You are on page 1of 221

st316

Industrial Catalysis CPT Prof. Dr. J.A. Moulijn Prof. Dr. F. Kapteijn Drs. A.E. van Diepen, MTD Dr. Ir. M.T. Kreutzer

Organic Chemistry and Catalysis Dr. J.C. Jansen

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

Equations to be known Problems

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

Minimum cost of overall process

Reactants

Reactor
Process requirements Maximum selectivity maximum conversion ease of scale-up high throughput low pressure drop .

Desired products Undesired products Unconverted reactants


Intrinsically safe WRAP . Environment & Safety

Figure 1.1 Aspects of reactor selection (after [1]).

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.

1.1 Reactors for solid-catalyzed reactions


Heterogeneously catalyzed reactions are two-, three-, or even more than three-phase operations. Solid catalyst and gaseous and liquid reactants are brought in contact to achieve the desired conversion. Some of the reactor types that are used are briefly presented here for background information with generalized remarks on their advantages and disadvantages. Gas-solid reactors are the most well-known two-phase catalytic reactor types operated continuously. The major reactors for solid-catalyzed gas phase reactions are the fixed-bed, the fluid-bed and the entrained-flow reactor. The fixed-bed reactor is either operated adiabatically (see Figure 1.2) or isothermally. In the latter case a multitubular reactor system is often chosen where the parallel reactor tubes are surrounded by a heat-exchanging fluid. The fixed character of the catalyst bed necessitates a long catalyst life. If deactivation occurs it should be possible to regenerate the catalyst inside the reactor.
Fluid (gas or liquid)

Inert beads Catalyst

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

Reactors for solid-catalyzed reactions

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

Figure 1.3 Fluid-bed reactor.

Figure 1.4 Entrained-flow reactor.

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

Reactors for solid-catalyzed reactions

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

Wire mesh Catalyst beads or granules

Schematic view of a vertical cross-section

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).

film, pocket, filament

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

Reactors for solid-catalyzed reactions

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.

few large bubbles small particles


Figure 1.7 Slurry reactor. The use of small particles results in high catalyst utilization, but the particles are subjected to attrition. Filtration is an energy intensive aftertreatment, therefore an optimum particle size is sought with respect to catalyst effectiveness, settling and mixing speed. In these well-stirred reactors a good heat supply or removal is achieved. Main advantages of the slurry reactor are the high catalyst utilization, the uniform temperature, and good external mass transfer. Disadvantages are catalyst attrition, the need for catalyst separation, and

1-5

Introduction

Reactors for solid-catalyzed reactions

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

0.1-0.5 0.2-0.8 0.05-0.2 0.1-5 500-1000 100-1000 <1

1-6

Introduction

Reactors for solid-catalyzed reactions

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

Cost Heat exchange Design difficulties

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

Reactors for solid-catalyzed reactions

1.2 Reactor design equations basic definitions


The derivation of the continuity equations or design equations for the ideal reactor types, viz. batch reactor, PFR (plug-flow reactor), and CSTR (continuous stirred-tank reactor), using mass balance equations is assumed to be basic knowledge. For the sake of uniformity and introduction of the symbols used, the integral equations for a single reaction in an isothermal reactor under steady-state conditions are presented in Figure 1.9.

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

Reactor design equations

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

DIFFUSION REACTION TRANSPORT PHENOMENA

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

m3gas m-3r m3liquid m-3r m3cat m-3r s -

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.

2.2 Catalyst Shape and Size in Relation to Reactor Type


Solid catalysts may be present in the reactor as a packed bed of particles (fixed-bed reactor, movingbed reactor), a flowing powder (slurry reactor, fluidized-bed reactor, entrained-flow reactor) or as a preformed matrix, i.e., a structured catalyst (monoliths). Figure 2.1 shows some catalyst shapes applied in industry.

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

Catalyst shape and size

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.

2.3 Classification of Solid Catalysts


Solid catalysts can be divided in bulk catalysts and supported catalysts. Examples of bulk catalysts are Raney nickel and the iron catalyst used in ammonia synthesis. Supported catalysts are used much more frequently in heterogeneous catalysis, especially when expensive metals are used. The function of the support is to maximize the dispersion of the active phase by providing a large surface area on which the active phase can be applied. Typical support materials are the acidic supports, viz., silica, alumina, and zeolites, and activated carbon. Sometimes the support also shows catalytic activity. For instance, the acidic supports are all active for the formation of carbenium ions to a greater or lesser extent. Examples of supported catalysts are CoO-MoO3/alumina, Pt/alumina, Pd/carbon, V2O5/TiO2, and Ni/monolith. 2-2

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.

2.4 Preparation of Bulk Metal Catalysts

2-3

Solid Catalysts

Preparation of bulk metal 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.

2.4.1 Raney Nickel


Raney nickel is the most important of the Raney catalysts (other Raney catalysts Co, Cu, Ru, Ag). It is extensively used in hydrogenation reactions in fine chemistry. Raney nickel is prepared by leaching an alloy of nickel with aluminum (obtained by fusion) in a concentrated alkaline solution. Most of the aluminum is dissolved and thus leaves a skeleton of nickel, with high surface area. The remaining aluminum probably has an essential function in stabilizing the small nickel particles. Although Raney nickel is very active for hydrogenations, it has the disadvantages that it is sensitive to poisoning and not very stable at high temperature. Furthermore, regeneration of the deactivated catalyst is not always possible. Often, the catalyst has to be melted with aluminum until the right NiAl alloy is formed again, after which the preparation procedure is repeated. Another important disadvantage, particularly for large-scale applications, is that the preparation procedure is not very reproducible.

2.4.2 Ammonia Synthesis Catalyst


Ammonia synthesis catalysts are prepared by fusion of a mixture of Fe3O4 and small amounts of alumina and K2CO3. Common shaping techniques (extrusion, pelletizing, etc.) are not practical with fused catalysts. Therefore, it is common practice to cool the melt, followed by crushing and size grading of the granules. Then the catalyst is reduced to iron metal in hydrogen. The resulting catalyst is highly porous and mechanically very strong. The reduced catalyst is pyrophoric. Therefore, the catalyst is either used in oxidic form with reduction carried out in situ in the ammonia synthesis reactor, or the reduced catalyst is passivated with nitrogendiluted air. A thin layer of iron oxide is formed, which protects the underlying iron layer from further oxidation. This layer is easily reduced in the ammonia synthesis reactor. The presence of alumina is essential, since it protects the catalyst against sintering. Additives such as CaCO3 are also used to enhance the catalyst performance. Due to the high iron content of the catalyst, the surface area is large.

2.5 Catalyst Supports


Supported catalysts are much more common in industrial processes than bulk catalyst. The support is the carrier for the active phase, it stabilizes the active phase, and provides access to the active 2-4

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.

2.5.1 Choice of Catalyst Support


The choice of the support depends on several factors. The first requirement is that it must be stable at reaction conditions. Many materials are not resistant against high temperatures. An example is the transformation of highly porous -alumina into -alumina (surface area typically 4 m2/g) at temperatures exceeding 1423 K. Activated carbon has high thermal stability, but is oxidized in an oxygen containing environment. Acidic or basic solutions cause many oxidic supports to dissolve. Additionally, catalysts requiring regeneration must be based on supports that are stable at regeneration conditions. For instance, hydrotreating and FCC catalysts must be stable in an oxidative environment at high temperature. The texture of the catalyst support, i.e., size and shape of pores, connection of pores, etc., influences the mass- and heat-transfer characteristics. Pores that are too narrow or too long will reduce the selectivity in partial oxidation processes, because the sequential reactions towards complete oxidation are favoured. In this respect, zeolites are a special case. Due to their pores of molecular dimensions, shape-selective catalysis is possible. The interaction between support and active phase must be optimal. One of the functions of the support is to stabilize the active phase against sintering. In this respect, strong interaction is favourable. However, the stronger the interaction, the more the chemical characteristics of the active phase will be modified. Usually this is not desired. Also, the support may be inert for a certain reaction or possess catalytic activity. Obviously, the cost of the support must be as low as possible. Other properties of interest are density (e.g., for separation by settling), heat capacity, and pellet shape, size, and. Most support materials are available in several shapes and sizes.

2-5

Solid Catalysts

Catalyst supports

2.5.2 Comparison and Uses of Supports


Table 2.2 summarizes the main characteristics of some support materials. Silica and alumina are the most common supports. Alumina is available in many forms of which a high surface area (-Al2O3) and a low surface area (-Al2O3) type are shown. Table 2.2 Characteristics of some support materials. Pore volume Material BET Surface area 2 -1 (cm3g-1) (m g ) 0.6 180 200 -Al2O3 3 5 -Al2O3 0.5 1.2 150 400 SiO2 0.2 0.7 130 160 SiO2-Al2O3 0.34 50 TiO2 (anatase) 0.1 - 2 200 1500 C (activated carbon) 0.2 0.2 0.6 400 700 Zeolites

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

Figure 2.2 Silica and silica-alumina.

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.

Brnsted acidity framework throughout the crystal H O O O Si Al O O O O O Si O O

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

luboils flashed 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.

2.5.3.2 Structure of the Framework and Configuration of the Pores

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:

Is the classic definition of Brnsted acidity site applicable to zeolites?

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

Structure code LTA FAV MOR BEA MFI

Mordenite >5 Beta MFI >5 >10

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

10-6 10-8 Configurational 10-10 10-12 10


-14

.1

1 10 100 zeolite meso macro

1000

pore size [nm] pore

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.

Table 2.6 Trends in zeolite properties as a function of the Si/Al ratio.

Si/Al

infinity [all silica]

hydrophilic hydrophobic n Brnsted sites acid strength thermal stability resistance to acid
900 K 1400 K

adsorption [cm /g]

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 Preparation of Supported Catalysts


Supported catalysts usually are manufactured by two basic methods, i.e., precipitation in which the active phase and support are made together, or impregnation of a preformed support with an active phase.

