Lecture Notes Catalysis Engineering
Lecture Notes Catalysis Engineering
Industrial Catalysis CPT Prof. Dr. J.A. Moulijn Prof. Dr. F. Kapteijn Drs. A.E. van Diepen, MTD Dr. Ir. M.T. Kreutzer
March, 2000
Preface
In the framework of the new curriculum we have introduced the lecture course Catalysis Engineering. The purpose of the course is to provide students with tools that enable the optimal design of a catalytic process. Various disciplines are integrated, and students are assumed to have knowledge of reactor technology and transport phenomena. The most recent developments in the area of catalysis and reactor technology are presented. Since no textbook covering this area is available at this time and the intention of this course is unique, the lectures will be based on texts of our own and journal articles. During the course, supplements to the first hand-outs might follow.
Contents
CONTENTS
1 Introduction 1.1 Reactors for solid-catalyzed reactions 1.2 Reactor design equations basic definitions Solid Catalysts 2.1 Introduction 2.2 Catalyst shape and size in relation to reactor type 2.3 Classification of solid catalysts 2.4 Preparation of bulk metal catalysts 2.5 Catalyst supports 2.6 Preparation of supported catalysts 2.7 Economical aspects Catalytic Reaction Kinetics 3.1 Introduction 3.2 Rate expression (single-site model) 3.3 Rate-determining step quasi equilibrium 3.4 Adsorption isotherms 3.5 Rate expressions (other models and generalizations) 3.6 Limiting cases reactant and product concentrations 3.7 Temperature and pressure dependence 3.8 Sabatier principle Volcano plot 3.9 Concluding remarks Case Studies I N2O decomposition II Complex kinetics III Kinetic coupling IV CFC conversion 4 Mass and HeatTransport Catalyst Effectiveness 4.1 Introduction 4.2 Extraparticle gradients 4.3 Intraparticle gradients 4.4 Comparison of criteria 4.5 Effect of particle transport limitations on behaviour Appendix 4.I
Contents 5 Catalyst Deactivation 5.1 Introduction 5.2 Catalyst deactivation qualitative description 5.3 Poisoning and fouling quantitative description Selectivity 6.1 Introduction 6.2 Reaction selectivity 6.3 Shape-selectivity 6.4 Catalyst modification for improved selectivity 6.5 Effect of catalyst deactivation on selectivity Strategies for Multiphase Reactor Selection [Krishna and Sie, 1994] 7.1 Introduction 7.2 Strategy level I catalyst design 7.3 Strategy level II injection and dispersion strategies 7.4 Strategy level III choice of hydrodynamic flow regimes 7.5 Case study of oil shale reactor selection 7.6 Closing remarks Special Topics The Design and Application of Porous Catalysts Special Topics Monolith Reactors
8 9
Introduction
INTRODUCTION
Every industrial chemical process, and catalytic processes are no exception, is designed to produce economically and safely a desired product or range of products from a variety of feedstocks. Ideally, one would like to produce safely and at the lowest costs, at a high throughput and the highest yield (100% selectivity and conversion), with minimum energy consumption and no negative environmental impact (see Figure 1.1). In the selection and design of a catalytic conversion process to approach this ideal several factors play a role, which often impose conflicting demands on the process. For instance, minimum costs of the overall process are aimed at, but safety and environmental requirements will add to the costs. In order to achieve an economically viable process operation a maximum selectivity of the desired product is striven for with, at the same time, a high conversion level of the reactants and a high throughput. Often, however, the selectivity decreases with increasing conversion, so a fine balance must be sought to optimize the yield of the product and to minimize or avoid completely the production of undesired by-products. There is a growing incentive for so-called zero emission plants. Energy consumption, having an impact on the operational costs and the environment, should be minimized by careful integration of the various unit operations. With respect to the catalytic reactor, integration relates to factors such as the required heat exchange duty, stirring power, recirculation rates, and pressure drop. Figure 1.1 depicts the various aspects that are related with the choice of a reactor system for catalytic operations. This forms the basic question of this lecture course: What engineering tools are presently available to enable the design of an optimal catalytic reactor and achieve optimal reactor operation.
Economics
Reactants
Reactor
Process requirements Maximum selectivity maximum conversion ease of scale-up high throughput low pressure drop .
1-1
Introduction This lecture course will focus on the engineering aspects of catalysts like structure/morphology, mass and heat transport, effectiveness, kinetics, selectivity, poisoning, and deactivation, and how they interact and can be used optimally. These basic elements, will be integrated in a strategy to optimize the choice of a catalyst and reactor, especially for three-phase operation. Examples of new and emerging catalytic processes will illustrate the usefulness of this integrated approach.
Figure 1.2 Adiabatic fixed-bed reactor; single fluid phase. Advantages of the adiabatic fixed-bed reactor are its simplicity, the high catalyst load per unit volume, little catalyst attrition, and little backmixing. Disadvantages are the high pressure drop, difficult temperature control, and the long diffusion distances. 1-2
Introduction
A reactor system similar to the fixed-bed reactor is the moving-bed reactor where the deactivation rate is relatively low, but too high for pure fixed-bed operation, e.g., in some catalytic reforming processes. In hydrodemetallization (HDM) a moving bed is also applied (Hycon process): the catalyst destroys the metal containing structures (porphyrins) and becomes slowly filled with metal sulfide deposits. Catalyst particles should be spherical and of uniform size for smooth flow through the bed. The fluid-bed reactor (Figure 1.3) is mainly chosen because of its good heat exchange properties and the uniform temperature distribution. Furthermore, due to the fluid-like behaviour of the bed, the catalyst can be added and withdrawn during operation for regeneration or replacement in case of fast deactivation. In addition, due to the relatively small particle size, the diffusion distances are short. On the other hand, disadvantages are the occurrence of catalyst attrition and backmixing, and the fact that fluid-bed reactors are difficult to scale-up.
Regeneration
The entrained-flow reactor (Figure 1.4) is used when very short contact times are required, so in case of highly active catalysts that deactivate fast. In fluid catalytic cracking (FCC) the recirculating catalyst also supplies part of the heat for the endothermal reaction. Depending on the catalyst loading one distinguishes dilute and dense phase risers. Advantages of entrained-flow reactors are the high mass-transfer rates, short contact times, and the possibility of continuous catalyst replacement. Disadvantages are the occurrence of catalyst attrition, reactor erosion, and the requirement to separate the catalyst from the product. Reactor types intermediate between fluid bed and entrained flow also exist, like internally circulating bed reactors with an internal riser section and an external annular section of a fluid/moving-bed type. Relatively new are the monolithic reactor types (shown in Figure 1.8 for gas/liquid/solid operation) that are mainly used in environmental applications [2]. Examples include the three-way catalyst (TWC) for exhaust gas cleaning of gasoline engines, the selective catalytic reduction catalyst (SCR) for NOx reduction with ammonia in lean-burn combustion flue gases, the oxidation of VOCs and the 1-3
Introduction
decomposition of ozone for air conditioning (airplanes, photocopiers). All have an extremely low pressure drop due to the structured character of the catalyst and can be used at high flowrates. The TWC has small channels (~1 mm) with very thin walls and is highly thermostable. The SCR catalyst has relatively large channels (~1 cm) and is tolerant for operation under dusty conditions. Another example of a structured reactor with low pressure drop is the parallel-passage reactor for flue gas treatment, which is shown in Figure 1.5.
Flue gas
Figure 1.5 Parallel-passage reactor. Liquid-solid reactors of the fluid bed or slurry type do exist but are much less encountered than their gas-solid analogues. In most cases a gas phase is involved, too, resulting in a three-phase operation. Many applications of gas/liquid/solid reactors are found in practice. Selective oxidations and hydrogenations, hydrotreating, hydrations and aminations of a liquid reactant (mixture) catalyzed by a solid catalyst are important industrial examples. Two extremes exist in three-phase operation: the trickle-bed reactor (Figure 1.6) and the slurry reactor (Figure 1.7).
dry,
unwetted, wetted
Figure 1.6 Trickle-bed reactor. In the trickle-bed reactor the gaseous and liquid reactants flow co-currently downwards over a packed bed of catalyst. The liquid phase must be distributed such that an even wetting of the catalyst is obtained. Uneven wetting may lead to local hot spots and runaways. The gas phase is the continuous phase. High gas flow rates are used to remove heat. Since the liquid covers the catalyst particles, 1-4
Introduction
reactant diffusion is much slower than in gas phase operation. The nature of the packed bed requires catalyst bodies of a few millimeters, so the catalyst effectiveness is restricted. Advantages of trickle-bed reactors are high catalyst load, no need for catalyst separation, no attrition, and very limited backmixing. Disadvantages are reduced catalyst effectiveness and selectivity, cocurrent operation, flow maldistributions, and large pressure drop. Upflow reactors are sometimes used for small scale testing. Here, the liquid phase is the continuous phase through which the gas bubbles rise and wetting problems are absent. Countercurrent operation may have certain advantages like removal of intermediate products to avoid side reactions, to overcome equilibrium limitations (hydrogenation of aromatics) or to avoid product inhibition (hydrodesulfurization). Also, temperature profiles along the reactor length differ which may be used advantageously. The major problem is flooding, i.e., entrainment of the liquid by the gas phase. Larger catalyst particles must be used to allow sufficiently high flow rates, thereby further lowering the catalyst efficiency. In slurry reactors (Figure 1.7) small catalyst particles (10-100 m) and gas bubbles are used suspended in a liquid phase by mechanical mixing. These reactors can be operated in a batch, semibatch or continuous mode. In the semi-batch mode often the gas-phase is supplied continuously.
1-5
Introduction
a high degree of backmixing. Separating slurry catalysts from the reaction mixture requires filtration, which is very energy-intensive. In bubble-column reactors, in order to reduce attrition, recirculation of the particles can be achieved by the gas bubbles (gas lift) instead of by mechanical agitation. External liquid recirculation offers the possibility to improve the gas-liquid mass transfer by special liquid ejector types. In case of gas and liquid upflow a three-phase fluidized bed reactor is produced. Here particles of a few millimeters are needed to retain them in the reactor and allow separation from the gas/liquid phase. Since three-phase operation is industrially very important and since this is the most challenging catalytic operation, Table 1.1 presents a comparison of the main characteristics of three-phase reactor types, while Table 1.2 gives a number of appreciation criteria. These can serve in considerations for choosing a proper reactor type for a given reaction. Several aspects will be further elaborated on later. Table 1.1 Characteristics of three-phase reactor types (after [3]). Catalyst in Fixed bed Characteristics suspension Bubble Stirred Cocurrent Cocurrent column tank down up 0.01-0.1 0.01-0.1 0.6-0.7 0.6-0.7 p 0.8-0.9 0.8-0.9 0.05-0.25 0.2-0.3 L 0.1-0.2 0.1-0.2 0.2-0.35 0.05-0.1 G dp (mm) 1-5 1-5 0.1 0.1 aS 500 500 1000-2000 1000-2000 2 -3 (m cat m reactor) 100-400 100-1500 100-1000 100-1000 aGL 2 -3 (m gas bubble m reactor) 1 1 (isoth.) <1 <1
3-Phase fluid bed Countercurrent 0.5 0.05-0.1 0.2-0.4 >5 500 100-500 <1
1-6
Introduction
Table 1.2 Comparison of different three-phase reactor types (after [3]). Appreciation criteria Suspended catalyst Three-phase fluid bed
Activity Highly variable, but often full utilization
Fixed bed
Selectivity Stability
Scaling-up
Highly variable: inter and intra mass transfer may reduce utilization, especially in fixed bed backmixing plug flow + Generally unaffected by as above mass / heat transfer backmixing often plug flow often + Essential for fixed bed Possibility of Catalyst replacement operation. Plug flow may continuous renewal. between batches to establish a poison adsorption Must have good overcome rapid front attrition resistance poisoning Consumption depends on feed impurities acting as Necessarily low catalyst poisons consumption Fairly easy to achieve Possibility of heat Generally adiabatic operation exchange in reactor Catalyst separation sometimes difficult: possible Fairly simple for a downward problems in pumps due to risks of deposits or cocurrent adiabatic bed erosion No difficulty: generally System poorly known, Large reactors can be built limited to batch and should be scaled up in provided a good liquid relatively small sizes steps distribution is achieved
New reactor types may be developed, like the application of structured reactors, which may have certain advantages in three-phase operation and can be operated both in co-, cross- as well as in countercurrent mode. A very recent development is the use of monolith reactors (Figure 1.8) in threephase operations. Their advantages are the low pressure drop, the good external mass transfer, the short diffusion distance, and the low adiabatic temperature rise. Disadvantages are the higher catalyst costs, importance of liquid distribution, and moderate catalyst load.
G, L, washcoat, substrate
Figure 1.8 Monolith reactor. Three-phase monoliths are not yet applied industrially. Only limited experience exists on laboratory and pilot-plant scale. 1-7
Introduction
r,V
Batch
t = c A0
XA
dX r A
F A0
r(X,z)
FA
Plug flow
0 =
XA V c A0 dX = c A0 FA0 r A 0
F A0 r,V
FA
CSTR
0 =
V c A0 X = c A0 A Ar FA0
Figure 1.9 Design equations for isothermal ideal reactors. The equations relate the concentration cA0 and molar flow rate FA0 of a component A in the feed to the reactor volume V, the reaction kinetics in terms of the rate expression r, the stoichiometric number A (positive for products, negative for reactants), and the conversion at the reactor exit or the batch operation time. The used reaction rate is a specific quantity having the dimensions of mol per second per unit reactor volume. It is often convenient to define a rate per unit volume of catalyst or even per unit catalyst mass rather than per unit of reactor volume. Therefore, various subscripts are introduced. The relations between the various specific rate definitions are given in equation (1) .
V rV = V p rv = W rw
[mol/s]
(1)
The production (or consumption) rate of a species per unit reactor volume is then given by equation (2).
RV , A = A rV
[mol A/s]
(2)
In catalytic reactions mass transfer from the fluid phase to the active phase inside the porous catalyst particle takes place via transport through a fictitious stagnant fluid film surrounding the particle and
1-8
Introduction
via diffusion inside the particle. Heat transport to or from the catalyst takes the same route. These phenomena are summarized in Figure 1.10.
PLUG FLOW MIXING DISPERSION
Figure 1.10 Phenomena in a packed-bed reactor. Transport resistances may lead to an under- of overestimation of the local conditions in the catalyst particle, so to a deviating reaction rate estimation. To correct for this the catalyst effectiveness factor is introduced in the design equation. This effectiveness factor may also include corrections for mass transport from a gas phase to a liquid phase in three-phase operation. Note that generally is a local variable depending on the local conditions in the reactor. Taking a plug-flow reactor under steadystate conditions as an example, equation (3) is obtained as the design equation in differential form.
dX A = A rV d (V FA0 )
(3)
Since catalysts are subject to deactivation by feed impurities, reaction products, or other unwanted phenomena, a time-on-stream dependent activity function is contained in the factor . This is a function of time-on-stream and local conditions and generally cannot be separated as is done in equation (3). In more general cases a set of differential equations (ODEs) is obtained for the plugflow reactor description, coupled with the set of differential equations describing the local catalyst behaviour in the reactor. The effect on reaction selectivity is taken into account simultaneously, since the set of ODEs is set up for all components and, in principle, includes all reactions. For a complete treatment the momentum and enthalpy balances should be included, too. The forthcoming chapters will focus in more detail on various scale-independent elements that are contained implicitly in equation (3), like the description of the catalyst characteristics, catalytic reaction kinetics, mass and heat transfer, and catalyst deactivation, and their impact on the reaction selectivity. These elements, combined with the basic principles of reactor engineering, provide the basis of an integral approach to optimally select and design catalytic reaction systems.
