is one slow step in the reaction mechanism, the solution for the rate of the catalytic reaction is straightforward. When the reaction mechanism has more than one slow step, the solution for the rate is more complicated, but the final expression is remarkably similar. The kinetic model may be formulated using kinetic equations for all steps or using equilibrium equations for all but the slowest steps. The latter approach reduces the computational effort and leads to a kinetic expression, which is far easier to analyze. However, if a step, which is slow in reality, is modeled by an equilibrium equation, the microkinetic model becomes unrealistic and it may in some cases be the simplest to treat a problematic step by a kinetic equation.

10.3 Types of models


Even with the limitations listed in the previous section, there are a very large number of models. We will our attention to the following 6 models: Molecules have a name, a stoichiometry and a stability. One of the reaction steps is assumed rate limiting, this step has a forward rate constant. This model maps onto the quasi equilibrium approximation, Section \ref{sec:solve/quasieq}. In this model the equilibrium constants for all steps are calculated from stoichiometry and from the stability of the molecules. The rate of the fast steps is not available. Molecules have a name and a stoichiometry. Each step has a thermodynamics. One of the reaction steps is assumed rate limiting, this step has a forward rate constant. This model is a trivial variation the previous model. The equilibrium constants for the steps are among the input parameters. The stability of the intermediates and the rate of the fast steps are not available. Molecules have a name, a stoichiometry and a stability. Each reaction step has a forward rate constant. This model maps onto the steady state approximation, Section \ref{sec:solve/steady}. 84

In this model the equilibrium constant for the steps are calculated from the stoichiometry and from the stability of the molecules. Molecules have a name and a stoichiometry. Each step has a forward and backward rate constant. This model is a trivial variation of the previous model. The equilibrium constants for the steps are calculated from the forward and backward rate constants. The stabilities of the molecules are not available. Molecules have a name, a stoichiometry and a stability. For each configuration of molecules at the surface, there are a number of possible events. The events occur randomly with a characteristic rate for each type of configuration. This model is the kinetic Monte Carlo simulation. In this model the equilibrium constants for the steps are calculated from the stability of the molecules. The rates are calculated by simulation. Molecules have a name and a stoichiometry. For each configuration of molecules at the surface, there are a number of possible events. Each event has a thermodynamics and occur randomly with a characteristic rate for each type of configuration. This is a trivial variation of the previous model. The equilibrium constants are among the input parameters. The rate is calculated by simulation. The stability of the molecules is not available. Hybrids between these 6 models are possible, e.g by specifying the thermodynamics by the thermodynamics for some steps and the stability for some of the molecules. However, these hybrids are much more difficult to work with than the 6 models defined above, in particular for large mechanisms.

10.4 Phases in a simulation


Simulations consists of 3 phases: initialization, equilibration and production. At the initialization, a configuration is generated or read from a file. The system is then equilibrated by propagation in time. When the system has equilibrated, the production phase starts. Data are stored on disk when the production phase starts and then after each major time step.

85

Minor timestep

Major timestep

Initialization Equilibration Production

Data

Data

Data

Data

Data

Data

The major time step is an abstraction. The program internally builds up the major time steps through a series of minor time steps. It is a complication that in KMC the minor time steps are of varying length. As we want the output from the simulation program stored at constant time intervals, special care is required. For mechanisms, there are two important aspects, stoichiometric and thermodynamic consistency. However, we must make the description of the kinetic model more concrete before we can implement the approximation schemes and arrive at a soluble model.

86

10.5 Input parameters


The input into a microkinetic model may be measured or calculated. Measured data for the catalyst The input parameters for a microkinetic model may be taken from measured adsorption and reaction rates for the catalyst, measured heats of adsorption together with thermodynamic data for the gas (or liquid) phase above the catalyst. This gives information of direct interest for the catalyst system considered, but often the interpretation of the experiments is difficult due to the fact that usually the state of the active surface is not known and may vary with the conditions of the experiment. Measured data for catatalyst models The input parameters can be taken from measurements on model systems. If the structure of the catalyst is known and one has a suspicion which is the active crystal surface and do experiments on this model with all the chemical phase, then one can isolate this phase, usually in the form of a single and structural characterization tools available in surface science. Calculated data Electronic structure theory has developed to a point where realistic bond energies and activation barriers can be calculated. Typically the model catalysts used in such calculations are even more idealized than in the surface science experiments (perfect surfaces, ordered overlayers etc.), but the insight into the details of the potential energy surface of the reaction is much greater.

10.6 Test of the model


Fragments of the kinetic model may be tested by detailed simulation of the same experiments as have been used to determine the parameters in the model. Failure to reproduce these experiments is entirely possible if a wrong reaction mechanism has been assumed. A sensitive test can be made by simulation of a reactor through numerical integration of the rate expression. This will test the model at higher pressures, frequently at higher coverages, in the presence of all reactants, intermediates and products. The model can be further tested for internal consistency. Steps treated through kinetic equations should be slow at least under some conditions or the model may obviously be 87

simplified. The significant parameters should obviously be determined rather directly from experiments. The coverages of the intermediates calculated under reaction conditions should not be greatly different from the coverages in the experiments used to determine the parameters.