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

Outer Egg shell Active phase

Middle Egg white

Inner Egg yolk

Support

Figure 2.10 Possible active phase distributions in a catalyst pellet.


The choice of the active phase distribution depends on factors such as the kinetics of the main and side reactions, the rate of diffusion inside the catalyst particle, and the poisoning mechanism (see Chapter 5). If, for instance, diffusion is slow compared to the rate of reaction, the inner part of the catalyst will not be used. Hence, catalytic material may be saved by leaving the inner catalyst part free from it, without losing catalyst activity. This is the so-called egg-shell catalyst. See also references [4, 5]. QUESTION: When would egg-white and egg-yolk catalysts be advantageous to use? The preparation of catalysts with a defined activity distribution requires special preparation methods. For instance, to produce and egg-white catalyst, a substance that adsorbs stronger on the support than the 2 - 16

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

1997 880 1600 1500 750 900

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

Catalytic Reaction Kinetics

Introduction

3 CATALYTIC REACTION KINETICS 3.1 Introduction


Rate expressions are indispensable in the application of catalyzed reactions, in the design of chemical processes, in particular reactor design, and in process start-up and control. Insight in the dependence of the reaction rate on catalyst variables, temperature, and concentrations of reactants, products, and other relevant species is required to predict the size of a catalytic reactor and the optimum operating conditions. Catalytic reaction pathways consist of reaction sequences formed by a series of elementary reaction steps. Hence, rate expressions are in general a function of many parameters, as depicted in equation (1). r = f (catalyst, T, pi or ci,. . . , ki . . . , Ki . . . , Keq) (1)

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

Catalytic Reaction Kinetics

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.

3.2 Rate expression (single-site model)


Consider the reversible reaction A B (e.g., an isomerization reaction), which proceeds according to the three elemental steps represented in Figure 3.1.
Elementary processes A B

1
A*

3 2
B*

Langmuir adsorption

Figure 3.1 Schematic representation of the reaction A B.

1. 2. 3.

A + *
A*

k1 k-1 k2
k-2

A*
B*

k3

B*

k-3

B + *

This kinetic model can be interpreted as follows: 3-2

Catalytic Reaction Kinetics

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

(2) (3) (4)

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

Catalytic Reaction Kinetics

Rate determining step

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.

3.3 Rate-determining Step - Quasi Equilibrium


The aforementioned reaction A B proceeds through three elementary processes in series. Consequently, the net rates of the individual steps are equal to the overall rate (Eq. 8). It can be imagined that the forward and backward rates of two of these steps are large compared to the third one. They might be so large that these steps can be considered to be in quasi-equilibrium. The third step is then called the rate-determining step (r.d.s.). Figure 3.2 depicts the situation in which the surface reaction is rate determining, while the adsorption and desorption steps are in quasiequilibrium.
r1 r-1 r2 r-2 rate determining r3 r-3 quasi-equilibrium r

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

Catalytic Reaction Kinetics


K3 = k 3 * p B = B k 3

Rate determining step

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)

Similar equations can be derived in case adsorption is the rate-determining step:

r=

N T k1 ( p A p B / K eq ) 1 + (1 + 1 / K 2 ) p B / K 3

(15)

or when desorption is rate determining: r= N T k 3 K1 K 2 ( p A p B / K eq ) 1 + ( 1 + K 2 ) K1 p A

(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

Catalytic Reaction Kinetics

Adsorption isotherms

3.4 Adsorption Isotherms


Step 1 in Figure 3.1 is an adsorption step. Under equilibrium conditions, the net rate is zero and the equilibrium is then described by equation (11). This does not mean per se that K1 is constant. It may vary with surface occupancy in the case of a non-uniform surface, due to the interaction of adsorbed species with each other. However, in the derivation of rate expressions it is generally assumed that one deals with a homogeneous surface. Under the assumptions that: the surface contains a constant number of identical adsorption sites, a site can contain only one molecule, and, no interaction takes place between adsorbed molecules,

the so-called Langmuir adsorption is operative.

3.4.1

One- component Adsorption

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 (-)

0.6 0.4 0.2 0 0 0.2 0.4 0.6 0.8 1

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

Catalytic Reaction Kinetics


1 = 1 +1 K A pA

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

(c, n constants, n > 1 )

(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

Catalytic Reaction Kinetics

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)

Figure 3.4 Multicomponent adsorption.

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

Catalytic Reaction Kinetics

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).

3.5 Rate Expressions (other Models and Generalizations)


3.5.1 Models involving Multiple Sites

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

Catalytic Reaction Kinetics


2 k2 sNT K A ( p A pB / K eq )

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

Models for Complex Kinetic Reaction Schemes

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

Catalytic Reaction Kinetics A* B + * with overall equilibrium constant K

Rate expressions other models

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 =

(kinetic factor )(driving force) (adsorption term)n

(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

Catalytic Reaction Kinetics

Limiting cases

3.6 Limiting Cases Reactant and Product Concentrations


In Section 3.2 and 3.4, the approach based on one rate-determining step with the other steps in quasiequilibrium was applied to simplify the derivation of rate expressions. A further simplification of the rate expressions is obtained when the product concentrations or partial pressures are negligibly small, i.e., at low conversions of a pure reactant feed stream. Similarly, at high conversions, when nearly only product is present, the rate expressions can also be simplified. Applying the single-site model, Eqs. 14-16, for the reaction A B at low conversion, the following expressions (often called initial rate expressions are obtained, provided that the feed contains pure A. Adsorption rate determining:

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

Catalytic Reaction Kinetics

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

The dehydrogenation of ethanol over a Cu-Co catalyst C2H5OH A CH3CHO R + H2 S

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

Catalytic Reaction Kinetics

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

Catalytic Reaction Kinetics 2. CO* + * C* + O* (r.d.s.)

Limiting cases

The rate at low conversions can be expressed as: r0 = k 2 sN T CO * = k 2 sN T K CO pCO (31)

(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

Oc 0.4 0.6 cu pa 0.8 nc y( 1.0 -)

600 550 500 450 400

Tem

ture per a

(K)

Figure 3.7 Dependence of CO dissociation rate on temperature and surface occupancy.

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.

3.7 Temperature and Pressure Dependence


Typical behaviour can be expected with respect to the temperature dependence of catalyzed reactions, due to the effect of (competitive) adsorption. This will be demonstrated with some simplified cases of the reaction A B, starting with a consideration of the rate-determining step

3.7.1 Transition-State Theory


According to the transition-state theory molecules react through unstable intermediates called transition-state complexes which then react to products. For instance, the surface reaction A* B*, which is considered to be the rate determing step, proceeds as follows:
k+
#

*A

*A#

kbarrier

*B

k-#

3 - 15

Catalytic Reaction Kinetics

Temperature and pressure dependence

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:

Derive equation (32).

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

Catalytic Reaction Kinetics

Temperature and pressure dependence

A*#
obs Ea = Ea2

Ea2

A*

B*
Figure 3.8 Energy diagram for reaction A B, strong adsorption of A.

3.7.2 Forward Reaction Temperature and Pressure Dependence


For the conversion A B, the overall activation energy is a complex function of the reaction enthalpies and activation energies of the individual elementary reaction steps. We will illustrate this by assuming that the surface reaction is the rate-determining step (Eq. 14) and that the backward reaction can be neglected. The rate expression for this single-site reaction can now be written as:
r= k2 NT K A p A 1 + K A p A + K B pB

(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

Catalytic Reaction Kinetics

Temperature and pressure dependence

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

values have been reported for the same reaction system.

3.7.3 Forward Reaction - Limiting Cases


Based on the previous analysis of the pressure and temperature dependence of the reaction order and the observed activation energy, four different cases can be distinguished:

3.7.3.1 Strong Adsorption of A


Strong adsorption of A means that KApA >> 1 and KBpB, and hence, equation (36) reduces to:
r = k2 NT

(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

concentration of active sites NT is independent of temperature.

3.7.3.2 Weak Adsorption of A and B


When A and B are only weakly adsorbed, KApA and KBpB << 1 and A, B 0, so the rate expression becomes:

3 - 18

Catalytic Reaction Kinetics


r = k2 NT K A p A

Temperature and pressure dependence (43)

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

Figure 3.9 Energy diagram for reaction A B, weak adsorption of A and B.

3.7.3.3 Strong adsorption of B


When the adsorption of B is very strong, so that KBpB >> 1 and KApA, and A 0, B 1, the rate expression becomes:
r= k2 NT K A p A K B pB

(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)

Note that - HB > - HA as a consequence of the stronger adsorption of B. 3 - 19

Catalytic Reaction Kinetics

Temperature and pressure dependence

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.

3.7.3.4 Intermediate Adsorption of A and B


For intermediate values of KApA and KBpB, expressions (39) and (41) apply for the reaction order and the observed activation energy, respectively. The reaction order will range between 0 and 1 for A, and between -1 and 0 for B. The observed overall activation energy will have intermediate values between the two extremes of cases 2 and 3:
E a 2 + H A <
obs Ea = E a 2 + ( 1 A )H A B H B

<

E a 2 + H A H B

(47)

3.7.3.5 Transitions between Limiting Cases


Transitions between the different situations discussed above will occur as a function of temperature, because the occupancies will vary. With increasing temperature the adsorption equilibrium constants decrease resulting in decreased occupancies. Thus, if starting with strongly adsorbing B at low temperature as example, a gradual transition can be envisaged from situation 3 via 4 to 2, during which: - the order of A remains nearly 1 - the order of B changes from -1 to 0, and - the observed overall activation energy will change from (Eq. 46) to (Eq. 44) In addition, large changes in partial pressures can result in changes in the apparent reaction order of a component. At low pA the reaction is first order in A, while at high pA the rate approaches a limit, as can be expected for Langmuir adsorption, and the reaction becomes zero order in A, see equation (39). Summarizing, this means that the overall observed activation energy and the apparent reaction orders of the components depend on the degree of coverage of the active sites, which in turn depends on the temperature and partial pressures. 3 - 20

Catalytic Reaction Kinetics

Temperature and pressure dependence

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)?