Notation
aGL gas/liquid interfacial area m2gas bubble m-3r
1-9
Introduction aS c dp Fi0 rv rV rW RV,A t V Vp W X external surface area of catalyst particles per unit volume concentration particle diameter molar flow of i at reactor inlet reaction rate per unit particle volume reaction rate per unit reactor volume reaction rate per unit catalyst mass production/consumption rate of A per unit reactor volume time reactor volume particle volume catalyst mass conversion
Notation & Literature m2cat m-3r mol m-3 m mol s-1 mol s-1 mp-3 mol s-1 mr-3 mol s-1 kg-1 mol s-1 mr-3 s m3 m3 kg -
Greek
G L p
volume fraction of reactor occupied by gas volume fraction of reactor occupied by liquid volume fraction of reactor occupied by catalyst paraticles overall effectiveness factor stoichiometric coefficient of i in reaction residence time based on feed conditions deactivation function
i 0
Literature
1. R. Krishna and S.T. Sie, Strategies for multiphase reactor selection Chem. Eng. Sci. 49(24A), 4029 (1994). 2. A. Cybulski and J.A. Moulijn (Eds.), Structured catalysts and reactors, Marcel Dekker, New York, 1998. 3. P. Trambouze, H. Van Landeghem and J.P. Wauquier, Chemical Reactors, ditions Technip, Paris, 1988.
1 - 10
Solid Catalysts
Introduction
SOLID CATALYSTS
2.1 Introduction
Solid catalysts are extensively used in oil refining and in the chemical industry. Their main advantages compared to soluble, homogeneous catalysts are their ease of separation and regeneration. The most important properties of a catalyst are activity, selectivity, and stability. In solid catalysts these are provided by two main components: Active catalyst phase: metals, oxides, sulfides, etc.; Support: provides particle size, shape, porosity, surface area, strength, etc.
Sometimes one material fulfills both functions. Examples are alumina catalysts and zeolites. Often so-called promoters are added in small quantities to enhance selectivity, activity and stability.
Irregular granule
Sphere
Pellet
Extruded Cylindrical
Trilobe
Ring
Minilith
Wagonwheel
Monolith ceramic
Monolith metallic
Foam
Figure 2.1 Examples of industrial catalyst shapes. The size and shape of catalyst particles are a compromise between conflicting demands: 2-1
Solid Catalysts
minimum pressure drop; minimum pore diffusion resistance = maximum catalyst utilization; maximum mechanical strength; minimum cost. Arrange the catalyst shapes of Figure 2.1 from best to worst with respect to the demands postulated above.
QUESTION:
Catalysts for fixed-bed reactors, are relatively large particles (typically several mm in diameter) in order to avoid excessive pressure drops. However, in large particles the diffusion distances are long, causing larger pore diffusion resistance. Mechanical strength should be sufficiently high to avoid formation of fines which causes increased pressure drop and eventually plugging of the reactor. The same requirements apply to catalysts for moving-bed reactors. An additional demand is that the catalyst should flow smoothly, so spherical particles of uniform size are recommended for this reactor type. Monolithic catalysts are increasingly used in fixed-bed reactors for processes requiring low pressure drop. The most familiar example is automotive exhaust gas cleaning. Monoliths typically have lengths of 2 100 cm with channels of 1 10 mm. In slurry reactors powdered catalysts are used with a diameter of typically 10 - 100 m. Mechanical strength is very important, because attrition can occur leading to decreased filterability. In many applications, in particular in fine chemicals production, the catalyst is separated from the reaction mixture by settling. Therefore, a high catalyst density is favourable. Still, often a post-filtration step is required. Catalysts for fluidized-bed and entrained-flow reactors are powders with a diameter of 20 to 100 m. Fluidization requires a well controlled particle size distribution. Again, mechanical strength is an issue, because catalyst fines formed by attrition have to be separated from the product and constitute a catalyst loss.
Solid Catalysts
Classification
Catalysts can be classified based on their chemical compositions. Table 2.1 gives a survey of some active catalyst components. The major groups are the metals, oxides/sulfides, and the aluminasilicates. Table 2.1 Composition of some solid catalysts. Catalyst Typical application Metals and alloys Ammonia synthesis Fe Hydrogenation Raney-Ni Catalytic reforming Pt/Al2O3, Pt-Re/Al2O3 Hydrogenation Pd/C, Pt/C Metal Oxides Partial oxidation of butane to maleic anhydride VPO4/SiO2 Selective catalytic reduction of NO by NH3 V2O5/TiO2 High temperature Water-gas shift Fe3O4 /Al2O3 Methanol synthesis, low temperature Water-gas shift CuO-ZnO/Al2O3 Solid acids Oxidation of SO2 SiO2 (silica) Alcohol dehydration Al2O3 (alumina) Old FCC process, precracking for FCC Amorphous silica-alumina (ASA) Current FCC process, xylene isomerization Zeolites Sulfides Hydrotreating CoS-MoS2/Al2O3 Hydrotreating NiS-MoS2/Al2O3 Chlorides Oxychlorination CuCl2/Al2O3 Polyethene (high density) TiCl4
Metal catalysts are typically used in hydrogenations, but also in hydrogenolysis, reduction and oxidation. Metal oxides have amongst others applications in selective oxidations and reductions. Metal sulfides are commonly used in hydrotreating, which requires hydrogenation and hydrogenolysis reactions. The solid acids are very suitable for reactions requiring the formation of carbenium ions, such as catalytic cracking. Often, catalysts are bi-functional. For instance, the catalyst used in catalytic reforming contains a hydrogenation/dehydrogenation catalyst (Pt), and an acid catalyst (Al2O3) which enhances isomerization.
2-3
Solid Catalysts
The preparation of bulk metal catalysts starts with fusing (melting) of a mixture of the components. The surface area of this fusion product is very low, so porosity must be created, for instance, by removing some component from the fuse.
Solid Catalysts
Catalyst supports
catalyst sites. Most supports have high specific surface areas, ranging from 100 to 700 m2/g. Examples are silica, alumina, silica-alumina, titania, zirconia, carbon, zeolites, and other molecular sieve type materials.
2-5
Solid Catalysts
Catalyst supports
Mean pore size (nm) 3.5 bimodal (5-50 and > 50) 10 9 1 (often bimodal) 0.4 1
The inorganic oxides dissolve in acidic or basic media. Therefore, activated carbon is used in these cases. In addition, the acidity of supports such as silica and alumina is not always desired (e.g., cracking activity) and more inert materials are used like TiO2, ZrO2, or activated carbon. Furthermore, silica can not be used under conditions of high temperature and steam atmosphere, because then evaporation takes place. Silica, Alumina, and Silica-Alumina Silica and alumina are both prepared by a method that starts with precipitation from a concentrated solution of the metal salt (see for example [1]). The resulting hydrogel, after purification can be formulated into various shapes, such as powders, pellets, extrudates, or spheres. Silica-alumina is a mixed oxide that can be produced by 'impregnation' of porous silica with an Al3+ solution. Silica-alumina contains Lewis and Brnsted acid sites. It is a remarkably strong acid, much more acidic than silica or even alumina alone. This is explained by the fact that in deprotonation of silica-alumina the negative charge is spread over the AlO4-unit as shown in Figure 2.2. Silica has no catalytic activity of its own. It is used as a support for Ni, Pd, or Pt in gas-phase hydrogenation reactions, and also as a support for V2O5 in the oxidation of SO2. -Alumina is used as support in, amongst others, catalysts for exhaust-gas treatment, hydrotreating catalysts, and some hydrogenation catalysts. -Alumina has also catalytic activity itself, and is used as catalyst for alcohol dehydration and in Claus plants for the conversion of H2S into elemental sulfur. Sometimes, a catalyst with low surface area is desired, for example in partial oxidation catalysts. alumina is suitable for this type of applications. Another example is steam reforming on Ni/-Al2O3.
2-6
Solid Catalysts
Catalyst supports
Zeolites (see Section 2.5.3) are a special type of silica-alumina compounds. They are crystalline and have a well-defined pore-structure with pores of molecular dimensions.
silica: Si O Si O Si OH O Si silica-alumina: Si O Si O Al HO Si O Si Si O O Si Si O O Si Si O Si O
H+
+ H+
Si O Al O Si
Activated Carbons Activated carbons are carbons with high specific surface area. Carbon is extensively used in industrial adsorption applications, but also in catalysis, because of its unique properties: stability in acidic and basic solutions; high adsorption capacity for organic molecules; wide variety of texture properties; resistance to high temperatures; not much erosion of pumps, etc.; no reduction of activity of active phase.
Disadvantages of activated carbons are their reactivity in oxidizing environments, their sometimes weak mechanical properties, and the formation of fines, which can make separation difficult. Activated carbons can be prepared from a wide variety of precursor materials such as wood, coal, lignite, nut-shells, sugar, and polymers. Usually, the manufacturing process consists of pyrolysis followed by activation (partial gasification) to increase the porosity and surface area. Table 2.3 gives a survey of industrial applications of activated carbon in catalysis. Table 2.3 Industrially applied reactions on carbon-supported catalysts. Reaction Active phase Zn(OCOCH3)2 Acetylene + acetic acid to vinyl acetate Acetylene + HCl to vinyl chloride HgCl2 Hydrogenations Pt, Pd, Rh, Ni Oxychlorination CuCl2 Selective catalytic reduction C 2-7
Solid Catalysts
Catalyst supports
2.5.3 Zeolites
2.5.3.1 Introduction
In the field of heterogeneous catalysis the zeolites, as crystalline, microporous, Brnsted acid catalysts (see Figure 2.3), belong to the most superb catalysts.
1-,2-,3-D pore configuration throughout the crystal pore length of crystal size pore diameters of few ngstrom
crystals of few m
Figure 2.3 Main features of the aluminosilicates denoted as zeolites: the Brnsted acidity, the micropore availability and the framework crystallinity. The more than 100 different structures are well refined on an atomic scale. Thus bond length and angles of the atoms comprising the catalyst material are known in detail. Depending upon the type of zeolite the microporosity shows a sharp, almost unique pore-size distribution which is in the range of molecular dimensions. This enables molecular shape selectivity. Based on chemical and physical compatibility of alumina and silica on an atomic scale, relatively strong Brnsted acidity is achieved in the so-called tectosilicate by isomorphous substitution of aluminum that is counterbalanced by protons. The catalytically active sites induced by a postsynthesis procedure of the zeolite are accessible for reactant molecules since all the atoms in the zeolite framework are exposed to the pore volume Many reactions in the oil refining and the petrochemical industry are accompanied by the formation of carbonaceous by-products (coke, also see Chapter 5) that deactivate the catalyst. Zeolites are thermally stable and can, therefore, be regenerated at high temperature by oxidation of these carbonaceous deposits. For competitive industrial application of zeolites, these high-tech catalysts demand time and massive effort in research and development. However, after elucidation of the best matching between the type
2-8
Solid Catalysts
Catalyst supports
of zeolite catalyst and specific reaction the high expectations of these excellent materials can be fulfilled. The contribution of zeolites in the industry has developed recently from modest to substantial with a bright outlook to the near future. For example, Shell Pernis today uses zeolites in all the processes to convert crude oil into products. Table 2.4 shows these processes, while some are treated in more detail in Chapter 8. Table 2.4 Use of zeolite in a modern oil refinery. Oil fraction Catalytic process paraffins aromatization olefins methanol gas oligomerization alkylation light ends naphtha isomerization hydrotreating reforming hydrotreating dewaxing dewaxing catalytic cracking hydrocracking hydrodesulfurization hydroconversion
Zeolite + + + + + + + + + + + + + + +
middle distillates
residue
Another example of a modern zeolite application is the production of 400.000 ton/annum of ethylbenzene at Shell Moerdijk with H-ZSM-5 zeolite, see Chapter 8. Almost no waste and no byproducts are formed compared to the conventional process that uses substantial amounts of hazardous catalysts that are not selective and burden the environment. The preparation and detailed studies on characterization and applications of zeolites are discussed in references [2] and [3]. The following sections present the relevant structures and pore configurations, from which the particular zeolite properties are derived. The industrial applications, based on these properties, are discussed in Chapter 8.
2-9
Solid Catalysts
Catalyst supports
The frameworks of zeolites are built up from SiO4 and AlO4- tetrahedra, as shown in Figure 2.4a as ball-and-stick models. By only drawing lines between the Si and Al positions and omitting the oxygen positions the structures can be displayed in the most simple way depicted in Figure 2.4b and Figure 2.5. a) b)
Zeolite A
O Al or Si
Zeolite Y
Figure 2.4 a) Ball and stick model; b) Connections between Si and Al. Oxygen positions are omitted for clarity.
A (LTA)
Beta
Y (Fau)
Figure 2.5 3-D configurations of zeolite pores of zeolite A, Beta, and Y. The presence of Al in the structure results in a net negative charge associated with each AlO4 tetrahedron. This charge is compensated by the presence of potentially catalytic sites such as metal cations or protons that are situated in the pores. In particular with a high Si/Al ratio the acid sites are considered to be strong, as shown in Figure 2.6. Even with ppm amounts of Al in the framework hexane cracking activity is observed. The Si/Al ratio in zeolite structures can vary from 1 to infinity.
2 - 10
Solid Catalysts
relative catalytic activity
Catalyst supports
100 * 10 * 1 * * * 10
-1
* *
** * *
10
100
1000
10000
ppm Al
Figure 2.6 Relative catalytic activity of H-ZSM-5 versus Al (ppm) in hexane cracking.
QUESTION:
The different connecting patterns of tetrahedra lead to different zeolite structures. A great variety in 1, 2-, and 3- dimensional pore configurations, shaped as interconnected cavities or as channels with a scala of individual pore diameters, is possible. These are controlled by the number of oxygen atoms, as shown in Table 2.5 and depicted in Figure 2.7 for a few zeolites frequently used in industry. The Si/Al ratio, in combination with the structure and pore configuration, determines the overall properties of these molecular shape selective Brnsted catalysts (see Table 2.6 in Section 2.5.3.3).
Table 2.5 Physical data of some frequently used zeolites. Type Si/Al n oxygen aperture pore volume (pore ring) [nm] [ml/g] A 1 8 .45 .30
Y >2.5 12 12 12 10 .75 .65 x .70 .64 x .76 .55 .35 .20 .25 .15
Ultimately, bifunctional catalyst systems consisting of H+ and Pt can be achieved by synthesis inside the zeolite pores. In the case of hydro-isomerization of C5 and C6 alkanes (see Chapter 8), refineries use so-called H-mordenite with Pt dispersed in the pores.
2 - 11
Solid Catalysts
Catalyst supports
Finally, regarding the pore system of zeolites, as shown in Figure 2.7, the micropore surface of the 3D pore configuration of the cubically shaped zeolite A is most accessible, while the 2-D pore configuration in zeolite ZSM-5 is somewhat less accessible. The pore volume of zeolite Mordenite has a reduced accessibility as it is a 1-D pore configuration with the pores running parallel to the fiber direction.