10.7 Significant parameters


While the number of parameters may be large, only a few of the parameters are usually significant. The significant parameters are easily determined by calculation of the sensitivity of the calculated rate at typical conditions to a small variation in the value of each parameter. For the NH3 synthesis, for instance, the rate of dissociation of N2 and the binding energy for N* are the most significant parameters and the kinetics of desorption of N2 in a TPD experiment is rather closely related to the kinetics of NH3 synthesis.

10.8 Applications of microkinetic modelling


Once the kinetic model passes the tests, the model may be used to understand details of the catalytic reaction, such as the origin of the macroscopic activation energy, the reaction orders or the structure of empirical kinetics. In spite of the shortcomings of the modelling, the real strength is that they can be used to understand {\em variations} in the catalytic activity from one system to another. The stability of the intermediates and the activation barriers are among the input parameters for the microkinetic model, and it is straight forward to calculate the effects of changes in stability for some or all the intermediates.

10.9 Ammonia synthesis


1) 2) 3) 4) 5) 6) 7) N2(g) N2 + * N* + H* NH* + H* NH2*+ H* NH3* H2(g) + 2* N2* 2N* NH* + * NH2* + * NH2*+ * NH3(g) + * 2H*

88

In this mechanism, the second step is known experimentally to be slow. The competition for adsorption sites is very important for the kinetics of a heterogeneous catalytic reaction. For this reason sites, *, are included as a reactant in the kinetic model. As a site must be either free or occupied by one of the surface intermediates, there is a conservation law for the coverages 1 X where X is the coverage by the intermediate X. In writing this equation we have implicitely defined X =1 to be saturation. With this convention, coverages may be interpreted as probabilities. For each of the fast steps, we get an equilibrium equation, e.g p K 1 N2 N2 po For a slow step we get an kinetic equation, e.g k2 2 r k 2 N2 * N K2 which expresses the net rate, r, as the difference between the forward rate, r+, and the backward rate, r.
" 

10.9.1 Solution for *


When the mechanism has only one slow step, the system of equilibrium equations and the rate equation may be solved with respect to 1
*

p0

K3K4 K5K6K p

K4 K 5 K 6 K7 pH

K5K6K7 p

The coverages by intermediates may be expressed by * and the partial pressures of the reactants and products.

89

K1 PN

p NH p
3

1 o 2 3 2 7 3 2 H2

p NH

p NH

p NH
1 o 2 1 2 H2

K 6 po

1 2 7

p p

1 2 H2 1 o 2

N2*

N*

K3K4 K5K6K p p NH
NH *

3 2 7

3 2 H2

K 4 K 5 K 6 K 7 pH p NH
3

NH 2 *

K5 K6 K7 p

1 o 2

p
*

1 2 H2

K 6 po

H*

1 2 7

p p

1 2 H2 1 o 2

K g pH

10.9.2 Calculation of the equlibrium constants


The equilibrium constants are expressed in terms of the molecular partition functions 2

q trans

mk B T h2

3 2

1 exp

kBT

q vib

q rot

kBT 2B 1

r 2 k2 K1

pN

p2 po NH
3

NH 3 *

p NH

p p NH po
3

K1

PN

2 *

90

q q trans q rot q vib exp

Eg

k BT

10.9.3 Calculation of reaction enthalpies


The enthalpy of formation for reactants and intermediates may be calculated from the expression for the partition function H# k B T 2

d ln q dT

The enthalpy obviously depends on temperature. However, even at T=0 the enthalpy of an intermediate may differ substantially from the electronic binding energy due to the zeropoint motion in vibrational degrees of freedom.

10.9.4 Calculation of the rate


The solution for the rate of the catalytic reaction is straight forward pN p
o

The parenthesis has a simple interpretation in terms of the thermodynamics of the gas phase while the interesting part of the kinetics comes from the factor *2.

10.9.5 Practical calculation


The calculation above closely follows the sequence of calculations when a reaction mechanism is turned into a rate expression. In principle, such a derivation is done once for a proposed reaction mechanism and the rate expression is then used repeatedly to calculate the reaction rate at different reaction conditions. The derivation of the rate expression takes, say, one hour of computation and the analysis of the rate expression takes, say, one week of programming followed by one month of 91

r 2K 2 K 1

1 NH 3 K g po

H2

3 2 {

computation. For this reason the derivation of the rate expression is usually repeated a few times to check for errors. Once we have derived a rate expression and carefully checked the derivation, the sequence of steps in the calculation of the rate is quite different from the the sequence of steps in the derivation of the rate expression. The main steps in the calculation of the rate from the rate expression are: Properties of intermediates are deduced from spectroscopic data. This is done once for each project. At the actual reaction temperature, the molecular partition function for all intermediates are calculated from the properties of each intermediate. The equilibrium constant for each step is alculated from the molecular partition functions. The coverage by free sites is calculated from the equilibrium constants and the partial pressures of the gases. The coverage by intermediates are calculated from the the coverage by free sites, the equilibrium constants and the partial pressures of the gases. The reaction rate is calculated from the rate constant, the equilibrium constants and the coverage by intermediates.