Example 3.3 Cracking of Alkanes

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.

3.7.4 Changing Rate-determining Step


Up to now it has been assumed that the rate-determining step remains the same with changing reaction conditions. However, the r.d.s. may change, especially under the influence of temperature [4], but also as a result of pressure changes [2]. Envisage the rate-determining step at low temperature is the desorption step. Generally, this implies that the activation energy barrier for desorption is the highest. A temperature increase will enhance the rate of this step more than the other steps with lower activation energies. Hence, another step can now become

3 - 21

Catalytic Reaction Kinetics

Temperature and pressure dependence

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.

Example 3.4 Dehydrogenation of Methylcyclohexane

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

Catalytic Reaction Kinetics

Temperature and pressure dependence

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.

3.8 Sabatier Principle - Volcano Plot


As apparent from equation (43), the larger the heat of reactant adsorption (larger KA), the larger the overall rate of reaction. A larger heat of adsorption enhances the surface coverage (and changes the reaction order), and consequently, the reaction rate. Often, a relation is found between the overall rate of reaction and the heat of adsorption. However, generally the rate passes through a maximum when the heat of adsorption increases, as demonstrated in Figure 3.13.

Rate

Heat of adsorption

3 - 23

Catalytic Reaction Kinetics

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.

3.9 Concluding Remarks


The preceding sections indicate how a useful approximate reaction rate expression can be derived for catalytic reactions, starting from an assumed kinetic model consisting of elementary reaction steps. The derivation is based on the following assumptions: 3 - 24 the reaction system is in a steady state;

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

Catalyst Performance Testing

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

J mol-1 J mol-1 J mol-1 K-1 -

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

empty/vacant active site

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

Catalytic Reaction Kinetics

Case study I

Chemical Kinetics of Catalytic Reactions - Case Studies Case Study I: N2O Decomposition
C.I.1 Oxide catalyzed N2O decomposition

For the decomposition of N2O over solid oxide catalysts: 2 N2O 2 N2 + O2

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

Catalytic Reaction Kinetics

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)

This rate expression reduces to Eq. (C.I.1) when K1 p N 2O and


0.5

( pO

K 3 ) << 1, e.g., at low partial pressure of N2O


0.5

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

The apparent order in N2O can be expressed as 1 N 2O .

C.I.2

N2O decomposition over Mn2O3, thermodynamic consistency

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

order in N2O 0.78


obs = 96 kJ/mol Ea

638 K
0.1

623 K 608 K 598 K


2.0 4.0 6.0 8.0 10.0

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

Catalytic Reaction Kinetics

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

Ea2 = 130 kJ/mol

H3 S 3

= =

92 kJ/mol 109 J/molK

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

Decomposition of N2O over ZSM-5 (Co, Cu, Fe)

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

Catalytic Reaction Kinetics

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)

0.6 0.4 0.2 0.0

Cu-ZSM-5 Fe-ZSM-5 Co-ZSM-5

793 K

733 K 688K 773 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

Catalytic Reaction Kinetics 3. O2 + * *O2

Case study I

which leads to the following rate of N2O decomposition on Cu-ZSM-5: rN 2O = 2 k 1 N T p N 2O 1+ k 1 / k 2 + K O2 pO2

(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

molar CO/N2O ratio

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

Catalytic Reaction Kinetics QUESTION:

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

catalysts. CO/N2O = 0 (mol/mol)

M Co Cu

Rate expression rN 2O =
rN 2O =

obs Ea (kJ/mol)

obs Contributions to E a (Eq.40)

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

catalysts. CO/N2O = 2 (mol/mol)

C3 - 6

Catalytic Reaction Kinetics M Co Cu Fe Rate expression rN 2O = k 1 N T p N 2O rN 2O = k 1 N T p N 2O 1+ K CO p CO 78 Ea1


obs Ea (kJ/mol)

Case study I
obs Contributions to E a (Eq. 40)

122 179

Ea1 Ea1 - COHCO

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

ratios. Are they logical?

C3 - 7

Catalytic Reaction Kinetics

Case study II

Case Study II: Complex Kinetics


Numerous examples of complex reactions can be found. Many oil refinery processes are notorious for their complex reaction schemes, involving many reversible reactions. Furthermore, many different reactants are present, which can influence the kinetics of others. For instance, the hydrodenitrogenation (HDN) of pure anilines (e.g., orthopropylaniline, OPA) is relatively easy, but their reaction is almost completely inhibited when other basic nitrogen-containing compounds, such as quinones, are present [22]. Section C.II.1 shows how the kinetics of complex reactions can be modeled with a LHHW model. Another example is hydrodesulfurization (HDS). Section C.II.2 shows that although the apparent reaction order of the mixture sulfur compounds in HDS is two, the reaction can be modeled as a combination of first order reactions.

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

NH2 PCHA PCHE

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

Catalytic Reaction Kinetics


k1

Case study II

NH2 OPA PB

NH2 OPA k6 k3 k5 PB

NH2 PCHA PCHE PCH


PCHE PCH

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

Partial pressure (kPa)

0.6

PCH PCHE PB

0.4

0.2

0.0 0 10 20 30 40 50 60

space time (cs)

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.

Partial pressure (kPa)

0.8

OPA

C3 - 9

Catalytic Reaction Kinetics


Competitive parallel steps Direct global routes
OPA + *

Case study II

OPA*

kb ka

HCs not adsorbed (weakly compared to N-s)


PB + *

Fast reaction steps

PCHA*

slow kc

PCHA + *

Only traces found

PCHE*

PCHE + *

kd The direct route to PCH

ke
PCH + *

Other hydrogenation functional sites ?

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

Catalytic Reaction Kinetics


1.8 1.6 1.4 C S (wt%), X (-) 1.2 1 0.8 0.6 0.4 0.2 0 0 10 20 30 40
3

Case study II

Gas oil Co/Mo-alumina 3 mm pellets Fgas oil = 52400 kg/h

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

Gas oil Co/Mo-alumina 3 mm pellets Vcat = 52.4 m3

0 0 0.2 0.4 0.6 0.8 1

1000 / WHSV (m3cat h / kg gas oil)

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

Catalytic Reaction Kinetics


S R R R S S R

Case study II

Thioethers
S

Dibenzothiophene

Thiophene
S S R R

Substituted dibenzothiophene

Benzothiophene
S

Figure C.II.4 Typical sulfur compounds.

S CH3

S CH3

S CH3 C2H5 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

Catalytic Reaction Kinetics

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

Sulfur concentration (wt%)

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 ]

1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 10 20 30 40 50

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

Catalytic Reaction Kinetics

Case study III

Case Study III: Kinetic Coupling


Kinetic coupling is a term to describe the coupling of two or more catalytic cycles on one catalyst, possibly with different sites, often by the competitive adsorption of different reactants or reactants and intermediates.
C.III.1 Aromatization of light alkanes

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

hydrogenated products hydrogenated products

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

Catalytic Reaction Kinetics


C.III.3 Series reactions; kinetic coupling between catalytic cycles

Case study III

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

Catalytic Reaction Kinetics

Case study III

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

Catalytic Reaction Kinetics

Case study IV

Case Study IV: CFC Conversion


C.IV.1 Introduction

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

Hr298 = - 156 kJ/mol


C3 - 17

Catalytic Reaction Kinetics

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

Hr298 = - 319 kJ/mol

C.IV.2

Development of Kinetic Model

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

80 60 40 20 0 400 450 500 550 600

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/(gh), H2/CCl2F2=3 mol/mol, p=0.3 Mpa [37]. Figure C.IV.3 Influence of HCl on hydrogenolysis reaction. Conditions: WHSV=1 g/(gh), 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

Catalytic Reaction Kinetics


Table C.IV.1 Gibbs energy of radicals which could be intermediates in the hydrogenolysis of CCl2F2 [39]. Radical Radical G0 (kJ/mol) G0 (kJ/mol) CF3 -549 CF 192 CClF2 -361 CCl 435 CHF2 -331 CH 540 CCl2F -194 CH2F -103 CHCl2 -9 CCl3 -9 CH2Cl 44 CH3 88 CF2 -254 F 32 CClF -51 Cl 72 CHF 35 H 184 CCl2 159 C 670 CHCl 238 CH2 329

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

Catalytic Reaction Kinetics

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

Figure C.IV.5 Simplified mechanism of CCl2F2 hydrogenolysis [37].

C3 - 20

Catalytic Reaction Kinetics

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)

Route 2: formation of methane

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

Catalytic Reaction Kinetics

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

100 80 60 40 20 0 440 460 480 500 520 540

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

Catalytic Reaction Kinetics

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:

k 1 sN T pCCl2 F2 pH2 1 + 1 S p HCl r6 = ( ADS ) 2

(C.IV.3)

CH2F2 production:

k 1 sN T pCCl2 F2 1 + S p HCl pH2 r5 = ( ADS ) 2

(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:

p HCl ADS = 1 + + K 2 p H 2 + ADS 2 K7 K 2 p H 2


1 + k 5 + K 3 k 5 p HCl k4 k4 K 7 K 2 pH2

(C.IV.7)

k1 p CCl2 F2 ADS 2 = k 5 k 6 K 3 p HCl k5 K 2 pH2 + k4 K7

(C.IV.8)

C3 - 23

Catalytic Reaction Kinetics

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

Catalytic Reaction Kinetics


Hydrogenation of benzaldehyde

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

Catalytic Reaction Kinetics 3. BALD* + 2 H# 4. BALC*

Literature

BALC* + 2 # BALC + *

r.d.s.