ZSM-5
Mordenite
Figure 2.7 Relationship between crystal forms and pore configurations. The micropore volume in zeolite A is far more accessible than in Mordenite. Arrows represent the directions in which molecules can move through the pores.
In view of the trend in (petrochemical) catalysis to shorten contact times, the reactions on the catalyst will proceed under transient rather than steady-state conditions, requiring that the penetration time of molecules in pore systems is as short as possible. Therefore, accessibility of the catalytic sites or maximal exposure of these sites to the reactant is important. This means that the number of pore entrances relative to the pore lengths, often determined by the crystal form and size, must be as high as possible to optimize the accessibility of the pore surface and thus the catalyst sites. As the pore diameters of the zeolite are close to the kinetic diameters of guest molecules, separation on a molecular scale is possible. For example, the separation of n-alkanes and branched alkanes is possible over zeolite A (0.45 nm pore diameter). The diffusivity of the molecules through the zeolite pores is determined beyond the so-called molecular and Knudsen regions in the so-called configurational region. In this region, where the molecules diffuse through pores of almost the same dimensions, small differences in size result in dramatic differences in diffusion coefficients as shown in Figure 2.8.
2 - 12
Solid Catalysts
D [cm2/sec] Molecular 1 10 10
-2
Catalyst supports
-4
Knudsen
.1
1000
Figure 2.8 Dependence of the magnitude of diffusivity of molecules upon the zeolite pore size between 0.5 and 0.8 nm (the zeolite pore windows) compared to Knudsen and molecular diffusion.
For reasons of mass transport through the pores it is important that the pores are as short as possible. Although the size of zeolite crystallites is in the order of a few micrometers, which seems small, this is still far too large in terms of mass transport. For example, imagine a substituted aromatic molecule with a length of 1 nm migrating through a crystallite of 3000 nm. This is comparable to a truck of 10 m driving through a very narrow tunnel of 30 km, with bends, obstacles, potholes and sites that interact with the van! QUESTION: A catalyst particle of zeolite is exposed to a reactant flow. How far does the reactant molecule penetrate in a pore system of a zeolite crystal?
A practical solution to improved accessibility of the catalytic sides in the pores and to decreased total residence time of molecules in the pores is to prepare very small crystallites, preferably in the order of some tens of nanometers. As a recognized approach and accepted strategy this is an important line of research today.
2.5.3.3 Properties
Table 2.6 gives the trends in particular properties of zeolites as a function of the Si/Al ratio. With large Al content in the framework a relatively large number of cations is present in the pores. This results in a rather hydrophilic pore behavior compared to a framework with a small number of Al or without any Al. Figure 2.9 shows a comparison between the adsorption equilibrium isotherms for
2 - 13
Solid Catalysts
Catalyst supports
water and n-hexane on the pore surface of NaX and Silicalite-1 (no Al, hence all-silica framework). It is well documented that so-called all-silica materials are strongly hydrophobic. These types of materials are, however, not applicable for catalysis but only for separation. Upon decreasing the Si/Al ratio more catalytic sites are available. However, with increased number of Brnsted sites the strength in acidity drops. The optimum in catalytic performance of Brnsted catalysts is thus a distinct number of sites. For example, for zeolite H-ZSM-5 the maximum Al content as expressed in the minimum Si/Al ratio is 8. For optimal performance in acid catalysis a Si/Al ratio of 24 is preferred.
Si/Al
hydrophilic hydrophobic n Brnsted sites acid strength thermal stability resistance to acid
900 K 1400 K
0.4 O 0.3 O O
O H2O
NaX
(Si/Al = 1.5) n-hexane n-hexane
0.2
silicalite-1
0.1 H2O 0.2 0.4 0.6 0.8 1.0 p/p0 (all silica)
Figure 2.9 Comparison of adsorption isotherms for water and n-hexane in silicalite-1 (hydrophobic) and NaX (hydrophilic).
2 - 14
Solid Catalysts
Catalyst supports
It is clearly observed that a polar molecule like water adsorbs preferentially in NaX compared to the a-polar n-hexane and the other way around in silicalite-1. QUESTIONs: Adsorption of guest molecules in zeolite pores is often based on three phenomena: 1) kin.diameter of guest molecule < diameter pore window 2) Cguest outside the pores > Cguest inside the pores 3) interaction with pore wall Based on (3). Why is the n-hexane molecule adsorbed more in NaX than in silicalite1?
2.6.1 Precipitation
An example of the use of the precipitation method is the manufacture of the CuO-ZnO/Al2O3 catalyst of ICI for methanol synthesis. This catalyst is manufactured by co-precipitation starting from a concentrated solution of Cu, Zn, and Al nitrates. By choosing the right conditions (pH, concentrations of the nitrates, residence time, and temperature), the desired co-precipitate is formed in a finely divided form of high surface area. After filtration and washing the precipitate undergoes calcination (treatment at high temperature in air) to yield the oxides. Precipitation yields a catalyst in which the components are well mixed on an atomic scale, whereas this is not the case with impregnation. It is usually easier to achieve a high loading of active phase by precipitation than by impregnation, and at these high loadings high dispersion can be achieved. However, at low loading the dispersion is relatively low. Another disadvantage of precipitation is that scale-up is difficult (it involves crystallization, one of the least understood chemical processes). Furthermore, the precipation technique is not suitable for use with noble metals, since always some of the metal is lost in the support phase. QUESTION: Why is the problem of the disappearance of the metal in the support phase greater with noble-metal catalysts than with other metal catalysts?
2.6.2 Impregnation
Preparation of catalysts starting with preformed supports, which are commercially available, is attractive, because a support with optimal properties, in particular its porosity structure, can be selected. Two methods can be used to add the active phase, i.e., dry impregnation and wet impregnation. Dry 2 - 15
Solid Catalysts
Supported catalysts
impregnation is also referred to as pore volume impregnation, because with this method the amount of solution containing the catalyst precursor (often a salt) used is just enough to fill the pore volume of the support. In wet impregnation the support is dipped into an excess quantity of the precursor solution. The drying process, which follows impregnation, is critical, because it affects the distribution of the active phase. During drying the solution in the pores becomes supersaturated and precipitation takes place. Inhomogeneous evaporation of the liquid (in bigger pores evaporation is faster) results in loss of dispersion of the active phase, and a non -uniform activity profile. After drying, generally treatment in air is applied at temperatures of about 770 870 K. This process is termed calcination. The aim of calcination can be to convert the precursor compound to an oxide, decomposition of anions, creation of porosity, and improvement of the mechanical strength. Care must be taken to avoid excessive temperatures, because they can induce the formation of a mixed oxide with the support (e.g., CuAl2O4, CoAl2O4). This may result in a lower activity or makes further activation (e.g., reduction) difficult. Supported metal catalysts are often produced by treatment in hydrogen at temperatures in the range of 570 770 K. The reduction temperature may influence the metal dispersion. Reduced catalysts are highly pyrophoric and must be handled with great care.
A homogeneous active phase distribution over the catalyst particle is not always desired. It is possible to generate profiles on purpose and in this way improve the catalyst performance. Figure 2.10 shows some possible active phase distributions.
Uniform
Support
Solid Catalysts
Supported catalysts
active phase precursor can be added. In the production of an egg-white platinum catalyst citric acid is added. Citric acid adsorbs in an egg shell and the Pt ions adsorb in a ring at the inside of the citric acid ring. See also reference [1, 6].
2.6.3 Ion-Exchange
Zeolites are by far the largest application for ion-exchange methods. After synthesis of the zeolite it contains sodium. To incorporate the acidity required for catalysis, this sodium may be exchanged by contact with solutions of mineral acids. Other metal ions for specific catalytic applications are introduced by subsequent ion-exchange with salt solutions.
2.7
Economical Aspects
The main sectors in which catalysis is applied are petroleum refining, chemical processing, and environmental protection. Tables 2.7 and 2.8 give an impression of the practical importance of catalysis. Table 2.7 gives a price indication of the most imortant refining catalysts. QUESTION: Why are catalytic reforming catalysts the most expensive?
Table 2.7 Price indication of refining catalysts (~1990). Process Price World catalyst market World process (NLG/kg) (kton / year) Capacity (kton / year) Catalytic cracking 34 350 500000 Hydrotreating 15 50 800000 Catalytic reforming 25 40 5 400000
Although the world refinery catalyst market is quite impressive, its value is modest when compared to the throughput of refinery feedstocks. Therefore, there is a great incentive for improvement of refinery catalysts. Illustrative is that an improvement of only 1% in selectivity to gasoline in the catalytic reforming process, leading to an extra 4 million ton, at a market value of 500 NLG / ton, represents 2000 million NLG / year. Table 2.8 divides the North American catalyst market according to the type of industry in which they are used.
2 - 17
Solid Catalysts
Economical aspects
Table 2.8 North American catalyst market (M$ / year). Industry or type 1992 Petroleum 580 Chemical 1200 Environmental 900 Spent 370 Biocatalyst 550
It is striking that the volume of catalyst used in environmental applications has increased enormously in the past 5 years (and before). The main reason has been the introduction of the three-way catalyst for automotive exhaust control. Emission control today is probably the largest catalyst consumer.
2 - 18
Solid Catalysts
Literature
Literature
1. J.A. Moulijn, Xu Xiaoding, F. Kapteijn and A.D. van Langeveld, Katalyse en Katalysatoren (Eng: Catalysis and Catalysts), TU Delft, 1997. 2. H. van Bekkum, E.M. Flanigen and J.C. Jansen, Introduction to zeolite science and practice, Stud. Surf. Sci. Catal. Vol. 58, 1991. (2nd edition in preparation) 3. J.C.Jansen, M. Stocker, H. Karge and J. Weitkamp, Advanced zeolite science and applications, Stud. Surf. Sci. Catal. Vol. 85, 1994. 4. R. Krishna and S.T. Sie, Strategies for multiphase reactor selection, Chem. Eng. Sci. 49(24A), 4029-4065 (1994). 5. E.R. Becker and J. Wei, Nonuniform distribution of catalysts on supports, J. Catal. 46, 363-381 (1977). 6. J.W. Geus and J.A.R. van Veen, Preparation of supported catalysts, in Catalysis, an integrated approach to homogeneous, heterogeneous and industrial catalysis, J.A. Moulijn, P.W.N.M. van Leeuwen and R.A. van Santen (Eds.), Ch. 9, Elsevier, Amsterdam, 1993.
2 - 19
Introduction
In heterogeneously catalyzed reactions reactant molecules adsorb on the catalyst surface (characterized by equilibrium constants Ki ), undergo modifications on the surface to form adsorbed products with rate constants ki, and these products finally desorb. The surface composition and structure of the catalyst determine its overall activity and selectivity. Therefore, it is important to relate constants, such as ki and Ki to the chemical reactivity of the catalyst surface. Many different modes of adsorption and rearrangements of molecule fragments are possible on catalyst surfaces, so many reactions can occur in parallel. Therefore, catalysts may exhibit satisfactory activity but at the same time have a low selectivity. It is common procedure in process development to carry out an optimization program aimed at increasing the selectivity of the catalyst system selected. Selectivity can often be increased by the use of promotors or catalyst modifiers to create particular surface sites that enhance the desired reaction but suppress undesired reaction paths. In homogeneous catalysis ligands play a similar role as surface sites on solid catalysts with respect to selectivity. In a solution many different catalytic complexes may be present with different catalytic activities. Types and concentrations of ligands determine the structure of the catalyst complexes to a large degree. Thus, catalytic activity and selectivity can be tuned by selecting the right ligands and reaction conditions. Similarly, in biocatalyis enzymes play the role of active sites. Besides activity and selectivity, stability is crucial in catalysis applications (see Chapter 5). Catalyst deactivation can have a kinetic origin. For instance, deactivation might occur by a serial reaction mechanism in which an intermediate can undergo a reaction to form a substance that is a poison for the active catalyst sites. Frequently encountered examples are oligomerization and coke formation. Adsorption on a solid catalyst surface and complex formation in homogeneous catalysis share the same principle, i.e., the total number of sites is constant. Therefore, the rate expressions for reactions 3-1
Rate expression
on heterogeneous and homogeneous catalysts have a similar form. Partial pressures are usually used in rate expressions for gas-phase reactions, while concentrations are used when the reactions take place in the liquid phase. The approach discussed in the next sections for the derivation of rate expressions can be applied to both homogeneous and heterogeneous catalyst systems. Two principles are important: Catalysts do not affect the position of the overall equilibrium of a reaction, they only affect the rate at which it is achieved; The total number of catalytically active sites is assumed to be a constant. In heterogeneous catalysis this site density is expressed as the number of sites per unit mass of catalyst or per unit surface area, in homogeneous catalysis as the concentration of catalytically active complex molecules. In fact, the above assumption is not always justified. There are examples where the number of active sites is a function of process variables such as temperature.
1
A*
3 2
B*
Langmuir adsorption
1. 2. 3.
A + *
A*
k1 k-1 k2
k-2
A*
B*
k3
B*
k-3
B + *
Rate expression
1. Molecule A adsorbs on an active site * at the catalyst surface under formation of an adsorbed complex A*. 2. This adsorbed complex reacts on the active site by rearrangement to an adsorbed complex B*. 3. Finally, product B desorbs from the active site, liberating the site for a new catalytic cycle. Since the three steps are considered to be elementary processes, their rates can be directly derived from their reaction rate expressions:
r1 = r+1 r1 = k 1 p A N T * k 1 N T A
r2 = r+ 2 r 2 = k 2 N T A k 2 N T B
r3 = r+ 3 r3 = k 3 N T B k 3 p B N T *
Note that under steady-state conditions the rate of each reaction step equals the overall net rate, r. *, A, and B represent the fractions of the total number of sites that are vacant, or occupied by A and B, respectively. NT represents the total concentration of active sites with possible dimension mol(kg cat)1 . For this gas-phase reaction, partial pressures of A and B are preferred in the rate equations, but for liquid-phase reactions molar concentrations should be used. Conservation of the total number of active sites leads to the site balance expression:
1 = * + A + B
(5)
Under steady-state conditions, which are usually satisfied in continuous-flow processes, the concentrations of the vacant sites, sites occupied by A, and sites occupied by B, do not change with time:
d A = 0 = k1 p A * k 1 A k 2 A + k 2 B dt d B = 0 = k 3 p B * k 3 B k 2 B + k 2 A dt
(6)
(7)
From equations (5-7), the unknowns *, A, and B can be expressed as functions of the rate constants and partial pressures and then substituted in the overall net rate equation:
r = r1 = r2 = r3
Finally, equation (9) is obtained for the reaction rate expression:
(8)
3-3
r=
N T k1 k 2 k 3 ( p A p B / K eq ) (k1 k 3 + k1 k 2 + k1 k 2 ) p A + (k 1 k 3 + k 2 k 3 + k 2 k 3 ) p B + (k 1 k 2 + k 1 k 3 + k 2 k 3 )
(9)
with Keq = K1 K2 K3 (Ki = ki / k-i) being the overall equilibrium constant for the reaction. Rate expression (9) has been derived for a relatively simple kinetic model by application of the site balance and the steady-state hypothesis. More complex models will result in more complex expressions (see references [1] and [2] for general derivation), which are hard to handle. Fortunately, usually some sort of simplification is justified.
r = r2 - r-2
Figure 3.2 Visualisation of the quasi-equilibrium and rate-determining steps. The lengths of the arrows are proportional to the rates of the relevant steps.