10.9.6 Parameters
The input parameters for the model are the thermodynamics of the gas phase, chemisorption energy and spectroscopic properties for the intermediates, the kinetic parameters for the rate limiting step and the number of active sites on the catalyst. No reference to experimental data for catalytic reaction rates are made in the determination of the input parameters. Only few of the input parameters are critical for the prediction of high pressure reaction rates. The critical kinetic parameters are the parameters for N2 adsorption and dissociation on Fe(111) and on K/Fe. Fe(111) EN2 38 kJ/mol EN H2# A2 91 kJ/mol 47.4 kJ/mol 1.37 10 11 s1 K/Fe 51 kJ/mol 91 kJ/mol 47.4 kJ/mol\ 5.89 10 10 s1

92

EN2 and EN are the ground state energies of N2* and N*, respectively. H2# and A2 are the activation energy and the prefactor, respectively, for reaction step 2.

10.9.7 Stability of intermediates 10.9.8 The NH3 concentration


NH3 concentration in a catalyst bed operating at 100 atm, 673 K, inlet 25 % N2 and 75 % H2, outlet 19.2 % NH3 corresponding to 75 % approach to equilibrium.

Significant variation in coverages occur from inlet to outlet. The most drastic variations take place in the first approx 3 % of the bed. In this reaction zone the reaction rate is high. It is a general feature at all temperatures and pressures that the coverages and the reaction 93

rate change dramatically when the NH3 concentration increases from zero to a few percent of the equilibrium concentration.

10.9.8.1 Coverage by *
Coverage by * through a catalyst bed operating at 100 atm, 673 K, inlet 25 % N2 and 75 % H2, outlet 19.2 % NH3 corresponding to 75 % approach to equilibrium.

Throughout the bed the coverage by free sites is small. In the present case the coverage by free sites defreases from 5102 at inlet to 2104 at outlet. This is remarkable as the synthesis of NH3 proceeds through the trapping of N2 on the free sites. The most important parameter in the calculation of * is the groundstate energy for N* which is determined from experiments. The NH3 synthesis rate is proportional to *2. Since the calculated synthesis rate is in reasonable agreement with independent measurements, it is unlikely that the low value of * should be significantly in error.

94

10.9.8.2 Coverage by N*
Coverage by N* through a catalyst bed operating at 100 atm, 673 K, inlet 25 % N2 and 75 % H2, outlet 19.2 % NH3 corresponding to 75 % approach to equilibrium.

The surface species N*, NH*, NH2* and NH3* are all mainly formed from NH3. For this reason the coverage by these four species is close to zero at inlet. The sequence of coverages under synthesis is N* > NH* > NH2* > NH3*. This sequence is determined by entropy rather than enthalpy and is independent of temperature and pressure.

10.9.8.3 Coverage by N2*


Coverage by N2* through a catalyst bed operating at 100 atm, 673 K, inlet 25 % N2 and 75 % H2, outlet 19.2 % NH3 corresponding to 75 % approach to equilibrium.

N2 is a weakly adsorbed species. It is mainly formed from N2. The coverage by N2 is small under synthesis conditions. 95

We will later find that N2* is also too small to influence the macroscopic kinetic parameters. The kinetics of NH3synthesis might thus have been adequately described without explicitely including N2* in the mechanism. However, the existence of N2* is important for the understanding of two aspects of the adsorption of N2 on Fe: the origin of the uniquely low sticking coefficient for N2 the description of chemisorption of N2 at low temperature. Neglecting N2* is thus conceptually less simple than the explicit consideration of N2* in the reaction scheme.

10.9.8.4 Variations with temperature

* increases rapidly with temperature far from equilibrium. Closer to equilibrium the increase is more moderate. The increase is most pronounced at low temperatures and low pressure. * decreases with temperature. The variation is most pronounced at smal consersions. At zero conversion, * is essentially zero at all temperatures. * decreases with temperature at higher conversions, while it increases with temperature at higher conversions. This complicated behaviour is caused by the decreasing tendency of H2 to adsorb with increasing temperature combined with the competition between H* and the much more strongly adsorbed N*.

96

10.9.9 The reaction rate


Reaction rates calculated for a catalyst operating at 100 atm, 673 K and inlet gas with composition 25 % N2 and 75 % H2. The exit NH3 concentration is 19.2 % corresponding to 75 % approach to equilibrium.

Significant variation in coverages occur from inlet to outlet. The most drastic variations take place in the first approx 3 % of the bed. The reaction rate is high at inlet due to the relatively high coverage by free sites followed by a rapid decrease in the reaction rate as NH3 is formed. It should again be stressed that N* is mainly formed from NH3 and not N2.

10.9.10 The turnover frequency


For a catalyst operating at 673 K, 100 atm and 28 % approach to equilibrium the turnover frequency is 0.029 s1. This can be further interpreted by analyzing the forward and backward rate. Each active site turns over 0.031 times each second, 0.030 times in the forward rate and 0.001 times in the backward direction. Each turnover results in the synthesis of two NH3 molecules. Taking the surface coverage into account the turnover frequency at these conditions corresponds to the synthesis of 71 NH3 molecules per free site per second. At inlet conditions the turnover frequency corresponds to each site synthesizing 250 and decomposing 0 NH3 molecules per second.

10.9.11 Lifetime of intermediates


We do not know the forward and backward rate of the fast steps. For this reason we can only estimate the lifetimes of the intermediates.