With the site balance for hydrocarbon adsorption sites:

1 = * + BALD + BALC and for the hydrogen adsorption sites:


1 = # + H The rate expression for the disappearance of benzaldehyde becomes:
r3 = K H2 pH2 k obs K BALD C BALD (1 + K BALD C BALD + K BALC C BALC ) 1 + K H p H 2 2

(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

4 MASS AND HEAT TRANSPORT CATALYST EFFECTIVENESS


Intrinsic catalytic reactivities and selectivities can be disguised by various phenomena of non-ideal reactor behaviour and mass- and heat transport limitations in the catalyst bed and on the particle level. Ideal plug-flow or ideal mixed-reactor behaviour can be assured if certain guidelines are satisfied. Criteria, based on observed rate data, and diagnostic experimental tests are presented that can be applied to test the absence of mass- and heat transport limitations. In this chapter criteria will be derived that can be used to check whether one operates under conditions that result in deviations of not more than 5% in reaction rate from the ideal situation, i.e.,

rate observed = 1 0.05 rate ideal

(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

Gas/liquid/solid slurry reactor


Liquid film Bulk gas (bubble) 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.

4.2 Extraparticle Gradients


4.2.1 Mass transfer, isothermal
Extraparticle mass and heat transfer are most conveniently analysed by means of the so-called film model, in which the flux from the bulk fluid to the particle is defined in terms of mass and heat transfer coefficients kf and h, respectively. A mass balance over the film layer yields:

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

a' k f (cb cs ) 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

4.2.2 Heat transfer, nonisothermal


Analogously to mass transfer the following equation is used for external heat transfer:
a' h ( Tb Ts ) = rv ,obs ( H r )

(8)

Combination with Eq. (3) yields:

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)

k v ( Ts ) 1 E Tb = exp a 1 1 = exp b k v ( Tb ) 1 + e Ca RTb Ts leading to equation (12):

(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

= 0.5 0.2 0.1

10

0 -0.5 -0.2 -0.1

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 =

E a ( Hr ) k f cb rv ,obs < 0.05 h Tb R Tb k f a 'cb

(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

Tb=500 K Ea(kJ/mol): 120


80

1.2 1.1 1.0

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).

4.2.3 Determination of Mass and Heat Transfer Coefficients


Both criteria for extraparticle gradients contain observables and can be calculated based on experimental observations of reaction rates. For heat and mass transfer coefficients in packed beds various correlations exist in terms of dimensionless numbers. Table 4.1 surveys the most appropriate ones for laboratory reactors [7-10]. Values of kf and h for gases in laboratory systems range between 0.1-10 m s-1 and 100-1000 J K-1 s-1 m-2, respectively. In the case of monoliths other correlations should be used because of the different geometry [11-13].

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
f C pf f

Re p =

f udp f

f f Dif

Gases
Sh = 0.357

Re p 0.641 Sc1 3

Nu =

0.428

Re p 0.641 Pr 1 3

3 < Rep < 2000

Sh 0.07 Re p
Liquids
Sh = 0.250

Nu 0.07 Re p

0.1 < Rep < 10

b
1.09

Re p 0.69 Sc1 3

Nu =

0.300

b
. 131

Re p 0.69 Pr 1 3

55 < Rep < 1500

Sh =

Re p1 3 Sc1 3

Nu =

Re p1 3 Pr 1 3

0.0016 < Rep < 55

Monoliths

Sh = Nu (1 + B1 Re Sc

d H 0.45 ) L d Nu = Nu (1 + B1 Re Pr H ) 0.45 L

where:

Re =

f u dH ; f

dH =

M w
1 M

4 Ach Och

and:

Nu =

2.976 (square) 3.657 (circular) 3.660 (hexagonal) 0.095 (catalytic monoliths)

B1 =

4-6

Catalyst effectiveness

Extraparticle gradients

4.2.4 Multiplicity Multiple Steady States


Sometimes more than one solution can be found to the mass- and heat balance equations. Some examples are given below.
1. Isothermal external

For the CO oxidation over a Pt catalyst with excess oxygen, the following equation for the observed reaction rate can be derived:

rv ,obs = a ' k f (cb c s ) =

(1 + Kc s )2

kv ' cs

(14)

in which K is the adsorption constant of CO, and cb and cs refer to the concentration of CO. QUESTION:
Suggest a kinetic scheme that is in agreement with equation (14).

Figure 4.5 shows the observed reaction rate and the rate of mass transfer as a function of the surface concentration at three different bulk concentrations.

akfcb

reaction rate

supply rate slope: -akf

robs

cs
Figure 4.5 robs and mass transfer rate as a function of the CO surface concentration cs; concentration hysteresis.

Figure 4.5 shows that depending on the bulk concentration one, two, or three solutions to Eq. 14 (steady states) are possible. In the case of three solutions, two stable situations are possible and one unstable situation. This phenomenon is known as a concentration hysteresis. QUESTION:
Which is the unstable situation?

2. Nonisothermal external - theoretical

When both mass- and heat transfer limitations exist, also multiple steady states can occur. Solving of the mass- and heat balance (Eqs. 3 and 8) for a first-order reaction after rearrangement yields: 4-7

Catalyst Effectiveness

Extraparticle gradients

k v (Ts ) 1 Ts Tb = a ' k f + k v (Ts ) e Tb

(range : 0 1)

(15)

in which the left-hand side roughly represents heat production and the right-hand side heat removal. Figure 4.16 shows both as a function of the surface temperature.
1.0 0.8

0.1 0.02

heat removal heat production

LHS, RHS

0.6 0.4 0.2 0.0

0.005

Tb

Ts

Ts,max

Figure 4.6 Temperature hysteresis. Heat production and removal as a function of Ts; LHS (), RHS (- - -) = left-hand side, right-hand side of Eq. 15, b = 20; e = 0.5; n = 1; variation of Da.

QUESTIONs: Rewrite the left-hand side of Eq. (15) in terms of Da, b, and Ts / Tb. How can one vary Da? Why is e not affected? How can e be varied without affecting Da? For operation conditions at which the rate of reaction is large compared to the maximum mass transfer rate, i.e., at large Da, the rate of heat production is larger than the rate of heat removal, resulting in accumulation of heat. This will lead to a so-called runaway. Decreasing Da gives rise to one, two, or three steady states. In the latter case only two situations are stable. Upon cycling the temperature one can observe ignition and extinction phenomena resulting in a so-called temperature hysteresis.
3. Nonisothermal external practical example

Another example of the existence of multiple steady states is in the oxidation of ammonia over a Pt/Rh wire gauze: 4 NH3 + 5 O2 4 NO + 6 H2O - Hr = 226 kJ / mol NH3

Start-up of the reaction in an industrial reactor takes place by directing a hydrogen torch to a spot at the top of the reactor and slowly introducing the ammonia-air mixture. The catalytic reaction starts and propagates through the reactor. However, if the torch is prematurely removed, the rate of reaction is not sufficient to sustain the required reaction temperature and the reaction is extinguished. Farrauto and Lee [14] simulated the start-up situation by preheating the air to different temperatures. When the reaction is ignited, the surface temperature of the gauze will rapidly increase, as shown in 4-8

Catalyst effectiveness

Extraparticle gradients

Figure 4.7 for an uncoated and a coated gauze. The coated gauze was obtained by covering the gauze with a layer of Pt, increasing the dispersion, resulting in higher activity.

1000 800 Ts (K) 600 400 300

extinction

ignition 400 500 Tb (K) 600

Figure 4.7 Comparison of ignition and extinction temperatures of an uncoated (- - -) and a coated () gauze in ammonia oxidation; p = 8 bar; 10% NH3 in air. Adapted from [14]

Figure 4.7 shows that the ignition temperature of the coated gauze is much lower than that of the uncoated one (520 K and 620 K, respectively). Once ignition has taken place, lowering the preheat temperature to below 480 K results in extinction of the reaction on the uncoated gauze, while the coated gauze does not extinguish until about 300 K.

4.3

Intraparticle gradients

Since the reactants have to diffuse to the active sites in the interior of catalyst particles, while at the same time they can react, a concentration profile over the particle diameter can develop. Mathematically this profile is described by a differential equation derived from a mass balance over a small shell of the catalyst particle. For a slab geometry and general kinetics:

Deff

d 2c rv = 0 dx 2

(16)

in which Deff is the effective diffusion coefficient and rv the rate based on unit volume of catalyst.

4-9

Catalyst Effectiveness

Intraparticle gradients

4.3.1 Diffusivity in Porous Media


The effective diffusion coefficient depends on the morphology of the porous material. Various transport mechanisms can be responsible for the movement of molecules in pores. These are summarized in Figure 4.8.

Viscous flow

Capillary condensation

Surface diffusion

Molecular diffusion

Knudsen diffusion

Activated diffusion (zeolites)

Figure 4.8 Transport mechanisms in pores.

In large pores (> ~10 nm) the diffusivity is determined by the molecular diffusivity (DAB ~10-4 - 10-5 m2/s for gases), in small pores by the Knudsen diffusivity (DK,i ~10-6 - 10-7 m2/s). The molecular diffusivity can be calculated using published correlations, see Tables 4.2 and 4.3 [7, 15], or by extrapolation from a known value at other conditions of temperature and pressure, using the following simple correlation for molecular diffusion of gases:
1 1.75 D AB p tot T

(17)

QUESTION:

Calculate the liquid diffusivity of thiophene using the Tables in Appendix 4.I.

The Knudsen diffusivity of component i in a cylindrical pore of radius ro equals:

D K ,i = 97 ro

T Mi

(18)

In a transition region the Bosanquet formula can be used that accounts for both contributions: 1 1 1 = + D D D AB K ,i m2s-1 (r in m) (19)

4 - 10

Catalyst effectiveness

Intraparticle gradients

Table 4.2 Correlations for the determination of binary diffusivity values of gases [7, 15].

D AB = 18.583 10 8

1 1 T 3 M + M B A 2 ptot AB D , AB

m 2 s 1

with:

T ptot MA MB

AB D,AB

= = = = = =

temperature (K) total pressure (atm) molecular weight of A (g/mol) molecular weight of B (g/mol) collision diameter for pair AB () collision integral for pair AB (-)

AB = 0.5(A + B) A and B can be obtained from tables (see Appendix 4.I).