The rate expression is now obtained as follows. Starting with the rate-determining step:
r = r+ 2 r 2 = k 2 N T A k 2 N T B
and using the quasi-equilibrium conditions:
(10)
K1 =
k1 = A k 1 * p A
A = K 1 p A *
(11)
3-4
B =
p B * K3
(12)
enables the elimination of the unknown quantities A and B from (10). The remaining * can be eliminated by use of the site balance (5):
* =
1 1 + K1 p A + p B / K 3
(13)
The resulting rate expression for the case that the surface reaction is rate determining is given by equation (14):
r=
N T k 2 K1 ( p A p B / K eq ) 1 + K1 p A + p B / K 3
(14)
r=
N T k1 ( p A p B / K eq ) 1 + (1 + 1 / K 2 ) p B / K 3
(15)
(16)
QUESTION:
The terms in the denominator have a physical meaning. Explain this for Eq. (14), (15), and (16).
If besides reactant A and product B also other species are present that adsorb on the active sites, they make these sites unavailable for reaction, and hence lower the reaction rate. The effect of such inhibitors (I) should be included in the rate expression and can be summarized for the considered reaction (and surface reaction rate determining) in equation (17):
r= k(p A p B /K eq ) 1 + K A p A + K B pB +
pI
(17)
In equation (17), KA, KB, and KI represent the adsorption equilibrium constants of the components A, B, and I, respectively. k represents the apparent (observed) overall reaction rate constant. It will be clear that KA equals KI and KB equals 1/K3 in equation (14).
3-5
Adsorption isotherms
3.4.1
From the foregoing it can be easily derived that in case of one-component adsorption, the fractional coverage of the adsorbed species A can be described by equation (18). Figure 3.3 shows the graphical representation of equation (18) for various values of KA ( K1).
100
0.8
10
KA (bar-1)
A (-)
A =
K A pA 1+ K A pA
(18)
1 0.1
PA (bar)
Figure 3.3 Surface coverage as a function of pA for several values of KA.
Three regions can be distinguished for the Langmuir isotherm: 1. At low values of KApA, a linear relation exists between A and pA, with slope KA. This is a relation equivalent to Henrys law. 2. At high values of KApA , A approaches 1, i.e., nearly all sites are occupied. This can be easily seen by rearrangement of equation (18):
3-6
Adsorption isotherms
(19)
3. For values of KApA near 1, i.e., at moderate coverage, the complete equation (18) must be used. Alternatively, the Freundlich isotherm (Eq. 21) can be used. With increasing temperature, KA decreases and a transition from situation 2 to 1 via 3 can be expected. KA is expressed in units of (pressure)-1. If atm-1 (1 atm is the standard thermodynamic reference state for gases) is used, KA can be expressed as:
o o o G A S A H A ln K A = = RT R RT
or
0 H A = RT 2
ln K A T
(20)
in which Go, Ho, and So represent the Gibbs free energy, enthalpy and entropy of adsorption, respectively. Both the enthalpy and entropy of adsorption are generally negative since adsorption is an exothermic process, and the molecule loses at least one translational degree of freedom. The assumptions for the Langmuir isotherm imply an ideal surface, but few real systems will fulfil this ideal under all conditions. Experimental determination of the heat of adsorption ( H 0 ) as a function of the surface coverage shows that the heat of adsorption usually decreases with increasing coverage. This indicates that catalyst surfaces are not uniform and/or that the adsorbed molecules exhibit a mutual interaction. Adsorption isotherms that take this coverage dependence into account are amongst others the Freundlich and Temkin isotherms. The Freundlich isotherm assumes a logarithmic dependence of H 0 on A, which leads to equation (21) [3]:
/n A = c p1 A
(21)
and is originally empirical. The parameters n and c usually both decrease with increasing temperature. By adjusting n and c, most data can be fit by this model. n represents the mutual interaction of adsorbed species. n > 1 means that adsorbed species repulse each other. The Temkin isotherm takes into account the varying heat of adsorption by assuming a linear relationship of - Ho with A [3] and using equation 18:
o o H A = H A 0 (1 A )
(22)
o H A 0 A0 = a 0 exp RT
A =
RT ln( A0 p A ) o H A 0
(23)
o where H A 0 is the heat of adsorption at zero surface coverage, and a0 are constants, and A0 is
independent of surface coverage. The Temkin isotherm is valid at moderate surface coverage. QUESTION: Derive equation 23 for moderate coverage A (representative value 0.5).
3-7
Adsorption isotherms
In kinetic modelling practice, these isotherms are hardly used since the derivation of rate expressions becomes a cumbersome job (see e.g., [1]), in particular for multicomponent systems. This situation might rapidly change with the increasing use of powerful computers, but the simple physical picture of the catalyst surface will always remain the advantage of the Langmuir model.
3.4.2
Multicomponent Adsorption
Within the Langmuir approach, the general expression for the fractional coverage by component A in the case of multicomponent adsorption is given by equation (24) and illustrated in Figure 3.4.
A =
(1 + K
K A pA
A
pA +
pI
(24)
3.4.3
Dissociative Adsorption
In the foregoing, only molecular adsorption has been considered. However, some molecules (e.g., H2, CO) can dissociate upon adsorption, and hence, two sites are required, as illustrated in Figure 3.5.
Figure 3.5 Dissociative adsorption of H2. The dissociative adsorption of hydrogen proceeds through the following adsorption equilibrium:
3-8
Adsorption isotherms
H2 +
2 * 2 H*
In analogy with molecular adsorption, the coverage with hydrogen can be derived as a function of the adsorption equilibrium constant and hydrogen partial pressure: K H2 pH2 1 + K H2 pH2
H =
(25)
At low values of K H 2 p H 2 (this is at low coverage), the coverage is proportional to the square root of the partial pressure of H2. QUESTION: Derive equation (25).
Often multiple sites are involved in a catalytic process. This is especially the case for dissociation reactions. The same procedure as applied in Section 3.1 for a single-site model can be used for the derivation of the rate expression for a dual-site model. The procedure is exemplified for the dissociation reaction A 2 B, which is thought to proceed through the following three elementary steps: 1. 2. 3.
A + *
A* + * B*
k1 k-1 k2 k-2 k3 k-3
A*
2 B* B + *
1 1 2
Since steps 1 and 2 must proceed once per overall reaction and step 3 twice, the so-called stoichiometric numbers of the steps are one, one, and two, respectively. By application of the steady-state hypothesis (surface occupancies do not change with time), a site balance, and the assumption that the surface dissociation reaction is rate determining, while the other steps are in quasi-equilibrium, the following rate expression is derived:
3-9
Rate expressions
r=
(1 + K A p A + K B p B )2
(26)
in which Keq = K1K2K32. The numerator includes a parameter s, which represents the number of nearest neighbours of an active site. The necessity of this parameter in the rate equation can be understood as follows. In step 2, the rate-determining step, adsorbed A reacts with an empty site. This reaction is only possible when there is an empty site next to A. Therefore, the rate of this reaction is proportional to the concentration of adsorbed A (NTA) multiplied with the number of adjacent sites s times the chance that they are empty *. Using the total concentration of empty sites (NT*) instead, would lead to overestimation of the reaction rate. KA and KB are the adsorption equilibrium constants of A and B and equal to K1 and 1/K3, respectively. The denominator is now squared compared to a single-site model. This indicates that in the rate-determining step two active sites are involved.
3.5.2
Eley-Rideal Models
The models described above are termed Langmuir-Hinshelwood-Hougen-Watson (LHHW) models, named after the scientists that contributed a lot to the development of these engineering models. The characteristics of these models are that adsorption follows the Langmuir isotherm, and that reaction takes place between adsorbed species. Sometimes, one encounters Eley Rideal models, whereby a molecule directly from the gas phase with a surface complex: A + B* C*
3.5.3
So far we have treated only reactions consisting of one adsorption, one reaction, and one desorption step. However, real reaction schemes are often much more complex. Despite this complexity often only one or two steps are kinetically significant (i.e., the slowest steps) so that the kinetic model can be reduced to a simple one. The reaction A B for instance could in fact be built up of many elementary steps: A + * A* X1* X2* X3* B* B + * in which X1* to X3* are adsorbed intermediates. The model contains 5 reactive intermediates and 12 rate constants. Here, the approach of the most abundant reaction intermediates (mari) together with the use of the site balance (Eq. 5) is useful to eliminate some steps from the kinetic model, reducing the model to only the kinetically significant steps. A lot of possible surface species can be eliminated based on experimental result or thermodynamic data (see Case study IV). For instance, assuming the adsorption of A is the rate-determining step and that A* is the most abundant intermediate, all 5 quasi-equilibria following the adsorption of A can be lumped to yield an overall quasi-equilibrium:
3 - 10
Hence, the treatment of this multistep reaction network is very much simplified. If B* would also be present on the surface in an appreciable amount (i.e., desorption of B kinetically significant) the model would simplify to the same model as before for the reaction A B (Section 3.3). QUESTION: Derive the rate expressions for both cases (surface A most abundant, and surface A plus B most abundant).
3.5.4
Generalized Models
Other variants of kinetic models can of course be derived. Froment and Bischoff [4] present an extended treatment of this approach. It follows that rate expressions based on sequences of elementary steps of which one is rate determining can be expressed in the general form: rate =
(27)
The kinetic factor always contains the rate constant of the rate-determining step, together with the total concentration of active sites, and adsorption equilibrium constants. The driving force represents the chemical affinity of the overall reaction to reach thermodynamic equilibrium. It is proportional to the concentration difference of the reactants with respect to their equilibrium concentrations. The driving force term does not contain parameters associated with the catalyst, consistent with the fact that the catalyst does not affect chemical equilibrium. The adsorption term represents the reduction of the overall rate due to adsorption, with the individual terms denoting the distribution of the active sites over the different intermediate surface species and vacancies. It may contain square roots of partial pressures, indicating dissociative adsorption. The power n in the rate expression indicates the number of sites involved in the rate-determining process, and usually has a value of 0, 1, or 2. Larger values reported in literature merely represent values obtained by fitting. In homogeneous catalysis, enzyme reactions, and reactions of microbial systems the same types of equations are used as in the LHHW models. In the latter disciplines, however, they are often referred to as Michaelis-Menten relations.
3 - 11
Limiting cases
r0 = N T k 1 K A p A0
Surface reaction rate determining: r0 = N T k 2 K A p A0 1 + K A p A0
(15a)
(14a)
In the extreme case of low KA pA0 equation (14a) reduces to: r0 = N T k 2 K A p A0 (14b)
which is a similar expression to equation (15a), and cannot be distinguished from it experimentally. QUESTION: Why can equations (14b) and (15a) not be distinguished experimentally?
On the other hand, if KA pA0 >> 1, equation (14a) reduces to equation (14c):
r0 = N T k 2
Desorption rate determining: r0 = N T k 3 K A K 2 p A0 1 + ( 1 + K 2 )K A p A0
(14c)
(16a)
When desorption is rate limiting, the surface is nearly fully occupied and expression (16a), for K2 and KA pA0 >> 1, reduces to:
r0 = N T k 3
(16b)
3 - 12
Limiting cases
Figure 3.6 represents the pressure dependences of r0 for these three cases. The effect of temperature is also shown.
Adsorption
T1
Surface reaction
T1
Desorption
T1
r0
T2 T3
T2 T3
T2 T3 pA0
pA0
pA0
Figure 3.6 Dependence of the initial reaction rate on pressure and temperature for three different rate-determining steps. T1 > T2 > T3.
Evidently, this initial rate pressure dependence gives a quick insight in which kinetic model describes best the experimental results of the reaction under consideration. Initial rate experiments are ideal for model discrimination purposes where one tries to select the best kinetic description of a process. However, two important aspects must be realized: The product partial pressure may be low, but when the product is strongly adsorbed not all the terms in the denominator can be neglected. In this case, only the numerator can be simplified. Other components in the reaction mixture may compete for adsorption sites and occupy part of the active sites. Hence, a variation of the reactant pressure will have less effect than would be expected. For practical purposes often high conversions are needed, which requires complete models. Derive simplified rate expressions in the case of high conversion of A (or large amount of product B (>> A) added to the feed).
QUESTION:
Example 3.1
Ethanol dehydrogenation
is believed to proceed according to the following kinetic model [5], in which hydrogen is assumed to adsorb in molecular form: 1. A + * A* 2. A* + * R* + S* (r.d.s.) 3. R* R+* 4. S* S+* The full rate expression for this model is represented by equation (28): 3 - 13
Limiting cases
r=
k 2 s N T K A p A - p R p S / K eq
(1+ K A p A + K R p R + K S p S )
(28)
When the products are weakly or moderately adsorbed the initial rate expression becomes: r0 = kK A p A (29)
(1+ K A p A )2
In order to validate the proposed model, and to determine the model parameters k (= k2sNT) and KA from rate versus pressure data, equation (29) can be transformed into (30): pA KA 1 = + pA r0 kK A kK A
(30)
If the model fits the data, equation (30) yields a straight line. However, if step 2 would have been the reaction A* R* + S, in which hydrogen desorbs directly as it dissociates from ethanol, no straight line would be obtained, indicating that the chosen model is incorrect. Similarly, if the adsorption step or one of the desorption steps would have been rate limiting, equation (30) would not fit the data. Indeed, at initial ethanol pressures of 1 to 10 bar, a good fit was obtained. The reaction rate constant k increased from 1.07 to 5.58 mol (hr-1 g.cat-1) when the temperature increased from 500 to 560 K, while the adsorption equilibrium constant KA decreased from 0.74 to 0.44 atm-1 in the same temperature range [5]. These trends are consistent with the usual behaviour of rate constants and adsorption constants.
Example 3.2
CO dissociation over Rh
In view of the decreasing crude oil reserves, C1-chemistry, i.e., chemistry based on natural gas (mostly methane), will probably play a major role in the future production of chemicals and automotive fuels. In principle, methane can be converted directly, for instance through oxidative coupling to form ethene. However, methane is very thermodynamically stable, and hence is not very reactive. Therefore, current practice to convert natural gas into higher hydrocarbons proceeds by an indirect route in which it is first converted to synthesis gas (a mixture of hydrogen and carbon monoxide) at high temperature. Subsequently, hydrocarbons can be produced by either Fischer-Tropsch synthesis or via methanol and the Methanol-to-Gasoline process. The key step in these processes is the dissociation of CO. Rhodium is known to be very active for this reaction [6], for which the kinetic model is: 1. CO + * CO*
3 - 14
Limiting cases
(1+ K CO pCO )2
Figure 3.7 shows the dependence of the rate on temperature and surface occupancy.
800
600
Rate
400
200
0 0.2
Tem
ture per a
(K)
There is an optimum surface coverage for CO dissociation, which would be expected from the first equality of equation (31). QUESTION: Explain qualitatively the trends in Figure 3.7.