97

Assuming that the sticking coefficient for N2* is 102 and refering to a catalyst operating at 673 K, 100 atm and 10 % to equilibrium, the coverage by N2* is typically 410, from which the average lifetime for N2* can be estimated to be 11013 s. For an Natom on the surface, the upper limit is essentially zero at inlet conditions. At 673 K, 10.1 MPa, and 10 % approach to equilibrium, the upper limit to the lifetime is 1.10 s for N*, 0.18 s for NH*, 24 ms for NH2* and 0.4 ms for NH3*.

10.9.12 Activation enthalpy


The activation enthalpy for the catalytic reaction may be calculated from the micro kinetic model as H # k BT

d ln r + dT

where r+ is the forward rate of NH3 synthesis pN p


o 2 *

r + K 2 K1

The result is H# = H2# + H1 2H1N2* 2 (H3+H4+H5+H6+3/2 H7)N* 2 (H4+H5+H6+H7)NH* 2 (H5+H6+1/2 H7)NH2* 2 H6 NH3* H7 H* The terms in the activation enthalpy can be interpreted as follows:

98

Enthalpy H1 (H3 + H4 + H5 + H6 + 3/2H7) (H4 + H5 + H6 + H7) (H5 + H6 + 1/2H7) H6 1/2H7

Reaction N2* N2 + * N* + 3/2 H2 NH3 + * NH* + H2 NH3 + * NH2* + 1/2 H2 NH3 + * NH3* NH3 + * H* 1/2 H2 + *

This table allows us to interpret the activation enthalpy for the synthesis as the activation enthalpy for N2 + 2* 2N* plus the averaged enthalpy to be supplied to create two free sites. The average is formed by weighing by the coverage for each intermediate. The activation enthalpy for synthesis is not constant but depends on the temperature and the surface coverages.

10.9.13 Reaction orders


The reaction orders for N2, H2 and NH3 can be defined as

99

d ln r +

d ln

pi po

No assumptions on N2 H2 or NH3 have been made. From the forward rate of NH3 pN p
o 2 *

r + k2 K1

the reaction orders can be calculated N2 H2 NH3 = = = 12N2* 3N* + 2NH* + NH2* H* 2N*2NH* 2NH2*2NH3*

The reaction orders are not constant but depend on the surface composition.

As N2* is very small under all experimentally feasible synthesis conditions the reaction is always first order in N2. At very small conversions H* is large and H2 is close to 1. This indicates that the 100

reaction is inhibited by H2 at low conversion. At high conversion N* is close to 1, H2 is close to 3 and NH3 is close to 2. At these conditions H2 increases the reaction rate by decreasing N*.

10.10 Why does it work ?


One of the important conclusions of the microkinetic modeling is that even large changes in some parameters do not affect the overall agreement between model and experiment much. The reason is that a larger sticking probability through the principle of detailed balance implies that the TPD rate increases. To get the correct TPD peak, the N bonding energy on the surface must then be increased, this leads to a smaller coverage of free sites which compensates for the larger sticking probability. To see that this is a more general conclusion consider the kinetics in more detail. Using the rate equations given above we can write the rate of NH3 production as

2


where we have also used the equilibrium of reaction step (1) to express the forward rate directly through the gas phase N2 pressure. Likewise, using that all steps after step (2) are also in equilibrium we can express the coverage of free sites as p H2 Kg

K1K2

p NH p
o

1 2
3

giving the following expression for the synthesis rate: p N2 p


o

2


r k

2 N*

Kg

p H2

p NH3

po

r k

K1 K 2

p N2

2 N*

3 2

N*

101

In the last equation we have introduced the N2 TPD rate,


2


r TPD k

2 N*

and it is seen that the synthesis rate can always be expressed as the product of the TPD rate and the approach to equilibrium. If, therefore, one uses a consistent description of the synthesis rate and the TPD rate, including the principle of detailed balance and makes a reasonable fit to the TPD spectrum in the relevant coverage and temperature range, one cannot avoid getting a good description of the synthesis rate. Since the N2 TPD peak is in the same temperature range as typical synthesis temperatures it is, in hindsight, not so surprising as first thought, that something good has come out of relatively simple modeling. The key to the success is consistency.

r r TPD K g

p N2

p H2

p NH3

102

11 Experimental methods
Although experimental methods are outside the scope of this manuscript, we have actually touched upon the computational aspects of a number of experimental methods. Pure surfaces to be used as catalyst models may be prepared in a number of ways: A filament may be purified by heating to high temperature in UHV. For most applications it is a complication that the area of the filament is small. Metal films may be prepared by evaporation. For some applications it is a complication that the crystallography andsometimes even the area of the film is not welldefined. A catalyst may be examined in its working state. However, most catalysts are extremely reactive in its working state and it may be necessary to perform measurements before the catalyst is activated or after it has been passivated. In insitu measurements the catalyst is examined in its reactive state. The advantage is that the catalyst is known to be in the relevant state and this can easily be verified by direct measurement of the catalytic activity. The disadvantage is that the catalyst must be inside some kind of reactor and there is not much choice with respect to pressure, temperaure or gas composition. In exsitu measurements the catalyst is outside a proper reactor and it is in a state which is more or less different from the active state. The advantage is that we have a wide choice of temperature, pressure and gas composition. The disadvantage is that we dont really know how far the state is from the active state and we have no easy way of finding out.