The collision integral D,AB is determined using the following procedure. First the Lennard-Jones potential is calculated starting from:

AB

= A B k k k

A / k and B / k are so-called Lennard-Jones force constants and can be obtained from tables (see
Appendix 4.I). From

AB
k

one can determine the Lennard-Jones potential

kT

AB

which can be found in tables together

with the accompanying value for D,AB.

In zeolitic materials, where the molecules are of pore size dimensions, activated configurational diffusion takes place for which even lower diffusivity values, below 10-8 m2/s hold [16]. In this case diffusion resembles that of surface diffusion, migration over the surface in the adsorbed phase, which explains the sometimes unexpected high diffusional flux [17]. A more rigorous approach for the combined effect of the various diffusion modes is based on the Maxwell-Stefan approach [17], for porous materials often referred to as the dusty gas model.

4 - 11

Catalyst Effectiveness

Intraparticle gradients

Table 4.3 Correlations for the determination of binary diffusivity values of liquids [7, 15].

D AB = 7.4 10 15

T BM B ~ V A 0.6

m 2 s 1

with:

= B =
T

MB = = ~ VA =

temperature (K) association factor of solvent B: 2.6 for water 1.9 for methanol 1.5 for ethanol 1.0 for benzene, heptane, ether, etc. molecular weight of solvent B (g/mol) viscosity of solution (Pa s) molar volume of solute A (cm3/mol)

The molar volume of various elements and substances are tabulated (see Appendix 4.I). The molar volume of a molecule is obtained by adding the contributions of the various fragments.

Additionally, due to the presence of solid material the volume in which diffusion can take place is reduced by a factor p, the particle porosity. The tortuous path increases the diffusion length for a molecule relative to the spatial coordinate. Moreover, the tortuosity reduces the concentration gradient in this direction. Both effects are included in the tortuosity factor p, and are illustrated in Figure 4.9 below.
only fraction open for diffusion combined to tortuosity tortuous path longer

Flux direction
component gradient in flux direction gradient dc/dx direction

Figure 4.9 Effect of porosity and tortuous path on diffusion.

The effective diffusivity can hence be expressed as:

4 - 12

Catalyst effectiveness

Intraparticle gradients

Deff =

p D 0.05D - 0.1D p

(20)

The ususual treatment of diffusion-reaction problems leads to a balance equation such as equation (16). The effective diffusivity Deff is defined by the following relationship between flux and concentration gradient:

J i = Deff ,i

dci dx

(21)

in which Ji is the flux of species i based on the catalyst surface area (i.e., the number of moles i crossing the surface by diffusion) and Deff given by equation (20). QUESTION:
The molecular and Knudsen diffusion coefficients differ orders of magnitudes. Does this mean that the diffusion fluxes in, e.g., zeolitic materials (Knudsen or configurational diffusion) are extremely small?

4.3.2 Isothermal internal mass transport


The form of the steady-state mass balance will depend on the particle shape (sphere, cylinder, trilobe, slab, etc.). For a slab geometry and first order irreversible reaction equation (22) is valid:

Deff

d 2c kvc = 0 dx 2

(22)

With the boundary conditions:


x=L x=0 c = cs dc/dx = 0

(at surface of slab) (at center of slab)

the solution of the concentration profile is (see also box below):

x cosh c L = cs cosh( )

(23)

in which is known as the Thiele modulus which is the square root of the ratio of the reaction rate and the diffusional rate in the particle:

4 - 13

Catalyst Effectiveness
kv Deff Derive equations 23 and 24.

Intraparticle gradients

=L

(24)

QUESTION:

Figure 4.10 shows the dimensionless concentration c* (= c / cs) as a function of the dimensionless distance in the slab x* (= x / L).
1.0

1.0

0.1

0.8

c*

0.6

0.4

L x+dx x

2.0
0.2

10.0
0.0 1.0 0.8 0.6 0.4 0.2 0.0

x*

Figure 4.10 Dimensionless concentration as a function of dimensionless distance for various values of the Thiele modulus. Hyperbolic functions:
4.0

sinh( x) =

e e 2 x e + ex cosh( x) = 2 sinh( x) 1 = tanh( x) = cosh( x) coth( x)


x

3.0

2.0

cosh
1.0

sinh

Limits: x>3: x < 0.3 :

tanh
tanh (x) 1 tanh (x) 0
0.0 0.0 0.5 1.0 1.5 2.0

When the Thiele modulus is small no internal concentration profile exists. In case of large values, due to the existence of a concentration profile, the catalyst is not effectively used, and an effectiveness factor is defined as:

4 - 14

Catalyst effectiveness
dc Ap dx x = L observed reaction rate 0 i = = = rate without internal gradients rv ,chem ( c s ,Ts ) V p rv ,chem ( c s ,Ts ) V p

Intraparticle gradients

Vp

rv ( c ,T ) dV

Deff

(25)

Only for simple reaction kinetics an analytical expression for the effectiveness factor can be given, e.g., for a first order reaction in a slab of thickness 2L, Eq. 26. In comparing catalyst activities and in kinetic studies one needs data that are not disguised by concentration gradients. For a 5% tolerance level the criterion for the absence of internal diffusion limitations is:

i =

tanh( )

= 1 0.05

(26)

QUESTION:

Derive equation 26.

The result for the slab geometry can be generalized to other catalyst geometries. For any geometry the characteristic length should then be defined as: L= Vp Ap =

1 a'

(27)

to obtain essentially the same result as for the slab geometry given in Eq. 26. Hence, for a slab L = L, for a cylinder L = R / 2, and for a sphere L = R / 3. Figure 4.11 is a plot of i as a function of for all three geometries.

sphere
slab: cylinder:

cylinder

slab

i =
i =

tanh( )

I 1 (2 ) 2I 0 (2 )
i

sphere:

i =

1 1 1 tanh(3 ) 3
0.1 0.1

10

Figure 4.11 Internal effectiveness factor as a function of the Thiele modulus for a 1st order irreversible reaction.

Figure 4.11 clearly shows that in the limit of going to zero or infinity, all three geometries yield the same result for the internal effectiveness factor i:
4 - 15

Catalyst Effectiveness

Intraparticle gradients

0 i 1 i
1

A generalization can also be made to account for different reaction orders (see box below).

Derivation of the generalized effectiveness factor and Thiele modulus

The mass balance of Eq. 16 can be rewritten as follows:

d dc Deff = rv (c) dx dx
Multiplying the left and right side of the equation with Deff and integrating over c once with dc dc dc = dx and using = 0 yields: dx x =0 dx

Cs dc dc d D D dx = Deff rv (c) dc eff eff Cc dx dx dx

The left part of this equation is of the form

y' y dx for which holds:


y = 2 y ' y dy

y' y dx =

1 2

( y 2 )' dy = 1 y2 2

Hence, for the flux (= y) we obtain:

Deff

cs dc = 2 Deff rv (c) dc cc dx

and for the internal effectiveness factor:

dc dx x = L i = = rv (c s ) L Deff

2 Deff rv (c) dc
cc

cs

rv (c s ) L

So now the Thiele modulus has been defined as = 1/i for (Eq. (28)). In this region of strong diffusion effects, the pellet center concentration approaches zero, or, for reversible reactions, ceq.

4 - 16

Catalyst effectiveness

Intraparticle gradients

The following result for the generalized Thiele modulus, which is independent on the reaction kinetics and defined such that in the limit of i 1 [6, 18], is obtained:

= L

rv (c s )
2 Deff rv (c) dc
cc cs

(28)

in which cc is the concentration in the center of the catalyst particle, which becomes zero (irreversible reaction) or ceq (reversible reaction) for large Thiele moduli. Equation (28) for an irreversible nth order reaction (n > -1) becomes:

=L

k v ( n +1 ) n 1 2 cs Deff

(29)

Figure 4.12 shows the internal effectiveness factor as a function of the generalized Thiele modulus for a number of irreversible reactions.
3rd order 2nd order 1st order 0th order
1

0.1 0.1

10

Figure 4.12 Internal effectiveness factor versus generalized Thiele modulus at various reaction orders ( r = k cn) .

QUESTIONs: Derive equation (29). Explain the behaviour of the zero order reaction.

4 - 17

Catalyst Effectiveness

Intraparticle gradients

4.3.2.1 Reversible reactions


For the reversible first order reaction

A B
k

k+

K eq =

k+ k

the rate can be expressed as (no B fed):

r = k+

1 + K eq c As c A K eq 1 + K eq

(30)

leading to the following expression for the Thiele modulus:

=L

k+ Deff

1 + K eq K eq

(31)

Figure 4.13 shows the internal effectiveness factor as a function of the Thiele modulus at various values of Keq.
1

Keq

10 1 0.1

0.1 0.1

10

Thiele 1st order = L

k+ Deff

Figure 4.13 Internal effectiveness factor versus Thiele modulus for reversible reaction.

Clearly, equilibrium constants smaller than 1 markedly decrease the effectiveness factor. QUESTION:
Derive equation (31).

4.3.2.2 Langmuir type rate expressions


The generalized effectiveness factor and Thiele modulus are equally well applicable to Langmuir type rate expressions, as shown in Figure 4.14 for the reversible reaction A B for which rate expression (32) holds (surface reaction rate determining). 4 - 18

Catalyst effectiveness
rv = kK A p A 1 + K A pA

Intraparticle gradients

(32)

(zero order)
100
1

10

0.1

0 (1st order)

0.1 0.1

10

0 = L

kK A Deff (1 + K A pAs )

Figure 4.14 Internal effectiveness factor versus a Thiele modulus 0 for a reaction with a Langmuir type rate expression; effect of KA.

For reaction systems in which both the reactant(s) and the products adsorb on the active sites, i.e., reactions for which the following general rate expression holds:
rv = kK A p A i = A, B , etc.

1+

pi

(33)

the generalized Thiele modulus can be expressed as [18]:

= 0 f (B ) = L

B kK A Deff (1 + K i pis ) 2 (1 + B ) B ln (1 + B )

(34)

with

1 = B

1 + K i p is 1 Deff , A p As K i Deff ,i

B > 1

i = A, B, etc.