*A
*A#
kbarrier
*B
k-#
3 - 15
It is assumed that the reacting complex (*A) is in equilibrium with the transition-state complex (*A#), and that the number of molecules in the transition state that react to the product (*B) per unit of time is given by the frequency kbarrier. This latter step is assumed to be rate limiting. The reaction rate constant can then be computed from expression (32):
# k+ k = k barrier k#
= k barrier K #
(32)
QUESTION:
Expression (32) is valid when energy exchange is fast compared to the overall reaction rate. Since K # is an equilibrium constant, it can be written as in equation (33), where G#, H#, and S# are the Gibbs free energy, the enthalpy, and the entropy differences between the transition state and the ground state, respectively.
- G # K # = exp RT
S # H # exp = R - RT
(33)
As long as quantum-mechanical corrections can be ignored, the rate of reaction of the transition-state complex is the same for all reactions and is given by equation (34):
k barrier = k BT h
(34)
where kB is the Boltzmann constant and h is Plancks constant. The overall rate constant then becomes:
k=
S # H # k BT Ea exp = k 0 exp h RT RT R
(35)
S#/R may be taken as a constant, because it only varies slightly with temperature. Furthermore, since
the exponential term is so much more temperature-sensitive than the pre-exponential term, this latter term may also be taken as a constant, resulting in the second equality of equation (35).
H# can be identified with the activation energy Ea (neglecting a contribution of RT) of the reaction (see Figure 3.8). In Figure 3.8, H represents the reaction enthalpy for the reaction A* B*. More
detailed treatments are presented by Maatman [7], Boudart, and Djg-Mariadassou [2], and Van Santen and Niemantsverdriet [8].
3 - 16
A*#
obs Ea = Ea2
Ea2
A*
B*
Figure 3.8 Energy diagram for reaction A B, strong adsorption of A.
(36)
where KA = K1 and KB = 1/K3. The dependence of the reaction rate (36) on pressure and temperature is determined by contributions of both the numerator and denominator. Often, rate expressions are presented in a power law form:
n A nB r = kp A pB
(37)
Comparison with equation (36) suggests that the powers nA and nB will not be constants. They can be extracted as follows:
ni =
ln r ln pi
(38)
Applying equation (38) to (36), and using the adsorption equilibrium relationships (Eqs. (11) and (12)), together with equation (13), yields:
nA = 1 A
and
nB = B
(39)
Hence, the apparent reaction orders are related to the fractional surface coverages. From equation (39) it follows that nA varies from 0 to 1, and nB from 1 to 0, depending on the conditions.
3 - 17
Like the pressure dependence, the temperature dependence is also often expressed in an empirical form, in which an apparent overall rate constant is used (like in Eq. (37)). The observed (or apparent) activation energy, following from equation (37), can be expressed as: ln r obs Ea = RT 2 T p
(40)
The observed activation energy can now be derived from equation (36) in a similar way as the derivation of the reaction powers as a function of surface coverage:
obs Ea = E a 2 + ( 1 A )H A B H B
(41)
The observed activation energy contains contributions from the rate-determining step (Ea2) and from the adsorption enthalpies of A and B, the latter depending on the fractional occupancies. Obviously,
obs Ea will depend on the experimental conditions. Therefore, it is not surprising that a wide range of
(42)
Physically this implies that the whole catalyst surface is covered with A (A 1, B 0). Therefore, varying the partial pressure of A does not influence the reaction rate. The reaction is said to be zero
obs order in A (and B). The overall activation energy is E a = E a 2 (see Figure 3.8), provided the
3 - 18
The reaction is now first order in A and zero order in B. The observed overall activation energy will be lower than in the previous case:
obs Ea = E a 2 + H A
(44)
where H A represents the enthalpy of adsorption of A, which is negative since adsorption is an exothermic process. This result can be understood from the energy diagram in Figure 3.9. Due to the low occupancy of A, for reaction to occur, adsorption of gaseous A is required, and hence A gains adsorption enthalpy. Subsequent surface reaction requires overcoming the activation energy.
A*#
obs Ea = E a 2 + H A
A(g) + * A*
Ea2
HA
(45)
The reaction is first order in A and minus one in B (B decreases the reaction rate strongly by competitive adsorption). The surface is nearly completely covered with B. Therefore, the initial state for the reaction is gaseous A and adsorbed B (see Figure 3.10). For A being able to react, firstly a molecule of B must desorb with accompanying desorption enthalpy. Subsequently, A adsorbs, gaining adsorption enthalpy, and reacts through the surface reaction, where the activation energy barrier has to be overcome. Thus, the observed activation energy is higher than in the previous cases:
obs Ea = E a 2 + H A H B
(46)
A*#
obs Ea = E a 2 + H A H B
B+*+A
- HB
Ea2
HA
A*
B* + A
Figure 3.10 Energy diagram for reaction A B, strong adsorption B.
<
E a 2 + H A H B
(47)
Analogously, limiting cases can be distinguished for the dual-site model, in which the order of A can even become negative (see for example equation 26). This is common for dissociation reactions. QUESTION:
How does the apparent activation energy behave in the CO dissociation over Rh (see example 3.2)?
An excellent illustration of the LHHW theory is catalytic cracking of n-alkanes over ZSM-5 [9]. For this reaction, the observed activation energy decreases from 140 to -50 (!) kJ/mol when the carbon number increases from 3 to 20. The decrease appeared to linearly depend on the carbon number as shown in Figure 3.11. This dependence can be interpreted from a kinetic analysis that showed that the hydrocarbons (A) are adsorbed weakly under the experimental conditions. The initial rate expression for a rate-determining surface reaction applies (Eq. 14a), which in the limiting case of weak adsorption of A reduces to equation (43). The activation energy is then represented by equation (44).
200 100
Ea2 Eaobs
kJ/mol 0
-100 -200
H A
Carbon number
Figure 3.11 Observed activation energy, reaction activation energy, and adsorption enthalpy in the cracking of n-alkanes over ZSM-5 (adapted from [9]).
Measurement of the adsorption enthalpy HA revealed a linear decrease with carbon number. By applying equation (44) the activation energy for the surface reaction, Ea2, was estimated. The data in Figure 3.11 clearly shows that Ea2 has reasonable values, while it remains fairly constant for n > 8, supporting the kinetic interpretation.
3 - 21
rate limiting. The temperature dependence of the overall rate will behave as depicted in Figure 3.12. This figure illustrates that the observed overall activation energy decreases with increasing temperature upon a change in the rate-determining step.
adsorption r.d.s.
ln robs
desorption r.d.s.
1/T
Figure 3.12
Change of observed activation energy due to changing rate-determining step as a function of temperature.
In instances where the rate-determining step changes upon a change in conditions one cannot comply with the assumption of only one rate-determining step. To obtain an adequate rate expression, valid over the whole temperature range under consideration, two or even more steps should be assumed not to be in quasi-equilibrium, but rate determining.
An example of a reaction, in which more than one step is rate determining, is the dehydrogenation of methylcyclohexane (M) to toluene (T) over a platinum catalyst in the presence of hydrogen (to reduce coke formation). This reaction was studied by Sinfelt et al. [3, 10]. Results of initial rate experiments suggested the following sequence of elementary steps: 1. M + * M* 2. M* T+*
r.d.s.
For this kinetic model the initial rate expression in equation (48) holds:
r0 = k 2 NT K M pM 1+ K M p M
(48)
In a similar way as in Example 3.1, this equation can be linearized and the constants (k2NT) and KM determined. Plotting the constants as a function of temperature in an Arrhenius plot yields the activation energies for surface reaction (Ea2 = 125 kJ/mol) and methylcyclohexane adsorption enthalpy (-HM = 79 kJ/mol), respectively. However, this equation was checked by adding aromatics, and it was found that no inhibition occurred by, e.g., benzene, which is known to adsorb strongly on the catalyst. This and other observations suggest that methylcyclohexane is not in a state of adsorption equilibrium, but adsorbs irreversibly. Furthermore, it is assumed that toluene is the most abundant reaction intermediate (mari), 3 - 22
and that it desorbs irreversibly. These considerations lead to the following kinetic model with two rate-determining steps: 1. M + * M* 2. M* T* 3. T* T+*
fast
The initial rate expression takes a similar form as the previous one:
r0 = k 3 N T (k 1 / k 3 ) p M 1 + (k 1 / k 3 ) p M
(49)
Note that the surface concentration M has been assumed to be small in deference to the observation that aromatics hardly affect the rate. Now, the constant previously attributed to the surface reaction (k2NT), in fact is that of desorption of the strongly bound toluene (k3NT, Ea3 = 125 kJ/mol), while the constant previously attributed to adsorption of methylcyclohexane (KM) really is the ratio of toluene desorption to methylcyclohexane adsorption (k1 / k3 exp( E a1 + E a 3 ) ). The activation energy for adsorption of methylcyclohexane can then be determined to be 46 kJ/mol. This example illustrates that similar rate expressions can be obtained from different kinetic models. So, kinetics cannot prove the correctness of a reaction model, they can only support or reject it.
Rate
Heat of adsorption
3 - 23
Sabatier principle
Figure 3.13 Volcano plot for the overall reaction rate as a function of the heat of adsorption.
Plots like Figure 3.13 are called Volcano plots. Such plots have been measured for very different reactions, e.g., formic acid decomposition, ammonia synthesis, hydrodesulfurization, and hydrodenitrogenation reactions. As explained above, the increase of the reaction rate is due to the increased site-coverage with the reactant (Eq. 18). Once the optimum surface coverage has been achieved, the reaction rate reaches a limit. The rate is then controlled by the rate of product formation. The decrease in reaction rate with further increase in heat of adsorption (surface coverage) is due to the increased activation energy for desorption; not only the reactants will be adsorbed more strongly, so will the products (Eq. 20). Clearly, an optimum for the interaction strength between the catalytically active surface and the adsorbates exists, resulting in a maximum in the rate of reaction (the Sabatier principle). To the left of the maximum, the reaction has a positive order in the reactants, whereas to the right the order has become zero or even negative (see Eq. 26). This kinetic dependence can be used to test whether a Volcano plot is due to Sabatiers effect or not. In practice, the variation of the rate of reaction with adsorption strength and the occurrence of a maximum in this rate have the following consequence for kinetic modeling of heterogeneous catalysts. Usually, the assumption of a homogeneous surface is not strictly valid. It would probably be more realistic to assume the existence of a certain distribution in the activity of the sites. However, certain sites will contribute most to the reaction, since these sites activate the reactants most. This might result in apparently uniform reaction behaviour, and can explain why Langmuir adsorption often provides a good basis for the reaction rate description. This also implies that adsorption equilibrium constants determined from separate adsorption experiments can only be used in kinetic expressions when coverage dependence is explicitly included. Otherwise, these constants have to be extracted from the rate data. Several authors have derived rate expressions for non-uniform catalyst surfaces. Boudart and DjgMariadassou [2] show that relations are obtained with a mathematical similarity to those obtained for a uniform surface. In the rate expression for ammonia production, the Temkin isotherm (Eq. 21) has been used for a long time. This isotherm accounted for a, supposedly, heterogeneous adsorption behaviour [1]. Recently, however, it has been shown that the LHHW approach can account for the data over a pressure range of 300 bar [11] without the assumption of heterogeneity. This and other examples demonstrate the usefulness of the LHHW approach for reactions of practical importance.
Catalytic Reaction Kinetics the surface for adsorption and reaction is uniform; the number of active sites is constant, irrespective of reaction conditions; adsorbed species do not interact, apart from their reaction paths.
Sabatier principle
The form of the resulting expression differs from rate expressions for classical homogeneous gasphase reactions due to the presence of a denominator representing the reduction in rate due to adsorption phenomena. The individual terms of this denominator represent the distribution of the active sites among the possible surface complexes and vacancies. Expressions of this type are termed Langmuir-Hinshelwood-Hougen-Watson (LHHW) rate expressions. The steady-state approach generally yields complex rate expressions. A simplification is obtained by the introduction of one or several rate-determining step(s) and quasi-equilibrium steps, and further by the initial reaction rate approach. For complex reaction schemes, identifying the most abundant reaction intermediates (mari) and making use of the site balance can simplify the kinetic models and rate expressions. In practice, useful relations result even for the non-ideal heterogeneous surfaces of solid catalysts. Some reasons can be: similarity of mathematical relations for uniform and non-uniform adsorption models; Sabatiers principle of the optimum site activity. Optimum sites contribute most to the reaction, resulting in an apparent uniform behaviour.
This chapter has neither dealt with the aspects of kinetic model selection/discrimination and parameter estimation, nor with the experimental acquisition of kinetic data, since these subjects are beyond its scope. Moreover, in interpreting the observed temperature dependence of the rate coefficients in this chapter we assumed to be dealing with intrinsic kinetic data. As discussed in Chapter 4 and the literature [12-15], parasitic phenomena of mass and heat transfer may interfere, disguising the intrinsic kinetics. Criteria are presented there to avoid this experimental problem.
3 - 25
Notation
Notation
a0 A0 c ci Ea h kB kbarrier ki Ki n ni NT pi r ri R s t T
Greek
constant in Temkin isotherm adsorption constant at zero surface coverage constant in Freundlich isotherm concentration activation energy Plancks constant Boltzmann constant number of molecules reacting per unit time reaction rate constant for reaction i equilibrium constant of reaction i constant in Freundlich isotherm reaction order in i total concentration of active sites partial pressure of component i reaction rate (overall) reaction rate of reaction i ideal gas constant number of nearest neighbours of active site time temperature
atm-1 atm-1 atm n mol m-3 J mol-1 Js J K-1 s-1 s-1 (for first order reaction) atm-1 or m3 mol-1 mol (g cat)-1 or mol (m2 cat)-1 atm, kPa mol m-3 s-1 mol m-3 s-1 J mol-1 K-1 s K
G H S i
constant in Temkin isotherm Gibbs free energy Enthalpy Entropy fraction of total number of sites occupied by i
Subscripts
0 + eq g obs
initial or at zero coverage forward reaction backward reaction equilibrium gas phase observed / apparent
Superscripts
0 3 - 26
standard conditions
Catalyst effectiveness #
Other
Literature
transition state
Literature
1. M.I. Temkin, The kinetics of some industrial heterogeneous catalytic reactions, in Advances in catalysis, Vol. 28, Academic Press, New York, p. 173 (1979). 2. M. Boudart and G. Djg-Mariadassou, Kinetics of heterogeneous catalytic reactions, Princeton University Press, Princeton, New York, (1984). 3. C.N. Satterfield, heterogeneous catalysis in industrial practice, 2nd ed., McGraw-Hill, New York, p.61 (1991). 4. G.F. Froment and K.B. Bischoff, Chemical reactor analysis and design, 2nd ed., Wiley, New York, (1991). 5. J. Franckaerts and G.F. Froment, Chem. Eng. Sci. 19, 807 (1964). 6. T. Koerts, The reactivity of surface carbonaceous intermediates, Ph.D. Thesis, Eindhoven University of Technology, 1992. 7. R.W. Maatman, Site density and entropy criteria in identifying rate-determining steps in solidcatalyzed reactions, in Advances in catalysis, Vol. 29, Academic Press, New York, p. 97 (1980). 8. R.A. van Santen and H. Niemantsverdriet, Chemical kinetics and catalysis, Plenum, (1995). 9. J. Wei, Ind. Chem. Eng. Res., 33, 2467 (1994). 10. H.H. Lee, Heterogeneous reactor design, Butterworth, Boston, p. 77 (1985). 11. P. Stoltze and J.K. Nrskov, An interpretation of the high-pressure kinetics of ammonia synthesis based on a microscopic model J. Catal. 110, 1 (1988). 12. J.J. Carberry, Physico-chemical aspects of mass and heat transfer in heterogeneous catalyis, in Catalysis, science and technology, Vol. 8, Springer, Berlin, p. 131 (1987). 13. O.A. Hougen and K.M. Watson, Chemical process principles, Vol. III, Wiley, New York, (1947). 14. J.M. Thomas and W.J. Thomas, Introduction to the principles of heterogeneous catalysis, Academic Press, London, (1967). 15. R. Mezaki, H. Inoue, Rate equations of solid-catalyzed reactions, University of Tokyo Press, Tokyo, (1991). F. Kapteijn and J.A. Moulijn, Kinetics and transport processes, in Handbook of heterogeneous catalysis Vol. 3, Ch. 6, G. Ertl, H. Knzinger and J. Weitkamp (Eds.), VCH, Weinheim, 1997.