11.1 Titration of active sites


The density of active sites at the catalyst surface is of course of great interest. The density of sites may be determined by two methods In volumetric chemisorption the catalyst is activated, gas is removed by evacuation at a relatively high temperature and the catalyst is then exposed to a known amount of gas. From the final pressure and the volume of the cell, the amount of adsorbed gas may be calculated. If the desorption rate is measured quantitatively in a TPD experiment, the number of active sites in the surface can be determined by integration of the desorption rate across the peak.

103

Temperature programmed reaction

Temperature programmed desorption is one limit of a more general technique, temperature programmed reaction (TPR). In TPR the catalyst is heated in a reactive gas and this opens the possibility of adsorption of the reactive gas and the reaction between the adsorbate and the gas in addition to the desorption of the adsorbate. In an intermediate form, TPD is performed in a flow of an inert gas instead of in a vacuum.

11.2 Rate measurements


Reaction rates are measured in either a plugflow or in a backmix reactor.

104

105

12 Ideal reactors
12.1 Rate
The reaction rate for a catalytic reaction obviously depends on the amount of catalyst, pressure, flow, temperature and composition of the gas. Reaction rates are often reported on different basis The rate per m3 of catalyst bed. The rate per m2 of catalyst area. The rate per kg of catalyst. The space velocity is the flow reported as the volume flow (m3/s) divided by the volume of the catalyst. The rate of reaction may be reported as the space velocity corresponding to a particular concentration of product in the exit gas. This is very convenient for a quick estimate of the necessary size ot the size of the catalyst bed, but very awkward for any other purpose. The reaction rate for the catalyst is the product of the density of sites and the turn over frequecy.

12.2 The tank reactor


The flow in Fi with temperature Ti and mole fractions x1i, x2i, ..., xni, the flow out is Fe with temperature Te and mole fractions x1e, x2e, ..., xne. The mixture in the tank has the same temperature and composition as the product. As the temperature of the flow in and out of the reactor may be different, the enthalpy of compound j hji} at temperature Ti may be different from the enthalpy hje at temperature Te. The heat transfered to the reactor is Q . The mass and energy balances, are
&

106

In 1 ... n sum energy Fix1i ... Fixni Fi Q Fi


 &

+ Produced + r1V + + rnV


i

= Out = Fex1e = = Fexne


i

+
j

x ji h ji + 0

The equations for the tank reactor are "larger" than the equations for the plugflow reactor. However, there is no integration for the tank reactor. For the adiabatic tank reactor, Q 0 . Unless the reaction happens to be thermoneutral, Ti and Te will be different. For the isothermal tank reactor Ti=Te and hji=hje for j=1... n.
&

12.3 The semibatch reactor


The equations controlling the operation of ideal reactors are the energy balance and a mass balance for each reactant and product. We assume that n reactants, products and inerts participate in a reaction with stoichiometric coefficients 1, 2, ... n$. The reaction rate ([mol/s/m^3]) is r

1 d ni V i dt

where V is reaction volume (m3) and ni ([mol]) is the amount of compound i. We use the convention that heat, work and flow is positive when they go into the system. Heat released from the system is a negative amount of heat transfered to the system etc.
i

The mole fractions add up to unity

xi 1

Initially, the reactor is partly filled with some of the reactants in the liquid phase. The reaction starts when other reactants are added, either as liquid or as gas.

107

= Fe
j

F e x je h je

Some of the reaction mixture may be withdrawn from the reactor during the reaction. As a result of the flow into and out of the reactor, the volume of the reaction mixture may change. The reaction rate is r = r(T,c1... c_n) For compound j, the molecular mass is Mj, the heat capacity is Cj and the amount in the reactor is nj. As the temperature of the flow in and out of the reactor may be different, the enthalpy hji of compound j at temperature Ti may be different from the enthalpy hje at temperature Te. The density, , of the reaction mixture depends on the reaction temperature and the composition = (Te, c1, ..., cn) The mass and volume of the reaction mixture is m nj M j
'

If j is added a gas the rate of addition of compound i is n ji F i x ji where xji is the mole fraction of j in the gas and Fi ([mol/s]) is the gas flow into the reactor. However, if it is added as a liquid n ji V i c ji where cji ([mol/m^3]) is the concentration of j in the feed and V i [m3/s] is the flow rate into the reactor. The rate of withdrawal for compound j is n je c je V e where cje is the concentration of j in the product and is V e the flow rate out of the reactor. The mass energy balance (Initial + in + produced = out + final) for component j is
  & & & & & & & &

and the energy balance is 108

n j n ji r
 & 

V n je n j dn j

The solution for the mass and volume of the reaction mixture are dm n ji M j Vr n je M j jMj dt dn j dV 1 dT Mj V dt dt dt The temperature develops as
&  & & !   ' &

12.3.1 The plugflow reactor


We want to calculate the mole fractions, x1,..., xG , coverages G+1,...G+S, and flow F [mol/s] through a catalyst bed operating at temperature T [K] and pressure p [Pa]. The parameters for the catalyst is the porosity , the skeletal density [kg/m3], the density of sites s [mol/m2] and the specific area a [m2/kg]. The parameters for the catalyst bed is the cross sectional area A [m2] and the length L [m]. The mass of the bed is m= LA(1) A slice of length dl has mass dm = A(1) dl contains a gas volume dVg = A dl and the amount of gas