(35)

Figure 4.15 shows the internal effectiveness factor as a function of the generalized Thiele modulus at different values for the adsorption factor B.

4 - 19

Catalyst Effectiveness

Intraparticle gradients

B=
1

(zero order)
100

0 - 0.5 - 0.95

0.1 0.1 1 10

0
Figure 4.15 Effectiveness factor versus Thiele modulus for general Langmuir type rate expression. Adapted from [18]

If the catalyst predominantly adsorbs the reactant A and the adsorption is very strong the reaction becomes a zero order reaction, which is reflected by the parameter B approaching infinity. On the other hand, in the case of weak adsorption, the reaction follows first order kinetics, and B approaches zero. QUESTION:
What is the physical meaning of negative values for B?

Figure 4.16 shows the effect of the adsorption factor on the generalized Thiele modulus.
strong product adsorption
B 2(1+ B ) B ln(1+ B )
6 5 4 3 2 1 0 -1 0 1 2 3 4 5 6 7 8 9 10

strong reactant adsorption

Figure 4.16 Effect of B on Thiele modulus.

It can be seen that strong reactant adsorption, i.e., large B, decreases the Thiele modulus, while strong product adsorption, i.e., negative B, has the opposite effect. The Thiele modulus is not affected when both reactants and products adsorb weakly (B = 0). Similar expressions as Eqs. (34) and (35) can be derived for reversible reactions and dual site models (see Chapter 3). Then only B changes its meaning. 4 - 20

Catalyst effectiveness

Intraparticle gradients

4.3.2.3 Expansion
When volume change accompanies a reaction, e.g., in the case of the first-order gas phase reaction A nB this volume change must be accounted for. This is done through the volume change modulus:

A = (n 1) y A0
with yA0 the initial mol fraction of A.

(36)

The effect of volume changes on the effectiveness factor is shown in Figure 4.17 for first-order kinetics at different values of the Thiele modulus.
1.6 1.4

0 0.3 0.5 1 2


i i

1.2 1.0 0.8 0.6 0.4 -1.0

0.0 1.0 2.0 3.0

A
Figure 4.17 Ratio of effectiveness factors with and without expansion; first-order kinetics. Adapted from [19].

Figure 4.17 was obtained by applying the following approximation for the Thiele modulus in the presence of volume changes [19]:

' = (1 + A )m
with:
m 0.4 for expansion m 0.35 for contraction

(37)

4 - 21

Catalyst Effectiveness

Intraparticle gradients

This expression can be substituted in the equations for the effectiveness factor (e.g., Eq. 26). For large Thiele modulus ( > 5) the ratio i / i (with i the effectiveness factor in case of volume change) becomes independent on and for first-order reactions takes the form:

1 2 i ' 1 = 2 2 ln(1 + A ) i A A

(38)

4.3.3 Diffusion control?


As during a kinetic study one measures the observed reaction rate and not the intrinsic rate, one cannot determine whether the criterion of Eq. 26 is satisfied. Therefore, the following combination is introduced that yields a procurable quantity, and is referred to as the Wheeler-Weisz modulus [1, 6]:

= i 2 =

observed rate ' diffusion rate'

(39)

From series expansion the following criterion for negligible internal diffusion limitation, called the Weisz-Prater criterion, results for an nth order reaction [2]:

= i 2 =

rv ,obs L2 n + 1 < 0.15 Deff c s 2

(40)

Note that cs is not an observable. However, its value may be calculated from Eq. 4. Alternatively, using cb instead is often justified. Figure 4.18 shows the effectiveness factor as a function of the Wheeler-Weisz modulus for different catalyst geometries, indicating that criterion (40) holds for the generalized Thiele modulus. Due to the definition of L, the internal effectivenes factor as a function of the Wheeler-Weisz modulus is fairly independent of the catalyst geometry.

4 - 22

Catalyst effectiveness
slab cylinder
1

Intraparticle gradients

sphere

0.1 0.1

i2

10

Figure 4.18 Isothermal internal effectiveness factor as a function of the Wheeler-Weisz modulus for different catalyst geometries.

4.3.4 Nonisothermal internal transport


The derivation of the internal temperature gradient can only be performed numerically, even for simple kinetics. Again the 5% criterion is used for the internal effectiveness factor. Assuming that the rate can be simplified into a temperature and concentration dependent part this yields:

i =

E T rv (c, T ) k (T ) f (c) k (T ) = v v = exp a s 1 = 1 0.05 rv (c s , Ts ) k v (Ts ) f (c s ) k v (Ts ) RTs T

(41)

where the concentration effects are assumed to be negligible in order to derive a criterion for the temperature effects only. In the nonisothermal case the reaction temperature in the center of the particle will be higher than the temperature at the external surface for exothermic reactions and lower for endothermic reactions. Since we focus on small deviations from the isothermal behaviour the rates in the particle center will be higher or lower, respectively, than at the surface conditions. An expression for the temperature rise in the particle is obtained from the mass (Eq. 16) and heat balance for a particle:
p ,eff

d 2T = rv ( H r ) dx 2

(42)

Effective thermal conductivity values of porous materials, p ,eff , range from 0.1 to 0.5 W K-1 m-1 in gaseous atmospheres [15] and are only slightly larger than those of the gas phase. Straightforward combination of Eqs. 16 and 42 (first order reaction) and integration with boundary conditions: x=L c = cs T = Ts 4 - 23

Catalyst Effectiveness
dc dT = =0 dx dx

Intraparticle gradients

x=0

leads to a simple general result that relates the temperature and concentration profile over a particle: T Ts c c = i s Ts cs (43)

with:

i =

( H r ) Deff c s

p ,eff Ts

Ti ,max
Ts

(44)

the internal Prater number, which represents the maximum relative temperature difference across the particle or the ratio of the heat production rate and the heat conduction rate in the particle (typical values for exothermal reactions are from 0 to 0.3). Series development and combination of Eqs. 41 and 43 yields the criterion for absence of temperature effects [2, 20]:

s i i

Ea = R T s

( H r ) Deff c s p ,eff Ts

rv ,obs L2 D c eff s

< 0.1

(45)

This criterion contains observables, so an estimation can be made about the presence of temperature gradients. The result is not unexpected and is quite similar to that for external transport. The WheelerWeisz parameter represents the concentration profile, the Prater number the maximum heat production relative to the heat removal by conduction, and the dimensionless activation energy the sensitivity of the rate towards a temperature change. It can be shown that i i 2 represents the relative temperature gradient over the particle, like e Ca represents the relative temperature gradient over the film. In the application of the criterion cs and Ts are not always known, but they can be calculated from the external transport balances, Eqs. 3 and 8. Alternatively, using the values for the bulk phase generally yields a good estimate. It will be clear that in case of exothermic reactions the effectiveness factor can become larger than one. This is indicated by the numerical solution for the effectiveness factor in Fig. 4.19 for various values of i and s . Situations may arise in which concentration effects counterbalance the temperature effect, resulting in an effectiveness factor of around one. On the other hand, the temperature effect in the case of endothermic reactions only adds to the lowering of the effectiveness factor by a concentration profile.

4 - 24

Catalyst effectiveness

Intraparticle gradients

10

i
0.4 1

0.6

0.2 -0.2 0.1 0.1

10

Figure 4.19 Internal effectiveness factor as a function of the Thiele modulus for nonisothermal reactions at different values of the Prater number i and s = 10 (numerical solutions for a first order reaction).

4.4 Comparison of Criteria


It is interesting to determine which criterion to avoid mass- and heat transport disguises is most severe, in order to have a quick impression whether all criteria are satisfied. Table 4.4 summarizes the various criteria.

Table 4.4 Summary of criteria for negligible mass and heat transport.
Isothermal, external mass transport

Ca =

rv ,obs a ' k f cb

<

0.05 n

(6)

Nonisothermal, external mass and heat transport

b e Ca =

E a ( Hr ) k f cb rv ,obs < 0.05 h Tb R Tb k f a 'cb

(13)

Isothermal, internal mass transport

= i 2 =

rv ,obs L2 n + 1 < 0.15 Deff c s 2

(40)

Nonisothermal, internal mass and heat transport

s i i

Ea = R T s

( H r ) Deff c s p ,eff Ts

rv ,obs L2 D c eff s

< 0.1

(45)

4 - 25

Catalyst Effectiveness

Comparison of criteria

4.4.1 Combined In ternal and External Mass Transport - isothermal


If both external and internal mass transport limitations are expected to play a role the mass balance can be written as: a' k f (cb c s ) = rv ,obs ( c ,T ) = i rv ( c s ,Ts ) which for a first-order isothermal reaction leads to: cs = cb 1 k 1 + i v a' k f = 1 1 + i Da (47) (46)

For a first-order reaction the ratio cs / cb is by definition the external effectiveness factor e. The overall effectiveness factor is expressed as:

= ie =

i 1 + i Da

(48)

Introduction of the Biot number for mass transport Bim:

Bim =

kf L Deff

(49)

makes it possible to express the Dahmkhler number Da as a function of the Thiele modulus (Da from Eq. 7) : Da =

2
Bi m

(50)

The overall effectiveness factor then becomes (with Eq. 26):

tanh ( ) 1 + Bi m
Express c / cb as a function of , x / L, and Bim.

tanh ( )

(51)

QUESTION:

Figure 4.20 shows the dimensionless concentration profiles over the particle and the film layer for two values of and Bim. The controlling transport mechanism depends on the value of Bim. A large Bim indicates internal control, while a small Bim leads to external control. Generally, Bim is greater than 10 (20 to 100) for conditions in laboratory reactors, resulting in internal diffusion control being more important. 4 - 26

Catalyst effectiveness

Comparison of criteria

1.0 0.8

Bim
100

c/cb
0.6 0.4 0.2

1 10

film
5
0.2 0.4 0.6 0.8 1.0

x/L

Figure 4.20 c / cb versus x / L; slab geometry, first-order isothermal reaction.