3 - 27
Case study I
Chemical Kinetics of Catalytic Reactions - Case Studies Case Study I: N2O Decomposition
C.I.1 Oxide catalyzed N2O decomposition
Winter [16] found the following three limiting rate expressions depending on reaction conditions and catalyst: 1st order: strong O2 inhibition: moderate O2 inhibition:
r = k obs p N 2O
r = k obs r=
(C.I.1) (C.I.2)
( pO )0.5
2
p N 2O
1 + ( pO 2 K 3 )
k obs p N 2O
0.5
(C.I.3)
Furthermore, the reaction order of N2O varied from 0.5 to 1. QUESTIONs: Propose a kinetic model which would explain these limiting rate expressions, derive the rate expression, and show how this rate expression reduces to one of the three limiting rate expressions. ANSWER: The observation of orders of N2O < 1 implies that a N2O adsorption term must appear in the denominator of the rate expression. Therefore, the first step in the reaction cycle must be the reversible adsorption of N2O on a vacant catalyst site. This step would have to be in quasiequilibrium. Furthermore, the adsorption term in the rate expressions contains the square root of the O2 partial pressure. This implies associative desorption of O2 (the reverse of dissociative adsorption of O*), also in quasi-equilibrium. The link between the adsorption of N2O and the desorption of O2 is the reaction of adsorbed N2O to yield gaseous N2 and adsorbed oxygen, which is the rate-determining step. The kinetic model, amongst others reported by Vannice et al. [17], becomes: i 2 2 1
1. N2O + 2. N2O* 3. 2 O*
N2O* N2 + O* O2 + 2 *
r.d.s.
C3 - 1
Case study I
Steps 1 and 2 must proceed twice and step 3 once in each catalytic cycle. Step 2 must be an irreversible step since the overall reaction is irreversible. The kinetic model leads to the following rate expression:
r= k 2 N T K 1 p N 2O
1 + K 1 p N 2O + ( pO 2 K 3 )
0.5
(C.I.4)
( pO
Eq. (C.I.2) when ( p O 2 K 3 ) >> 1 and K 1 p N 2O , e.g., at high partial pressure of O2 or low Eq. (C.I.3) when ( p O 2 K 3 ) and 1 >> K 1 p N 2O , e.g., for low K1
0.5
values of K3
C.I.2
Vannice et al. [17] tested rate equation (C.I.4) by determining the rate of N2O decomposition over a Mn2O3 catalyst as a function of temperature and partial pressure of N2O and O2 and fitting Eq. (C.I.4) to the experimental data. From the dependence of the rate on the partial pressure of N2O at various temperatures they found a reaction order in N2O somewhat lower than 1 (0.78), suggesting moderate surface coverage with N2O. Figure C.I.1 shows the influence of the partial pressure of O2 on the rate of N2O decomposition at various temperatures and a partial pressure of N2O of 10 kPa.
0.4
pN2O = 10 kPa
r / 10-6 mol.s-1.g-1
0.3
648 K
0.2
638 K
0.1
0.0 0.0
pO2 / kPa
Figure C.I.1
Dependence of rate of N2O decomposition over Mn2O3 on oxygen partial pressure [17].
Figure C.I.1 clearly shows that oxygen has an inhibiting effect on the N2O decomposition reaction at all temperatures. C3 - 2
Case study I
Enthalpy and entropy changes for adsorption and desorption steps 1 and 3, and the activation energy for step 2, the rate-determining surface reaction, were calculated from the parameter estimates for k (k2NT), K1 and K3 at different temperatures (see Eqs. (20) and (35)) [17].
H1 S 1
= =
- 29 kJ/mol - 38 J/molK
H3 S 3
= =
QUESTIONs: Are these values thermodynamically consistent and consistent with the proposed kinetic model? ANSWER Both the enthalpy and entropy change of step 1, adsorption of N2O, are negative which is in agreement with the exothermicity of adsorption and the loss of at least one degree of freedom upon adsorption of a molecule. The values for step 3 are positive, because in step 3 desorption of O2 takes place, the reverse of adsorption. So, these values are thermodynamically consistent and consistent with the kinetic model. The adsorption of O2 is much stronger than that of N2O. At a temperature of 600 K the adsorption constants have a value of 3.5 and 207 atm-1, respectively (from [17], can also be calculated using Eq. (20)). However, the partial pressure of O2 will be lower than that of N2O, say 1 kPa versus 10 kPa, resulting in values of about the same order of magnitude in the denominator of the rate constant expression. Therefore, rate expression (C.4) must be used, which means a fractional order in N2O (e.g., the found 0.78) and the observed activation energy (96 kJ/mol) would have to be between the two extremes (analogous to Eq. (47), note that reaction order in O2 = - O):
E a 2 + H N 2 O <
obs Ea = E a 2 + ( 1 N 2O )H 1 + 1 H 3 2 O
<
E a 2 + H 1 + 1 H 3 2
(C.I.5)
101
<
obs Ea
<
147
In fact, the observed activation energy is slightly lower than but close to the theoretical minimum, which implies low surface coverages (the contribution of O2 desorption must be small). The result is reasonable and within experimental error.
C.I.3
An example of a reaction in which two steps are rate determining is the decomposition of N2O over Co-, Cu-, and Fe-ZSM-5 zeolites. The overall reaction is the same as above, but the kinetic model is different due to the different catalyst properties. On similar zeolitic catalysts, it was found that no oxygen inhibition occurred (in contrast with the situation on the oxide catalysts presented in Sections C.I.1 and C.I.2) and that N2O exhibited first order behaviour [18-21]. Therefore, the model is believed to consist of two irreversible reaction steps that are both rate determining:
C3 - 3
Case study I
1. N2O + * N2 + O* 2. N2O + O* N2 + O2 + * Note that the rates of step 1 and step 2 must be equal because O* is part of a series reaction. However, two molecules of N2O are converted in every catalytic cycle, and therefore rN 2O = N 2O r = 2r = 2r1 = 2r2 resulting in equation (C.I.7) for the total N2O conversion rate. (C.I.6)
rN 2O =
2 k 1 N T p N 2O 1+ k 1 / k 2
(C.I.7)
The ratio k1 / k2 in equation C.I.7 equals O* / * and hence determines the state of the active sites. If k1 / k2 >> 1 the difficult step is step 2 and the sites are covered with oxygen, while for k1 / k2 << 1 step 1 is more difficult. Figure C.I.2 shows the effect of oxygen on the N2O conversion [18], which is a measure of the reaction rate (at constant space time W/F).
1.0 0.8
743 K 833 K
X(N2O)
793 K
10
p(O2) / kPa
Figure C.I.2 Effect of O2 on N2O conversion at 0.1 kPa N2O and W / FN 2O = 2.87 x 105 g s/mol.
QUESTION:
Based on Figure C.I.2 can you decide which is the most and which is the least active catalyst?
Indeed, oxygen hardly or not inhibits N2O decomposition on Co- and Fe-ZSM-5, but its effect is appreciable on Cu-ZSM-5. Proof has been found that the adsorption of O2 on this catalyst is molecular rather than dissociative [18]. Therefore, for this catalyst an additional step is incorporated in the model:
C3 - 4
Case study I
(C.I.8)
Addition of CO (a component often encountered in N2O streams) to the reaction mixture, significantly alters the picture. Figure C.I.3 presents the influence of CO on the conversion of N2O (rate of decomposition).
1.0
Cu-ZSM-5 (673 K)
0.8
X(N2O)
0.6
Fe-ZSM-5 (673 K)
0.4 0.2 0.0 0.0
Co-ZSM-5 (693 K)
0.5 1.0 1.5 2.0
Figure C.I.3 Effect of CO on N2O conversion at 0.1 kPa N2O and W / FN 2O = 2.87 x 105 g s/mol.
When comparing the conversion without CO addition and at a CO/N2O ratio of 2, an increase is observed on all catalysts. Particularly on Fe-ZSM-5, the conversion enhancement is remarkable - from zero to 0.7. The effect of CO can be explained by the irreversible reaction of CO with adsorbed atomic oxygen, formed in step 1: 4. CO + O* CO2 + * Thus, CO enhances oxygen removal. On all catalysts step 4 is much faster than step 2 (as proven by fitting of the experimental data [18] and demonstrated by Figure C.I.3). Therefore, we can assume that oxygen is solely removed by CO and not by N2O (step 2 and 4 are parallel reactions with respect to O*). The kinetic model in the presence of CO is now determined by step 1 and step 4 (and step 3 on Cu). In the reasonable approximation that step 4 is much faster than step 1 equation (C.I.9) results for the N2O decomposition rate on all catalysts: rN 2O = k 1 N T p N 2O (C.I.9) Note that the factor 2 is missing from equation (C.I.9) since step 2 is not involved anymore.
C3 - 5
Case study I
In fact, on the Fe catalyst, k1 and k4 have the same order of magnitude. Derive the rate equation for N2O decomposition for this situation. In which case does this equation reduce to Eq. (C.I.9)?
The fact that without CO the conversion of N2O is much lower on Fe-ZSM-5 also shows that step 2 on this catalyst is much more difficult than step 1 (k1 / k2 >> 1), and thus equation (C.I.7) simplifies to: Fe catalyst, no CO: rN 2O = 2 k 2 N T p N 2O (C.I.10)
On Cu-ZSM-5, the conversion shows a maximum as a function of the CO/N2O ratio. This indicates that at higher CO partial pressures N2O decomposition is hindered by CO. Therefore, another reaction must be taken into account, namely the adsorption of CO: 5. CO + * CO*
The resulting N2O decomposition rate on Cu-ZSM-5 in the presence of excess CO is:
rN 2O = k 1 N T p N 2O 1+ K CO pCO
(C.I.11)
Tables C.I.1 and C.I.2 present a summary of the results, together with values for the apparent activation energies, without and with CO addition, respectively.
obs Table C.I.1 Rate expressions, E a , and contributions for N2O decomposition on M-ZSM-5
M Co Cu
Rate expression rN 2O =
rN 2O =
obs Ea (kJ/mol)
2 k 1 N T p N 2O 1+ k1 / k 2
2 k 1 N T p N 2O 1 + k 1 / k 2 + K O2 p O2
104 136
(1 O )E a1 + O E a 2
(1 O )E a1 + O E a 2 O H O
2 2
Fe
rN 2O = 2 k 2 N T p N 2O
168
Ea2
obs Table C.I.2 Rate expressions, E a , and contributions for N2O decomposition on M-ZSM-5
C3 - 6
Case study I
obs Contributions to E a (Eq. 40)
122 179
rN 2O = k 1 N T p N 2O
QUESTIONs: Why does the term K O2 p O2 not appear in the rate equation on Cu when CO is present? Explain why the partial pressure of CO does not appear in any of the rate equations of Table C.I.2 (except in the denominator with Cu). How can the parameters k1NT, k2NT, k1 / k2, K O2 , and KCO be determined? and
obs Ea ? obs Compare the values of E a on the different catalysts and at the different CO/N2O
C3 - 7
Case study II
C.II.1 Hydrodenitrogenation
Quinone (Q) is a typical nitrogen-containing compound present in oil fractions, and it is often used as a model compound in kinetic studies. Its denitrogenation over NiMo/Al2O3 can be represented by the following kinetic scheme [23].
N Q
N THQ1
NH2 OPA PB
N THQ5
N DHQ
PCH
Scheme C.II.1
Elucidation of the kinetics of this reaction is rather complex. Therefore, a subscheme was considered, viz., the denitrogenation of OPA, Scheme C.II.2 [22]. One pathway leads to the formation of propylbenzene (PB) and ammonia by C-N bond cleavage. The other pathway involves hydrogenation of the aromatic ring to propylcyclohexylamine (PCHA), followed by elimination of ammonia to propylcyclohexene (PCHE), and subsequent hydrogenation to propylcyclohexane (PCH). C3 - 8
Case study II
NH2 OPA PB
NH2 OPA k6 k3 k5 PB
Scheme C.II.2
Scheme C.II.3
The intermediate PCHA is not observed in the product mixture, and hence is not considered in the global kinetic scheme (Scheme C.II.3). A direct step from OPA to PCH occurs in the kinetic scheme, although its interpretation is not directly obvious. This step has to be included based on the product composition as a function of space time [22], as shown in Figure C.II.1.
OPA NiMo one site model 370C
1.0 5
0.6
PCH PCHE PB
0.4
0.2
0.0 0 10 20 30 40 50 60
Figure C.II.1 Product composition versus space time in the HDN of OPA over NiMo/Al2O3 at 623 K and 3.0 MPa.
Based on a sequential reaction scheme, OPA PCHE PCH, the curve for PCH would initially be much flatter than observed. Therefore, it must be formed also via another route. As PCHE obviously is the natural intermediate this result is surprising at first sight. However, an explanation can be found realizing that the only conclusion is that gaseous PCHE is not involved. The situation at the catalyst surface, as shown in Scheme C.II.4 provides the explanation for the apparent direct reaction step. The direct route to PCH in fact involves adsorbed PCHA and PCHE. PCHA is not found in the product because its desorption is slow compared to reaction to PCHE on the surface. PCHE either desorbs or reacts further to PCH.
0.8
OPA
C3 - 9
Case study II
OPA*
kb ka
PCHA*
slow kc
PCHA + *
PCHE*
PCHE + *
ke
PCH + *
Scheme C.II.4
A rate expression (Eq. C.II.1) can be derived for the reaction rate of OPA, based on the usual steadystate assumption, the site balance, and the fact that nitrogen-containing species adsorb more strongly on the catalyst than hydrocarbons. rOPA = (k a + k b )K OPA p OPA 1 + K OPA p OPA + K NH 3 p NH 3 (C.II.1)
Eq. C.II.2 represents the rate expression for the formation of PCH. ka 1 + k /k K OPA p OPA + k e p PCHE a d = 1 + K OPA p OPA + K NH 3 p NH 3
rPCH
(C.II.2)
Ammonia enters the denominator of the rate expressions since it is a product of de denitrogenation reactions, and adsorbs strongly on the catalyst.
C.II.2 Hydrodesulfurization
Hydrodesulfurization (HDS) of gas oils is an important process in oil refineries. The reactivities of the various sulfur-containing compounds differ strongly. In model studies (in which pure components were studied) it appeared that kinetics are in general close to first order. However, in a real feed, although the individual compounds show first order behaviour, second order kinetics are observed (see Figures C.II.2 and C.II.3).