RT 1

The contact time dt is

'

dn

p dV g RT

dT dt

n ji h ji h je njCj

dm

109

&

&

n j h je Q dt
& 

n j h ji dt 0

n je dt

n j dn j h je C p dT

The rate of reaction [mol s1} for gas g in the slice is

sa

r 1

rr

rg

dm

The mole fractions add up to unity x 1 i i

12.3.1.1 The isothermal plugflow reactor


We consider a small slice of volume dV during the time dt. The mass and energy balances for the slice are In + Produced Out 1 ... n sum energy Fx1dt ... Fxndt Fdt Q dt F
 &

+ + + +
i

r1dtdV ... rndtdV


i

From the sum we find


dF r dV

Substituting this equation into the mass balances for each of the compounds we find
1 i

dV dx n dV

Substituting into the energy balance we find 110

dx 1

x1

F ... ... r n xn

ny i

x i h i dt +

F dF

'

dt

dV g

dm

'

= = =
i

(F+dF)(x1+dx1)dt ... (F+dt)(xn+dxn)dt (F+dF)dt x i dx i h i dt

dtdV =

Even for simple reactions analytical solution is impossible. Instead we use an ODE solver to integrate the equations. The boundary values are the inlet flow and inlet concentrations.

Q
&

V


h i dV

111

13 Fit of kinetic data


The core problem in experimental kinetic studies is to determine kinetic parameters from measurements. A prerequisite for kinetic studies is a reliable knowledge of the thermodynamics of the system. The reaction always takes the system closer to equilibrium, the rate is zero at equilibrium and has opposite signs on each side of the equilibrium composition. The rate thus depends on the thermodynamics. However, kinetic measurements are not suitable for determination of thermodynamic parameters. Before we can begin studies of the kinetics, we must know the thermodynamics either from calculation or from measurements. Kinetic studies may be performed on several levels. We may assume something simple, such as a pseudofirst order reaction determine the rate constant. Suppose we know the reaction conditions and the desired exit composition. We can then use a small sample at these conditions and adjust the flow until the desired exit composition is reached. We then know a reliable values for F/m. We may fit the rate constant, k. We may fit the activation energy, E#, and preexponential factor, A. We may fit the activation energy, E#, the preexponential factor, A, and one or more reaction orders, i.

13.1 The simple approach


Let us assume we have a first order reaction A B with kinetics dc kc dt where c is the concentration of A. The solution is c c 0 exp kt

k c 0 exp

We then measure the rates, r, at different temperatures, T, and times t:

kt

112

T ... ... ...

t ... ... ... ... ... ...

r ... ... ... ... ... ...

We can easily rewrite Equation \ref{eq2} as ln r ln kc 0 kt For each temperature we prepare a graph t, ln r and determine the rate constant, k, by linear regression. Using the value of k at each of the temperatures, we then prepare an Arrhenius plot 1 ln k , T and determine A from the intercept, ln(A), and the activation energy, E#, from the slope, E#/R, by linear regression. You have without doubt seen this approach sketched in introductory textbooks. Most likely you have also tried applying it to your measurements and discovered that it failed miserably. Why does the method work for the "experimental" data in the textbook when it fails for your data ? The major problems are 1. If the values of r are uncertain, the values of k will be uncertain and most likely slightly wrong. When we then use these values to determine A and E#, the values will be uncertain and wrong. We have no idea how wrong and how uncertain A and E# are. 2. Calculation of the uncertainty for A and E# is quite complex. Most likely we have roughly the same errorbar for each of the concentrations, and roughly the same errorbar for each of the temperatures, but this implies different, asymmetrical "errorbars" for ln{c} and 1/T. 3. For kinetic expressions which are just a little more complicated than first oder kinetics, the rate cannot be integrated by hand. Very complex kinetic expressions are easy to treat by numerical methods, but analytical treatment is only possible for extremely simple kinetic expressions. and the minor problems are 1. In the linear regression using the integrated rate law, we actually determined k from the slope and ignored that kc0 was available from the intercept. But we know c0. We 113

thus fitted k twice in one fit and did this as if the two values were independent. In other words, we did not even do the fit properly. 2. In almost all situations we measure the inlet and outlet concentrations, together with the values of the flow, temperature and pressure, this is enough to calculate the rate. The rate itself can only be measured in exceptional cases.

13.2 A more general method


Let us consider a more complicated reaction |1| A1 + |2| A2 ... ... |n| An

13.2.1 Kinetic measurements


We then perform the kinetic measurements. We select pressures, p, flows, F, temperatures, T, and inlet concentrations x1i,...,xni We select a key component $r$ and measure the exit concentration, xre, of this component. The exit concentration of all other components may then be determined from a mass balance, the inlet concentrations and xre. It is important that we select a reference point and return to this point frequently. This should be the first and last point we measure, the first point we measure on each monday morning, the last point we measure on friday afternoon and, say, every seventh point we measure during the week. During the measurements the reference point is used to check for problems with the catalyst or with the equipment. The result should always be almost the same or we have a major experimental problem. During the analysis, the reference point is used to determine the uncertainty in the measurements. If the conversion is virtually zero for a given set of reaction conditions, it will be virtually zero at most other reaction conditions and you can fit this with A0 and any value of E# and 1,..n. Similarly, if the exit composition is near the equilibrium composition, you can fit the data with a large value for A. For this reason, kinetic measurements made at very small conversions and kinetic measurements made near equilibrium do not contain much information. The amount of catalyst and the reaction conditions must be selected to give exit compositions which are significantly different from both the input composition and the equilibrium compositions. It is frequently necessary to make several short series of kinetic measurements to determine a proper experimental procedure. When you have determined a suitable range of reaction conditions, you can perform the kinetic measurements without any prior knowledge on the values of the kinetic parameters. Just proceed with the measurements and leave the columns xrc and (xrexrc)2 114