Physical meaning of Bim

The mass transferred across the interface between the bulk fluid and the particle surface must be equal to the mass flux at the particle surface. Thus, for a first-order isothermal reaction occurring in a slablike catalyst particle: k f (cb c s ) = Deff dc dx (52)
x=L

So, in dimensionless form:


dc * dx * =
x* =1

kf L Deff

(1 c * ) = Bim ( 1 c*)

Bim =

rate of mass transfer rate of diffusion

(53)

or alternatively:

dc * dx * x* =1 Bi m = 1 c*
Limiting cases

Bi m =

concentration gradient at surface concentration gradient in film

(54)

Bim large : Bim small :

tanh( )

Bim

= i

(internal control) (external control)

1 Da

4 - 27

Catalyst Effectiveness

Comparison of criteria

For general reaction kinetics and catalyst geometry, the ratio of the criteria for internal and external mass transfer can be rearranged as follows (neglecting the reaction order in the Wheeler-Weisz modulus):

i 2
Ca

rv ,obs L2

Deff c s rv ,obs L

k f cb

kf L Deff

= Bim

(55)

Since the Biot mass number generally ranges between 20 and 100 for conditions in laboratory reactors, the Weisz-Prater criterion is more severe than the Carberry criterion, i.e., internal mass transport limitations are expected to occur earlier than external limitations.

4.4.2 Heat Transport External versus Internal


Comparing the criteria for intra- and extraparticle heat transport the ratio of these criteria yields globally:

Deff h s i i 2 hL Bi m = = Bi h b e Ca k f p ,eff p ,eff

(56)

The limits of the Biot heat number, Bih, are 0.01 and 50 [6], so the particular conditions will determine which criterion is the most severe. In laboratory reactors Bih is often < 1. It is obvious that decreasing the particle size will shift the largest gradient to the film layer around the particle. Under the assumption that inside a particle the reactant concentration becomes zero it can easily be shown that the ratio of the external and internal temperature gradients equals: Ts Tb (T )e Bi m = = T Ts (T )i Bi h or: Ca 1 Ca (57)

Te Ti ,max + Te

Bi m Ca Bi h Bi m 1+ Bi 1 Ca h 10 - 104 10-4 - 10-1 Gas-Solid Liquid-Solid

(58)

where [1]:

Bi m e = = Bi h i

QUESTION:

Derive equations (57) and (58).

4 - 28

Catalyst effectiveness

Comparison of criteria

This indicates that in liquid-solid reactions the external gradient is negligible, while in gas-solid reactions it will depend on the specific parameter values. Unlike in industrial operations (high flow rates, large particles) where as a rule of thumb the internal temperature gradient is assumed to be the largest [21], in laboratory operation the external temperature gradient is generally larger. Figure 4.21 shows the ratio of the external temperature gradient and the total gradient as a function of the Carberry number Ca at several ratios of Bim Bih .

(T )e (T )i +e

1.0 0.8 0.6

500 50 20 10 5

0.4 0.2 0.0 0.0

Bim/Bih

0.2

0.4

0.6

0.8

1.0

Ca
Figure 4.21 Temperature gradient ratio versus Ca.

The overall effectiveness factor in the case of in- and external mass transport limitations and an external temperature gradient (particle isothermal at Ts) for a first order reaction is plotted in Figure 4.22 as a function of the Carberry number.
10

e
1

(Ts)

0.5
0.1

0.2 0

0.01 0.001

0.01

0.1

Ca

Figure 4.22 Overall effectiveness factor versus Ca; Bim = 100; s = 20.

Whether the criteria for intra- or extraparticle heat transport limitations are more severe than the corresponding ones for mass transport depends on the absolute value of the products s i and

b e , respectively. In some cases the former product exceeds unity, but more often not [6], so
internal temperature gradients do not occur frequently. The latter product is a factor Bim Bih larger, generally >10 (gases), which means that in this case the extraparticle temperature gradient occurs 4 - 29

Catalyst Effectiveness

Comparison of criteria

earlier than an extraparticle concentration gradient. For labscale operation the following order in relative importance of the various gradients can be indicated:

(T grad )bed > (T grad )ext > (c grad )int , (T grad )int > (c grad )ext
The gradient over the catalyst bed is discussed in reference [22].

(59)

4.5 Effect of Particle Transport Limitations on Behaviour


The observed reaction rate can be expressed as follows:
rv ,obs = rv ,chem (cb , Tb ) = e i rv ,chem (cb , Tb ) (60)

If external mass transport limitations strongly dominate the rate becomes equal to the mass transfer rate: rv ,obs = k f a' cb um cb Lm 2

(61)

in which m represents the power of the Reynolds number in the Sherwood correlation in Table 4.1. Hence, a first order dependence is observed. Furthermore, the apparent activation energy is negligible, because the mass transfer coefficient (which is the observed rate constant) is fairly independent on the temperature. On the other hand, the observed rate constant does depend on the flow rate and particle size (through correlations in Table 4.1). If internal diffusion limitations dominate i 1 so for an nth order reaction:

rv ,obs L

k v Deff c

n +1 s

Ea L c s exp 2 RT s
1
n +1 2


( n +1) 2

(62)

So the rate of reaction is dependent on the particle size, has an apparent order of activation energy that is half the true activation energy. QUESTION: Derive equations (61) and (62).

and apparent

As an example of these so-called falsified kinetics Figures 4.23 and 4.24 show results obtained by Post et al. [23] for the Fischer-Tropsch synthesis of middle distillates: n CO + m H2 CnH2(m-n) + n H2O which has the following intrinsic reaction rate:

4 - 30

Catalyst effectiveness

Effect of particle transport limitations

rv = k v p H 2

(zero order CO)

(63)

Figure 4.23 is a plot of the internal effectiveness factor as a function of the Thiele modulus.
2

0.1 0.1

10

3
Figure 4.23 Fischer-Tropsch reaction; Internal effectiveness factor as a function of the Thiele modulus; Catalyst: Co-Zr/SiO2; H2/CO = 2; p = 21 bar; T = 473 513 K; dp = 0.38 2.6 mm [23].

Figure 4.23 shows that a large part of the experiments was performed in the internal diffusion limited region. Figure 4.24 gives additional prove that this is the case.
0.1

dp/mm

kv
0.01

0.38

2.4

1.4

0.001 1.90

1.95

2.00

2.05

2.10

1000/T

Figure 4.24 Fischer-Tropsch reaction; Arrhenius plots at three catalyst diameters; same conditions as Figure 4.23 [23].

The rate constant clearly depends on the temperature, so external transport limitations do not play a role. The fact that the rate constant increases with decreasing particle size shows that internal limitations are important. Furthermore, the activation energy obtained from the Arrhenius plots is not the true activation energy but Ea,true / 2. 4 - 31

Catalyst Effectiveness

Effect of particle transport limitations

At low temperatures reactions are generally kinetically controlled. With increasing temperatures one first enters the diffusion-limited region and at even higher temperatures the film (extraparticle) diffusion limited region. In Arrhenius plots this is nicely seen as a changing slope of the rate versus 1/T dependence, which can be used as an indication for the presence of limitations. Interesting examples of this behaviour can be found in the literature, e.g, for catalyzed gasification of carbon [24]. Bernardo and Trimm [24] studied the catalyzed steam gasification of carbon: C + H2O
Ni CO + H2

Figure 4.25, which is a plot of the observed reaction rate as a function of reciprocal temperature, clearly shows the three regimes.
5

Ea(kJ/mol) 0
1

61 1 0.75

robs

164
0.1

0.6 order n
0.01 0.9 1.0 1.1 1.2 1.3 1.4

1000/T

Figure 4.25 Carbon gasification; observed reaction rate versus reciprocal temperature; psteam = 26 kPa; catalyst: Ni/Al2O3 [24].

Changing activation energies, however, are not always indicative of the presence of limitations. The approach of thermodynamic equilibrium in the case of exothermic reactions can also cause this phenomenon, like in hydrogenation reactions [24]. Figure 4.26 illustrates this for the catalytic hydrogenation of carbon:
Ni

C + 2H2

CH4

4 - 32

Catalyst effectiveness
30 10

Effect of particle transport limitations

. eq

180 1.4

Ea(kJ/mol) order n

robs

-82 1.8 265 <1


0.9 1.0 1.1 1.2 1.3 1.4

0.1 0.03 0.8

1000/T

Figure 4.26 Carbon hydrogenation; observed reaction rate versus reciprocal temperature; phydrogen = 100 kPa; catalyst: Ni/Al2O3 [24].

QUESTION:

Explain the observed behaviour in Figure 4.26.

Also changes in rate-determining steps and catalyst deactivation might be causes of changing activation energies and reaction orders. Table 4.5 gives a survey of the various observations that can be made when mass transfer affects the isothermal kinetic behaviour of catalyst particles.
Table 4.5 Apparent rate behaviour depending on rate controlling regime (isothermal case). Controlling Process apparent order apparent activation energy dependence L Dependence U

Kinetics Internal diffusion External mass transfer


*)

n
n +1 2

Ea (true) Ea (true) 0

1/L Lm-2
*)

um
*)

m represents the power of the Reynolds number in the Sherwood correlation in Table 4.1.

Temperature gradients in endothermic reactions increase the effects of concentration gradients. In exothermic reactions the thermal effects can compensate the concentration effects or dominate completely. In the latter cases the apparent activation energy can become higher than the true one due to a kind of light-off [7]. Also in this case at the highest temperatures the apparent Ea approaches zero.