C3 - 10
Case study II
Conversion
Concentration
50
60
Catalyst volume (m )
Figure C.II.2
Second order behaviour in HDS of gas oil; concentration and conversion versus catalyst volume. Data from [24]
Figure C.II.2 shows that the sulfur concentration profile in the gas oil over the bed consists of a fast decrease followed by slow decrease, typical for second order behaviour. Indeed, a second order plot describes the data well, as shown in Figure C.II.3.
12
1 / C S - 1 / C S0 (1 / wt%)
10
Figure C.II.3
Second order behaviour in HDS of gas oil; Effect of space velocity on S removal [24].
Note that CS refers to the concentration of sulfur (in wt%) in the gas oil mixture. This is referred to as lumping. So, all the S-compounds are represented by one pseudo-component. For HDS in practice, the behaviour shown in Figures C.II.2 and C.II.3 means that first the most reactive compounds are desulfurized, followed by compounds that are less reactive. In fact we deal with apparent second order kinetics. Figure C.II.4 shows some typical sulfur compounds that are present in gas oils. Figure C.II.5 shows a gas chromatogram of a gas oil, treated at various temperatures. C3 - 11
Case study II
Thioethers
S
Dibenzothiophene
Thiophene
S S R R
Substituted dibenzothiophene
Benzothiophene
S
S CH3
S CH3
Feed
760 ppm
593 K
240 ppm
613 K
160 ppm
633 K
70 ppm
retention time
Figure C.II.5 Gas chromatograms of pre-hydrotreated gas oil after hydrotreating over NiW/Al2O3 at various temperatures, WHSV = 4.2 (g oil)(g cat)-1 h-1 [25].
C3 - 12
Case study II
The feed had already been hydrotreated to some extent, as can be seen from the absence of thioethers, thiophene, and benzothiophene. These compounds are the most reactive sulfur compounds. The disubstituted sulfur compounds are the least reactive as can be seen from their occurrence even after hydrotreatment at the highest temperature. Since individual sulfur compounds show first order kinetics (as determined by experiments with model compounds), it seems logical to model the HDS kinetics with several different first order reactions with different rate constants. A simulated example of such a model is shown below.
1.8
Simulated model data: Fgas oil = 52400 kg/h Second order: 3 3 -1 -1 -1 k = 10.6*10 (kg oil)(m cat) (wt%S) h CS0 = 1.64 wt% three-lump model, first orders: 3 C01 = 1.00 wt% k1 = 21.6*10 3 C02 = 0.48 wt% k2 = 4.1*10 3 C03 = 0.16 wt% k3 = 0.7*10
3 -1 -1 ki [(kg oil)(m cat) h ]
Sum 1 2 3
Catalyst volume (m )
Figure C.II.6 HDS: Simulated concentration profiles, three-lump model, first order reactions.
The symbols are simulated experimental data, obtained for the second-order reaction of Figure C.II.2 and C.II.3. The numbered curves represent first order reactions of lumps of sulfur-containing compounds with similar reactivities (e.g., 1: thioethers, 2: thiophene + benzothiophene, 3: others). The sum of these three curves is the curve through the symbols. So, a three-lump model consisting of first-order reactions fits an apparent second-order reaction well. Which compounds have been lumped together must be examined by carrying out model studies. QUESTION: Compare the reactor volumes needed (fixed-bed trickle-flow reactor) to obtain 95% conversion using the second-order model and the three-lump model. Do the same for 99% conversion.
C3 - 13
In the aromatization of light alkanes over a zeolite catalysts, cracking, which is catalysed by acid catalysts, yields high amounts of adsorbed hydrogen H*. The surface concentration of adsorbed hydrogen is not in equilibrium with gaseous H2 but much higher. Therefore, hydrogenation reactions are favoured and the selectivity for aromatics is low. However, if Ga is added to the zeolite, an escape route is provided for H*, i.e., Ga provides catalytic sites which enhance the desorption of hydrogen: zeolite: Ga: alkane 2 H* + ... 2 H* H2
So, in this case kinetic coupling is used to increase the reaction selectivity for aromatics.
C.III.2 Parallel reactions; kinetic coupling between catalytic cycles
A classic example of kinetic coupling is the hydrogenation in parallel of two different alkylaromatic compounds:
A1 A2
1 2
Separately, they would each cover the catalyst surface and react with a rate proportional to the rate constant k1 or k2. However, in simultaneous hydrogenation, both compounds compete for the same surface and their rates of hydrogenation will also depend on the adsorption constants K1 and K2. The relative rates of hydrogenation are proportional to selectivity for hydrogenation of A1 compared to A2 which is defined as:
S=
k1 K 1 k2 K 2
Prove equation C.III.1 for a LHHW kinetic scheme.
(C.III.1)
QUESTION:
C3 - 14
Often, the desired product is an intermediate in a series reaction network. Examples are the hydrogenation of alkynes to alkenes, which could react further to alkanes, and partial oxidation reactions. It is often difficult to obtain high selectivity for the desired intermediate product at high conversion.
Hydrogenation of alkynes
In the catalytic hydrogenation of butyne to butene: k1 k2 C-CC-C C-C=C-C C-C-C-C a high butene selectivity can be obtained even when k2 > k1, because butyne adsorbs much more strongly than butene: KBY > KBE. The reason is that the two reactions are kinetically coupled as both butyne and butene compete for the same catalyst sites. The selectivity for butene in this series reaction depends on the relative reaction rates of butyne and butene, which in turn depend on the rate constants, adsorption constants, and concentrations as follows:
p kK p rBY = S BE BY 1 = 1 BY BY 1 p BE k 2 K BE p BE rBE
while the yields on butene and butane depend on the conversion level. QUESTION: Prove equation C.III.2 for a kinetic scheme based on LHHW kinetics.
(C.III.2)
Meyer and Burwell [26] obtained the following product distribution (mol%):
2-butyne cis-2-butene trans-2-butene t-butene butane 22.0 77.2 0.7 0.0 0.1
So, at a conversion of 88 mol%, the selectivity for butenes is nearly 100%, mainly do to the stronger adsorption of butyne on the catalytic sites. A similar example, one that is of great industrial importance, is the hydrogenation of acetylene (C2H2). Acetylene is present in small amounts in the C2 product stream of ethane and naphtha crackers. Ethene is an important monomer and the polymerisation of ethene requires that the amount of acetylene is reduced to levels as low as 0.5 ppm. Acetylene adsorbs more strongly on the nobleC3 - 15
metal hydrogenation catalyst than ethene, but ethene is present in much larger quantities, so a substantial loss of ethene to ethane can be expected. In order to decrease the ethene hydrogenation rate, CO is added to the feed. The adsorption strength of CO is intermediate between that of acetylene and ethene, Kacetylene > KCO > Kethene and hence the hydrogenation of ethene is suppressed. QUESTION: The selective hydrogenation of acetylene in the C2 stream is carried out in a packedbed reactor. Guess the surface coverage of the catalyst with acetylene, ethene, and CO as a function of the axial coordinate.
Catalytic Reforming
The isomerization of n-pentane involves a three-step sequence consisting of dehydrogenation, isomerization, and hydrogenation. The dehydrogenation and hydrogenation steps occur on platinum sites; the isomerization step occurs on acidic sites as shown in Scheme C.III.1. Pt-function: n-C5 n-C5=
Surface diffusion Acid function: n-C5= i-C5= Surface diffusion Pt-function: i-C5= i-C5
Scheme C.III.1
So, a bifunctional catalyst is used, on which different catalytic cycles take place on different sites. Due to the presence of hydrogen in the gas phase (10 to 30 bar) to prevent coke formation as a result of overdehydrogenation, the concentrations of the alkenes will be low. Therefore, for isomerization of the alkenes to take place both catalytic functions should be in close proximity. Generally, the isomerizations will be the slowest step, while the (de)hydrogenation steps are at quasi-equilibrium (see for instance Froment and Bischoff [4]).
C3 - 16
Case study IV
Until recently, chlorofluorocarbons (CFCs) were produced for use in amongst others refrigeration applications. Technically, they possess very favourable properties, and, moreover, they are low-priced chemicals. However, it is now generally accepted [27] that fully halogenated CFCs are responsible for the depletion of the ozone layer, and that they contribute to the greenhouse effect. World-wide, production and consumption of CFCs is being terminated, but considerable amounts are still present. Recovery and subsequent destruction of these substances is a logical step. Many destruction techniques have been proposed [28-33], but only combustion has been applied on a commercial scale. Obviously, the conversion of CFCs into valuable chemicals is much more desirable. At Delft University of Technology, a catalytic process has been developed in which CCl2F2 (CFC-12) is converted into CH2F2 (HFC-32), which is an ozone-friendly refrigerant [34]. CCl2F2 belongs to the family of halogenated methanes. Figure C.IV.1 [35] shows all the possible C1 derivates containing Cl, F, and H. The arrows indicate the thermodynamic stability (e.g. CCl4 least stable, CH4 most stable), and not necessarily a reaction sequence.
CCl4
10
CCl3F
11
CHCl3
20
CCl2F2
12
CHCl2F
21
CH2Cl2
30
CClF3
13
CHClF2
22
CH2ClF
31
CH3Cl
40
CF4
14
CHF3
23
CH2F2 C2H6
170 32
CH3F C3H8
290 41
CH4
50
Figure C.IV.1 Thermodynamic relation between chlorinated and fluorinated methanes. The direction of the arrows indicates the thermodynamic stability at 298 K and atmospheric pressure [35].
All hydrogenolysis reactions starting from CCl2F2 are exothermic irreversible reactions. Although thermodynamically the reaction of CCl2F2 would proceed via CHClF2, CH2ClF, CH3F, and finally methane, a reaction towards CH2F2 might be hoped for (provided that a selective catalyst is available), because the carbon-fluorine bond is much stronger than the carbon-chlorine bond [36]. The reaction to be carried out is the hydrogenolysis of CCl2F2: CCl2F2 + 2 H2 CH2F2 + 2 HCl
Case study IV
The reaction enthalpy shows that the reaction is very exothermic. An important side reaction is the formation of methane, which is even more exothermic: CCl2F2 + 4 H2 CH4 + 2 HCl + 2 HF
C.IV.2
Figure C.IV.2 shows the conversion of CCl2F2 and the selectivities towards the most abundant products as a function of temperature [37]. The main product is CH2F2 but CHClF2 and methane are also formed. Other possible products such as CH2ClF and CH3F are hardly formed. HCl and HF are also formed during hydrogenolysis. HF does not affect the conversion and the selectivity towards CH2F2 [37], but addition of HCl results in lower selectivity towards CH2F2 as demonstrated in Figure C.IV.3. Interestingly, both Figures C.IV.2 and C.IV.3 show that the selectivity towards methane remains nearly constant as a function of conversion, and is not affected by addition of HCl.
conversion/selectivity (mol%)
100
Conversion/selectivity (mol%)
80 70 60 50 40 30 20 10 0
H2 H2/HCl
CCl2F2 Conv.
CH2F2
CHClF2 Selectivity
CH4
temperature (K)
Figure C.IV.2 Conversion (triangles) and selectivity in the hydrogenolysis of CCl2F2 over 1 wt% Pd/C as a function of temperature to CH2F2 (solid circles), CHClF2 (squares), and CH4 (open circles). Conditions: WHSV=1 g/(g h), H2/CCl2F2=3 mol/mol, p=0.3 Mpa [37]. Figure C.IV.3 Influence of HCl on hydrogenolysis reaction. Conditions: WHSV=1 g/(g h), H2/CCl2F2=3 mol/mol, H2/HCl = 1 mol/mol, p=0.2 Mpa [35].
Many possible pathways can be formulated for the formation of the observed reaction products, but the experimental results and thermodynamic data make a plausible choice possible. CCl2F2 is a rather inert molecule. Therefore, the first step in its conversion must be dissociative adsorption. In this step, two surface species can be formed, *CClF2 and *CCl2F. Since CH2F2 is the main reaction product, obviously *CClF2 is formed preferentially, which indicates that the C-Cl bond is much weaker than the C-F bond. This is in accordance with the dissociation energies (at 298 K), which are 318 and 460 kJ/mol, respectively [38]. Furthermore, CClF2 is thermodynamically more stable (see Table C.IV.1, which gives a survey of possible gas-phase reaction intermediates with their Gibbs free energy of formation [39]).
C3 - 18
Case study IV
The adsorbed CClF2 may either react with adsorbed hydrogen to yield CHClF2, loose a fluorine atom to give *CClF, or loose the second Cl atom to form *CF2. CH2F2 being the main product, the formation of *CF2 is most likely, also because it is a relatively stable species as apparent from Table C.IV.1. Its conversion to CH2F2 must proceed via *CHF2, which is even more stable. The effect of HCl on the conversion of CCl2F2 and the relative selectivity towards CH2F2 and CHClF2 can be explained by strong adsorption of HCl and the increased concentration of chlorine on the catalyst surface at higher HCl partial pressure (*CF2 *CClF2 equilibrium shifts to right). Therefore, adsorption of HCl must be taken into account in a kinetic model. Dissociative re-adsorption of CHClF2 and CH2F2 is neglected on the basis of their low reactivity (less than 3% of CCl2F2 reactivity, see Figure C.IV.4 [35]).
100
conversion (mol%)
80 60 40 20 0
CCl2F2
CHClF2
CH2F2
Figure C.IV.4 Comparison of the reactivity of CCl2F2, CHClF2, and CH2F2 in the catalytic hydrogenolysis over 2 g of 2wt% Pd/C. Conditions: T=510 K, H2/CFC=3, p=0.3 MPa, feed=16.5 mmol/h [35].
C3 - 19
Case study IV
The low reactivity of CHClF2 proves that CH2F2 is not formed via a reaction network of the type: CCl2F2
H2
CHClF2
H2
CH2F2
H2
etc.
If this would have been the case, the conversion of CHClF2 would have been of the same order as that of CCl2F2. It is not entirely clear which surface intermediates lead to the formation of methane, but the experimental results clearly indicate that they are different from those leading to CH2F2, i.e. methane is formed via a parallel rather than a sequential pathway. One possible route is through dissociative adsorption of CCl2F2 to *CCl2F, a kind of accident. The further pathway could proceed via *CClF, *CClFH, *CHCl, *CH2Cl, *CH3, and finally CH4. Because of the low stability and hence high reactivity of these intermediates, they will react to the thermodynamically most stable product, methane. Therefore, kinetically only the dissociative adsorption of CCl2F2 to *CCl2F is important for this route. In order to derive a kinetic model, the following assumptions were made by Wiersma [37]: Total number of active sites is constant; All surface species occupy only one catalyst site; All reactions take place at the catalyst surface: No gas phase reactions occur; CHClF2 is formed via reaction of *CClF2 with *H. The possible reaction of *CHF2 with *Cl is neglected for practical reasons; The sequential reaction of *CHF2 to methane is neglected. Methane is formed via route 2 (see Figure C.IV.5); No fluorine is present on the catalyst surface: *F is neglected.
These assumptions together with the above discussion lead to the simplified mechanism of Figure C.IV.5, which includes the most important reaction products and the surface species through which they are most likely formed. The intermediates in methane formation are not important kinetically.