blank in the table below. However, if you have an idea on the valued for A, E# and 1,..n use these values with the program and the reaction conditions p, F, T, x1i,...,xni to calculate an expected exit concentration xrc of the key component. If the agreement between xre and xrc is too poor you may want to check your reactor and reconsider your experimental procedures before you waste too much time. The table of experimental results may look like this: p p0 p1 T T0 T1 T2 p2 T1 T2 p0 p1 T0 T1 T2 p2 T1 T2 p0 T0 F F0 F1 F2 F1 F2 F1 F2 F1 F2 F0 F1 F2 F1 F2 F1 F2 F1 F2 F0 x10 ... xno x10 x12 ... ... xno xn2 x1i x10 x11 ... ... ... xni xno xn1 xre ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... xrc ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... (xre xrc)2 ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ...

In an experiment, the pressure is usually controlled manually, while the temperature and the flow are controlled electronically. For this reason, changes in pressure are made less frequently than changes in flow or temperature.

115

After and adjustment of the flow, the transient usually takes minutes while changes in temperature and the resulting transient frequently takes several hours. For this reason changes in temperature are made less frequently than changes in flow.

13.2.2 Thermodynamics
The equilibrium equation is
n

pi p
o

i 1

We verify that we know K, either from tabulated data in the literature or from measurement.

13.2.3 Choice of kinetic expression


We then assume some form of the kinetic expression, e.g r k

n i 1

and verify that we do not have any serious reason to believe that this form for the kinetics is wrong. It is important that we dont make the expression more complex than necessary. If we know the reaction must be first order in one of the components or zeroth order in another component, this information should be fixed a priori in the kinetic expression we will use in the following. It may be tempting to leave parameters you already know as free parameters. However, a leastsquares fit often uses extra flexibility to generate a slightly better fit and a drastic reduction of the reliability of the fitted values. What will frequently happen is the following: Suppose the true values are A=1 and B=2. You know that A is 1, but you dont know the value of B. You leave A free "just to see if it comes out right" and obtain the fit A=1.2 0.1, B=2.4 0.2$. However, if you fix A=1, you find a slightly poorer fit but a considerably better value for B, e.g B=2.1 0.1. We verify that we have a program which can treat the kinetic expression we have selected for the reactor we are going to use.

po

k K

n i 1

po

116

pi

pi

13.2.4 Fitting the parameters


If you have an idea on the valued for A, E# and 1,...,n use these values with the program and the reaction conditions p, F, T, x1i,...,xni to calculate an exit concentration xrc of the key component. (If you dont know the values for A, E# and 1,...,n, just make a wild guess to get started.) Calculate

x r e x rc

Most likely you will find that the calculated exit concentration, xrc, do not reproduce the measured exit concentration, xre. We then start the fitting procedure 1. Make a small random variation of A, E# and 1,...,n. 2. If the new parameters are unacceptable, keep the old paramets and go to step 1. 3. Use the program to update xrc and ( xrexrc)2 4. If this value is better than before, keep the new parameters else keep the old parameters. 5. Go to step 1. It may be necessary to adjust the meaning of small in step 1 as the fit improves.

13.2.5 Checking the fit


When the fit is not improved through a long series of guesses, there are two possibilities The fit is good. This is fine, but you should at least consider the possibility that the kinetic model is too general or has too many parameters. The fit is bad. Use the table to prepare plots (p,(xrexrc)2), (T,(xrexrc)2), (x1i,(xrexrc)2) etc. These plots will give valuable hints on exactly what it is in the experimental data that you kinetic model cannot fit. Dont forget to check that all your measurements of your reference points gives almost the same value. You might have a problem with the catalyst or with the reactor. Dont forget to plot (t,(xrexrc)2), where t is the number of hours from the start of the experiment until each measurement is made. This plot may send a loud message about experimental problems.

117

118

13.3 Bibliography
Catalysis Charles L. Thomas: Catalytic processses and proven catalysts. Academic Press (1970). Bruce C. Gates: Catalytic chemistry. John Wiley and Sons (1992) B. C. Gates, J. R. Katzer, and G. C. A. Schuit. Chemistry of catalytic processes. McGrawHill Book Company, 1979. Robert J. Farrauto and Calvin H. Bartholomew: Fundamentals of industrial catalytic processes. Blackie Academic and Professional, London (1998) J. A. Moulijn, P. W. N. M. van Leeuwen, and R. A. van Santen: Catalysis. An integrated approach to homogeneous, heterogeneous and industrial catalysis. Studies in Surface Science and Catalysis. volume 79. Elsevier, 1993. R. A. van Santen and J. W. Niemantsverdriet: Chemical kinetics and catalysis. Plenum (1995). Charles N. Satterfield: Heterogeneous catalysis in industrial practice. McGrawHill. 2nd edition (1991). G. A. Somorjai: Introduction to surface chemistry and catalysis. John Wiley and Sons, 1994. J. M. Thomas and W. J. Thomas: Principles and practice of heterogeneous catalysis. VCH (1997). R. J. Wijngaarden, A. Kronberg, K. R. Westerterp: Industrial Catalysis. Wiley VCH Encyclopedias J. R. Anderson and M. Boudart (editors): Catalysis: Science and technology. Springer. (19811997). 11 volumes. Volume 1 (1981) A brief history of industrial catalysis. An introduction to the theory of catalytic reactors. Catalytic activation of dinotrogen. 119

The FischerTropsch synthesis. Catalytic reforming of hydrocarbons. Volume 2 (1981) History of concepts in catalysis. Crystallography of catalysts types. Catalytic kinetics: Modelling. Texture of catalysts. Solid acid and base catalysts. Volume 3 (1982) History of coal liquefaction. Catalytic activation of dioxygen. Catalytic activation of carbon monoxide on metal surfaces. Chemisorption on nonmetallic surfaces. Chemisorption of dihydrogen. Volume 4 (1983) Catalytic processed in organic conversions. Nature and estimation of functional groups on solid surfaces. Kinetics of chemical processes on welldefined surfaces. Volume 5 (1984) Catalytic steam reforming. Automobile catalytic converters. Infrared spectroscopy in catalytic research. Xray techniques in catalysis. Volume 6 (1984) Catalyst deactivation an regeneration. Catalytic olefin polymerization. Metal catalysed skeletal reactions of hydrocarbons on metal catalysts. Dispersed metal catalysts. Volume 7 (1985) The history of the catalytic synthesis of ammonia. The electron microscope of catalysts. Surface structural chemistry. Volume 8. (1987) The historical development of catalytic oxidation processes. Catalytic methathesis of alkenes. Physicochemical aspects of mass and heat transfer in heterogeneous catalysis. Small scale laboratory reactors. EPR methods in heterogeneous catalysis. Volume 9 (1991) Determination of mechanism to heterogeneous catalysis. Dynamics relaxation methods in heterogeneous catalysis. Dynamics of heterogeneously catalyzed reactions. Volume 10 (1996) Application of NMR methods to catalysis. Glossary of terminology used to catalysis. 120

Volume 11 (1996) Hydrotreating catalysis. Science and technology. G. Ertl, H. Knzinger and J. Weitkamp (editors). Handbook of heterogeneous catalysis. WileyVCH. (1998). 5 volumes.

Experimental J. W. Niemantsverdriet: Spectroscopy in catalysis. VCH (1993). D. P. Woodruff and T. A. Delchar. Modern techniques of surface science. Cambridge University Press (1986). Historical Olaf A. Hougen and Kenneth M Watson. Chemical process principles. Part 1. Material and energy balances. John Wiley and Sons (1943) Olaf A Hougen, Kenneth M. Watson and Roland A. Ragatz. Chemical process principles Part 2: Thermodynamics. John Wiley and Sons. 1st edition (1947) 2nd edition. (1959). Olaf A Hougen and Kenneth M. Watson: Chemical process principles Part 3: Kinetics and catalysis. John Wiley and Sons (1947). H. Heinemann: A brief history of industrial catalysis. in John R. Anderson and Michel Boudart (editors): Catalysis. Science and technology. Volume 1. Springer (1981). M. Boudart. Kinetics of chemical reactions. Prentice Hall, 1968. Introductions Gary Attard and Colin Barnes: Surfaces. Oxford Chemistry Primers, number 59 (1998). Ian S. Metcalfe: Chemical Reaction engineering. A first course. Oxford Chemistry Primers, number 49. Oxford University Press (1997). Michael Bowker: The basis and applications of heterogeneous catalysis. Oxford Chemistry Primers. Volume 53. Oxford University Press (1998). 121

Kinetics Sidney W. V. Benson: Thermochemical kinetics. John Wiley and Sons (1968). Michel Boudart and G. DjegaMariadassou: Kinetics of heterogeneous catalytic reactions. Princeton University Press (1984). James A. Dumesic, Dale F. Rudd, Luis M. Aparicio, James E. Rekoske, Andres A. Trivino: The microkinetics of heterogeneous catalysis. American Chemical Society (1993) Reaction engineering. J. J. Carberry. Chemical and catalytic reaction engineering. McGrawHill, 1976. G. F. Froment and K. B. Bischoff. Chemical reactor analysis and design. John Wiley and Sons, 2nd edition, 1990. Octave Levenspiel: Chemical reaction engineering. John Wiley and Sons 2nd edition (1972) . L. D. Schmidt. The engineering of chemical reactions. Oxford University Press, 1998. H. Scott Fogler. Elements of chemical reaction engineering. PrenticeHall, 2nd edition, 1992. Theory Stephen Elliot: The physics and chemistry of solids. John Wiley and sons. (1998). R. I. Masel: Principle of adsorption and reaction on solid surfaces. John Wiley and Sons, 1996. R. A. van Santen and J. W. Niemantsverdriet: Chemical kinetics and catalysis. Plenum (1995). R. A. van Santen: Theoretical heterogeneous catalysis. World Scientific (1991). Andrew Zangwill: Physics at surfaces. Cambridge University Press (1988)

122

V. P. Zhdanov. Elementary physicochemical processes on solid surfaces. Plenum, 1991

123

S-ar putea să vă placă și