CCl2F2
Route 1 Route 2
CHClF2
CH2F2
CH4
*CClF2 *CCl2F
*CF2
*CHF2
C3 - 20
Case study IV
Because on the basis of the experimental results and gas-phase thermodynamic data, a large number of possible surface species have been eliminated, leaving only CClF2, CF2, CHF2, CCl2F, H, and Cl, the kinetic model can be reduced to the 8 kinetically important equations shown in Table C.IV.2 [37].
Table C.IV.2 Elementary steps in the hydrogenolysis of CCl2F2. Route 1: formation CH2F2 and CHClF2
Dissociative adsorption: CCl2F2 + 2 * * CClF2 + *Cl H2 + 2 * 2 *H Surface reactions: *CClF2 + * *CF2 + *H *CF2 + *Cl *CHF2 + * (3) (4) (1) (2)
Associative desorption: *CHF2 + *H CH2F2 + 2 * *CClF2 + *H CHClF2 + 2 * *Cl + *H HCl + 2 * (5) (6) (7)
CCl2F2 + 2 * *CCl2F + *F
(8)
The adsorbed CCl2F reacts through a series of surface intermediates to *CH3, which desorbs to form methane: *CH3 + *H CH4 + 2* (9)
This step and all the possible intermediate steps are much faster than adsorption of CCl2F2 (equation (8)). Therefore, adsorbed CCl2F reacts immediately, and its surface concentration is negligible. As stated above, adsorbed F is also neglected.
C.IV.2
Rate Expressions
Using only the most abundant reaction intermediates (called mari), it can be concluded that reactions (1) through (8) are kinetically significant. Reactions (1), (4), (5), (6), and (8) are assumed to be rate determining, while reactions (2), (3), and (7) are assumed to be in quasi-equilibrium. The rate C3 - 21
Case study IV
equation is found by using the quasi-equilibrium relationships, the site-balance (equation (C.IV.1)), and coupling of the reaction rates of reactions (1), (4), (5), (6), and (8). 1 = * + Cl + H + CClF2 + CF2 + CHF2 The following relations hold: r = r1 + r8; r4 = r5; r1 = r5 + r6 (C.IV.2) (C.IV.1)
The rate expressions for the formation of CHClF2, CH2F2, and methane are shown in Table C.IV.3. The adsorption term ADS (C.IV.7) can be simplified in several ways, e.g., by neglecting the adsorption of the carbon containing species (ADS2). This is justified when step 5 and/or step 6 are fast, resulting in low surface coverage with CClF2 and CF2. From equations (C.IV.3) and (C.IV.4) it follows that increasing the hydrogen partial pressure enhances the selectivity towards CH2F2, the desired product, while addition of HCl leads to more CHClF2. The combined rate of formation of CH2F2 and CHClF2 is equal to:
r1 =
k1 sN T pCCl2 F2
( ADS )2
(C.IV.9)
Hence, the rate equations predict a constant ratio of formation of CH2F2 and CHClF2 together relative to the formation of methane, consistent with the experimental results (Figure C.IV.2). The kinetic model describes the experimental results well as shown in Figures C.IV.6 and C.IV.7 below, while it also has predictive value [37].
conversion/selectivity (mol%)
conversion/selectivity(mol%)
100 80 60 40 20 0 0 1 2 3 4 5
W (g)
Figure C.IV.6 Prediction of conversion and selectivities as a function of catalyst amount (W) by kinetic model (conditions: T=490, P=0.28 MPa, H2/CCl2F2=10 mol/mol), (measured values: =conv. =sel. CH2F2, =sel. CHClF2, =sel. CH4). Adapted from [37]
temperature (K)
Figure C.IV.7 Prediction of conversion and selectivities as a function of temperature by kinetic model (conditions: P=0.5 MPa, H2/CCl2F2=10 mol/mol, WHSV=1 g/(g.h)), (measured values: =conv. =sel. CH2F2, =sel. CHClF2, =sel. CH4). Adapted from [37]
C3 - 22
Case study IV
This example shows that the use of rate expressions based on kinetic models consisting of elementary steps is very useful. Even though a large number of elementary steps are involved, practical and not too complicated rate expressions result by eliminating unimportant steps in advance.
Table C.5 Reaction rate expressions derived from the kinetic model [37].
CHClF2 production:
(C.IV.3)
CH2F2 production:
(C.IV.4)
with:
S=
k6 k 4 K 3 K7 K 2
(C.IV.5)
CH4 production:
r9 =
k 8 sN T pCCl2 F2
( ADS )2
(C.IV.6)
with:
(C.IV.7)
(C.IV.8)
C3 - 23
Literature
Literature
16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. E.R.S. Winter, J. Catal. 15, 144 (1969). T. Yamashita and A. Vannice, J. Catal. 161, 254 (1996). F. Kapteijn, G. Marbn, J. Rodriguez-Mirasol, and J.A. Moulijn, J. Catal. 167, 256 (1997). J. Leglise, J.O. Petunchi, and W.K. Hall, J. Catal. 86, 392 (1984). C.M. Fu, V.N. Korchak, and W.K. Hall, J. Catal. 68, 166 (1981). G.I. Panov, V.I. Sobolev, and A.S. Kharitonov, J. Mol. Catal. 61, 85 (1990). M. Jian, F. Kapteijn, and R. Prins, J. Catal. 168, 491 (1997). G. Perot, Catal. Today 10, 447 (1991). S.T. Sie, Revue de linstitute Franais du Ptrole 46(4), 501 (1991). H.R. Reinhoudt, unpublished work, (1996). E. F. Meyer and R.L. Burwell, JACS 85 (19), 2877 (1963). W. Brune, Nature 379, 486 (1996). United Nations Environmental Programme, Report of the Ad-Hoc Technical Advisory Committee on ODS Destruction Technologies, May (1992). J. Burdeniuc and R.H. Crabtree, Science 271, 340 (1996). S. Karmaker and H.L. Greene, J. Catal. 151, 394 (1995). S. Imamura, T. Shiomi, S. Ishida, K. Utani, and H. Jindai, Ind. Eng. Chem. Res. 29, 1758 (1990). D. Miyatani, K. Shinoda, T. Nakamura, M. Ohta, and K. Yasuda, Chem. Lett., 795 (1992). S. Okazaki and A. Kurosaki, Chem. Lett., 1901 (1989). Programme for Alternative Fluorocarbon Toxicity Testing, Report of PAFT-V: HFC-32, (september 1992). E.J.A.X. van de Sandt, A. Wiersma, M. Makkee, H. van Bekkum, and J.A., Moulijn, Recl. Trav. Chim. Pays-Bas 115, 505 (1996). J.R. Lacher, A. Kianpour, F. Oeting, and J.D. Park, Trans. Faraday. Soc. 52, 1500 (1956). A. Wiersma, Catalytic hydrogenolysis of CCl2F2 into CH2F2 process development, Ph.D. Thesis Delft University of Technology, 1997. R.C. Weast (Ed.), CRC Handbook of chemistry and physics, CRC Press Inc., Boca Raton, Florida, 64th ed., p. F-189 (1983). I. Barin, Thermochemical data of pure substances 3rd ed., VCH, Weinheim, (1995).
C3 - 24
Literature
The hydrogenation of benzaldehyde to benzyl alcohol is also a typical example of a series reaction in which the intermediate product is the desired one:
Depending on the reaction conditions, benzaldehyde can be hydrogenated to form benzyl alcohol, toluene, hydroxymethylcyclohexane, benzene and methane. However, with suitable reaction conditions and a catalyst a high benzyl alcohol yield can be obtained. This gas-liquid-solid reaction can be carried out in a monolithic reactor, which is an example of a structured catalyst. The monolith is covered with a layer of alumina, the washcoat on which nickel is applied [27]. Good results are obtained, as apparent from Figure C.III.1.
Figure C.III.1 Batch hydrogenation of pure benzaldehyde; composition change versus time (415 K, 20 bar) [27]. Toluene is only formed when all benzaldehyde has been converted. This implies that either the adsorption of benzaldehyde is much stronger than that of benzyl alcohol, or that the reaction rate for the formation of toluene is much lower than that for the formation of benzyl alcohol. Kinetic studies have proven that the latter is the case.
Figure C.III.1 shows that the rate of disappearance of benzaldehyde is independent on its concentration (zero order reaction). Furthermore, the hydrogen pressure did not influence the reaction rate much. This led to the assumption that hydrogen adsorbs on different sites than the hydrocarbons. A kinetic model which can explain the results is the following (neglecting formation of toluene, BALD = benzaldehyde, BALC = benzyl alcohol): 1. BALD + * 2. H2 + 2 #
BALD* 2 H# C3 - 25
Literature
BALC* + 2 # BALC + *
r.d.s.
(C.III.2)
(C.III.3)
(C.III.3)
A similar expression can be derived for the further hydrogenation of benzyl alcohol. The coupling between these two reactions is through the presence of both concentrations in the rate expression; benzaldehyde and benzyl alcohol compete for the same sites. The term containing the hydrogen pressure becomes unity at hydrogen pressures above 10 bar. The strong benzaldehyde adsorption (KBALD/KBALC 2) reduces equation (C.III.3) further to: r3 = k obs (C.III.4)
40. A.E. van Diepen, A.C.J.M. van de Riet, and J.A. Moulijn, Rev. Port. Qum. 3, 23 (1996).
C3 - 26
Catalyst effectiveness
Introduction
(1)
The various aspects that are to be considered to achieve a proper and efficient catalyst testing approach are presented. This approach applies to heterogeneous systems in which the catalyst is the solid phase and the reactants are in the gaseous and/or the liquid phase. The presence of a solid phase introduces complicating phenomena on which this chapter focuses. In this respect homogeneous catalysis is a limiting case and needs no separate treatment. The solid catalyst can be present as either a packed bed of particles, a structural system such as a washcoated monolith, a fluidized bed, an entrained bed, or in a liquid-phase slurry.
4.1 Introduction
Due to the consumption of reactants and the production or consumption of heat, concentration and temperature profiles can develop in the stagnant zone around and in the particle itself (Fig. 4.1).
Gas/solid reactor
Gas film
T c
Bulk gas
T c
Exothermal Endothermal
Bulk liquid
T c
Exothermal
Figure 4.1 Temperature and concentration gradients in and around a catalyst particle for exoand endothermic reaction in gas reaction (left) and for exothermic reaction in gasliquid reaction (right). 4-1
Catalyst Effectiveness
Introduction
In the following sections criteria are derived to ensure that the effect of these gradients on the observed reaction rate is negligible [1, 2, 3]. In gas/liquid/solid slurry reactors the mass transfer between the gas and liquid phase has to be considered, too (see references [4, 5]). In the following sections only catalytic gas-phase reactions will be addressed.
Ap k f ( cb cs ) = V p rv
(2)
This expression shows that the mass transfer rate is proportional to the concentration difference over the film. The observed reaction rate can then be expressed as follows:
rv ,obs = r (cs ) = a' k f (cb cs ) where a' = Ap V p is the specific particle area.
(3)
No transport limitations exist when cs cb. But, how can we determine the concentration at the surface cs? In order to relate cs to observable quantities, a dimensionless number Ca, the Carberry number [6], is introduced as follows:
Ca =
rv ,obs a' k f cb
cb cs cb
(4)
in which akf cb is the maximum mass transfer rate, which is obtained when the surface concentration equals zero. Ca relates the concentration difference over the film to procurable quantities and is therefore a so-called observable [1]. A criterion for the absence of extraparticle gradients in the rate data can be derived from the definition of an effectiveness factor for a particle. This should not deviate more than 5% from unity as criterion:
e =
r ( c ,T ) observed reaction rate n = v ,obs s s = (1 Ca ) = 1 0.05 rate at bulk fluid conditions rv ,chem ( cb ,Tb )
(5)
The - sign applies to positive reaction orders and endothermic reactions, the + sign to negative reaction orders and exothermic reactions. 4-2
Catalyst effectiveness
Extraparticle gradients
QUESTION:
Derive the equality e = (1 Ca)n. What are the physical meanings of the cases in which Ca approaches its limits, i.e., 0 and 1.
For an isothermal, n-th order irreversible reaction this results in the following criterion: Ca < 0.05 n (6)
Figure 4.2 shows the dependence of the particle effectiveness factor on the Carberry number and the reaction order.
10
-1
1
e
0.5 n= 1 2
0.1
0.01 0.001
0.01
0.1
Ca
Figure 4.2 Particle effectiveness factor versus Carberry number for various reaction orders (rv = kv cn).
Another dimensionless group which is often encountered in relation with mass transfer limitations is the Damkhler number (Da):
Da =
rate at bulk fluid conditions rv (cb , Tb ) = maximum mass transfer rate a ' k f cb
The Damkhler is not an observable but it can be related to the Carberry number as follows:
Ca =
rv ,obs a ' k f cb
rv (cb , Tb )
rv ,obs
rv (cb , Tb ) Da = e Da = a ' k f cb 1 + Da
(7)
4-3
Catalyst Effectiveness
Extraparticle gradients
(8)
Te ,max Ts Tb k f c b ( H r ) c b c s = = e Ca = Ca Tb h Tb cb Tb
(9)
where e is called the external Prater number. It represents the maximum relative temperature difference over the film or the ratio of the maximum heat production and heat transfer rates. Thus, temperature and concentration are coupled through e . A criterion for the absence of external heat transfer limitation can now be expressed based on the nonisothermal particle effectiveness factor as follows:
r ( c ,T ) k ( T ) f ( c s ) k v ( Ts ) c s e = v s s = v s = = 1 0.05 rv ( cb ,Tb ) k v ( Tb ) f ( cb ) k v ( Tb ) cb with:
n
(10)
(11)
e = (1 Ca )n exp b
1 = 1 0.05 1 + e Ca 1
(12)
Figure 4.3 shows the influence of the external Prater number e on the relationship between Ca and e for a first order reaction.
100
= 20 n=1
10
0.1 0.001
0.01
0.1
Ca
Figure 4.3 Nonisothermal particle effectiveness factor versus Ca at various values for e.
4-4
Catalyst effectiveness
Extraparticle gradients
It can be seen from Figure 4.3 that the concentration gradient is amplified for e < 0, i.e., in case of an endothermal reaction, whereas it is reversed for e > 0, i.e., when the reaction is exothermal. QUESTION: Explain the above observation.
Under the assumption that heat effects dominate the transport disguises, i.e., for small Ca values, the exponential dominates. Series expansion of this exponential leads to a simple result:
b e Ca =
(13)
A result like this is rather logical since it contains the three groups that determine the overall process. Ca determines the concentration drop over the film, the Prater number the maximum temperature rise or drop (exothermic or endothermic reaction) and the dimensionless activation energy expresses the sensitivity of the reaction towards a temperature change. Figure 4.4 shows that a temperature difference of only a few degrees between the bulk fluid and the catalyst is already critical for the value of the reaction rate constant.
kv(Ts)/kv(Tb)
1.5 1.4 1.3
40
Criterion 0.05
0 2 4 6 8 10
Te / K
Figure 4.4 Effect of temperature difference, Te = Ts Tb, on the reaction rate constant at different values for the activation energy Ea; Eq. (11).
4-5
Catalyst Effectiveness
Extraparticle gradients
Table 4.1 Correlation to calculate mass and heat transfer coefficients in packed beds [7-10] and monoliths [11-13]. Packed beds
Mass transfer
Sh = Sc = kf dp Dif
Heat transfer
Range of validity
Nu =
Pr =
hdp
f