4 - 33

Catalyst effectiveness

Notation

Notation
a Ap Ach B B1 c dH dp DK,,i Dif D Deff Ea f(c) h Ji kv kf L m Mi n Och p r0 rv rv, chem rv, obs rV, obs rW R T u Vp x y0 specific surface area of catalyst particle external surface area of particle cross sectional area of monolith channel defined in Eq. 35 constant in Table 4.1 concentration hydrodynamic channel diameter monolith particle diameter Knudsen diffusivity of component i molecular diffusivity of i in fluid phase f average diffusivity effective diffusivity in particle activation energy concentration function in rate expression heat transfer coefficient flux of component i rate constant per unit particle volume mass transfer coefficient characteristic catalyst dimension exponent in Eq. 37 molar mass of component i reaction order periphery of monoliths channel cross section pressure average pore radius reaction rate per unit particle volume intrinsic reaction rate per unit particle volume observed reaction rate per unit particle volume observed reaction rate per unit bed volume reaction rate per unit catalyst mass universal gas constant temperature superficial velocity particle volume coordinate molar fraction of reactant in reactor feed gas m-1 m2 m2 mol m-3 m m m2 s-1 m2 s-1 m2 s-1 m2 s-1 J mol-1 J m-2 s-1 K-1 mol m-2 s-1 (mp3 mol-1)n-1 s-1 m s-1 m kg kmol-1 m Pa m mol s-1 mp-3 mol s-1 mp-3 mol s-1 mp-3 mol s-1 mb-3 mol s-1 kg-1 J mol-1 K-1 K m s-1 m3 m -

Dimensionless numbers

Bim Biot number of mass Bih Biot number of heat 4 - 34

Catalyst effectiveness
Ca Da Nu Pe Pep Pr Re Rep Sc Sh
Greek

Notation

Carberry number Dahmkhler number Nusselt number Pclet particle number Pclet particle number Prandtl number Reynolds number Reynolds particle number Schmidt number Sherwood number

Prater number dimensionless activation energy wall thickness in monolith enthalpy of reaction porosity of bed volume change modulus porosity of monolith porosity of monolith porosity of particle overall effectiveness factor internal effectiveness factor internal effectiveness factor corrected for expansion external effectiveness factor thermal conductivity of fluid phase effective thermal conductivity dynamic viscosity kinematic viscosity of liquid density collision diameter for pair AB average residence time CSTR particle tortuosity factor Thiele modulus part of Thiele modulus in Eq. 34 Thiele modulus corrected for expansion Wheeler-Weisz modulus

m J mol-1 J m-1 s-1 K-1 J m-1 s-1 K-1 kg m s-1 m2 s-1 kg m-3 s -

w
A b M
Hr

p
i

i '
e f eff l AB p

4 - 35

Catalyst effectiveness
Subscripts

Notation

app b c chem e eq eff f i M obs p s t tot w

apparent in bulk phase core, center chemically controlled external equilibrium effective fluid phase intraparticle monolith observed particle at external particle surface tube total at the reactor wall

4 - 36

Catalyst effectiveness

Literature

Literature
1. J.J. Carberry, in Catalysis: Science and Technology, Vol. 8, J.R. Anderson and M. Boudart (Eds.), Springer-Verlag, Berlin, 131 (1987). 2. J.A. Moulijn, A. Tarfaoui and F. Kapteijn, Catal. Today 11, 1 (1991). 3. D.E. Mears, J. Catal. 20, 127 (1971). 4. F. Kapteijn, G.B. Marin and J.A. Moulijn, 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.), Studies in Surface Science and Catalysis, Vol. 79, Elsevier, Amsterdam, 251 (1993). 5. A.A.C.M. Beenackers and W.P.M. van Swaaij, Chem. Eng. Sci. 48, 3109 (1993). 6. G.F. Froment and K.B. Bischoff, Chemical Reactor Analysis and Design, John Wiley, New York, 2nd ed., 1990. 7. L.K. Doraiswamy and M.M. Sharma, Heterogeneous Reactions: Analysis, Examples and Reactor Design, Vol. 1: Gas-Solid and Solid-Solid Reactions, John Wiley and Sons, New York, 1984. 8. P. Trambouze, H. van Landeghem and J.P. Wauquier, Chemical Reactors. Design/ Engineering/ Operation, Ch. 11., Editions Technip, Paris, 1986. 9. A. Cybulski, M.J. van Dalen, J.W. Verkerk and P.J. van den Berg, Chem. Eng. Sci. 30, 1015 (1975). 10. C.N. Satterfield, Heterogeneous Catalysis in Industrial Practice, Ch. 11, McGraw-Hill, New York, 2nd ed., 1992. 11. R.D. Hawthorn, AIChE Symp. Ser. Vol. 70(137), 428 (1974). 12. R.K. Shah and A.L. London, AIChE -J. 22, 344 (1976). 13. A. Cybulski and J.A. Moulijn, Catal. Rev.-Sci. Eng. 36, 179 (1994). 14. R.J. Farrauto and H.C. Lee, Ind. Eng. Chem. Res. 29, 1125 (1990). 15. C.N. Satterfield, Mass Transfer in Heterogeneous Catalysis, MIT Press, Cambridge, 1970. 16. J. Krger and D.M. Ruthven, Diffusion in Zeolites, John Wiley, New York, 1992. 17. R. Krishna, Gas Sep. Purif. 7, 91 (1993). 18. P. Schneider, Catal. Rev.-Sci. Eng. 12, 201 (1975). 19. V.W. Weekman and R.L. Gorring, J. Catal. 4, 260 (1965). 20. D.E. Mears, Chem. Eng. Sci. 26, 1361 (1971). 21. F.M. Dautzenberg, in Characterization in Catalyst Development, ACS Symp. Ser. Vol. 411, American Chemical Society, Washington DC, 99 (1989). 22. J.A. Moulijn, X. Xiaoding, F. Kapteijn and A.D. van Langeveld, Katalyse en katalysatoren (Eng.: Catalysis and catalysts), Lecture notes, TU Delft, 1997. 23. M.F.M. Post, A.C. van 't Hoog, J.K. Minderhout and S.T. Sie, AIChE-J. 35 1107 (1989). 24. C.A. Bernardo and D.L. Trimm, Carbon 17, 115 (1979).

F. Kapteijn and J.A. Moulijn, Laboratory reactors, in Handbook of heterogeneous catalysis Vol. 3, Ch. 9, G. Ertl, H. Knzinger and J. Weitkamp (Eds.), VCH, Weinheim, 1997.

4 - 37

Catalyst Performance Testing


Table 4.I.1 Lennard-Jones force constants calculated from viscosity data. COMPOUND Acetylene Air Argon Arsine Benzene Bromine i-Butane n-Butane Carbon dioxide Carbon disulphide Carbon monoxide Carbon tetrachloride Carbonyl sulphide Chlorine Chloroform Cyanogen Cyclohexane Ethane Ethanol Ethylene Fluorine Helium n-Heptane

Appendix 4.I

/k (K)
185. 97. 124. 281. 440. 520. 313. 410. 190. 488. 110. 327. 335. 357. 327. 339. 324. 230. 391. 205. 112. 10.22 282.

()
4.221 3.617 3.418 4.06 5.270 4.268 5.341 4.997 3.996 4.438 3.590 5.881 4.13 4.115 5.430 4.38 6.093 4.418 4.455 4.232 3.653 2.576 8.88

COMPOUND n-Hexane Hydrogen Hydrogen chloride Hydrogen iodide Iodine Krypton Methane Methanol Methylene chloride Methyl chloride Mercuric iodide Mercury Neon Nitric oxide Nitrogen Nitrous oxide n-Nonane n-Octane Oxygen n-Pentane Propane Sulphur dioxide Water Xenon

/k (K)
413. 33.3 360. 324. 550. 190. 136.5 507. 406. 855. 691. 851. 35.7 119. 91.5 220. 240. 320. 113. 345. 254. 252. 356. 229.

()
5.909 2.968 3.305 4.123 4.962 3.61 3.822 3.585 4.759 3.375 5.625 2.898 2.789 3.470 3.681 3.879 8.448 7.451 3.433 5.769 5.061 4.290 2.649 4.055

4 - 38

Catalyst effectiveness
Table 4.I.2 Values of the collision integral D based on the Lennard-Jones potential.

Appendix 4.I

kT

D
2.662 2.476 2.318 2.184 2.066 1.966 1.877 1.798 1.729 1.667 1.612 1.562 1.517 1.476 1.439 1.406 1.375 1.346 1.320 1.296 1.273 1.253 1.233 1.215 1.198 1.182 1.167

kT

D
1.153 1.140 1.128 1.116 1.105 1.094 1.084 1.075 1.057 1.041 1.026 1.012 0.9996 0.9878 0.9770 0.9672 0.9576 0.9490 0.9406 0.9328 0.9256 0.9186 0.9120 0.9058 0.8998 0.8942 0.8888

kT

D
0.8836 0.8788 0.8740 0.8694 0.8652 0.8610 0.8568 0.8530 0.8492 0.8456 0.8422 0.8124 0.7896 0.7712 0.7556 0.7424 0.6640 0.6232 0.5960 0.5756 0.5596 0.5464 0.5352 0.5256 0.5136 0.4644 0.4170

12
0.30 0.35 0.40 0.45 0.50 0.55 0.60 0.65 0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 1.25 1.30 1.35 1.40 1.45 1.50 1.55 1.60

12
1.65 1.70 1.75 1.80 1.85 1.90 1.95 2.00 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9

12
4.0 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 5.0 6. 7. 8. 9. 10. 20. 30. 40. 50. 60. 70. 80. 90. 100. 200. 400.

4 - 39

Catalyst Performance Testing

Appendix 4.I

Table 4.I.3 Additive volume increments for the estimation of the molar volume Vb at the normal boiling temperature. Substance Vb increment cm3/mol Air 29.9 Ammonia 25 Bromine 27 Carbon 14.8 Chlorine, terminal as R-Cl 21.6 medial as R-CHCl-R 24.6 Fluorine 8.7 Helium 1.0 Hydrogen (in compound) 3.7 Hydrogen molecule 14.3 Mercury 15.7 Nitrogen 31.2 in primary amines 10.5 in secondary amines 12.0 Oxygen, molecular 14.8 doubly bound 7.4 methyl esters and ethers 9.1 ethyl esters and ethers 9.9 higher esters and ethers 11.0 acids 12.0 joined to S, P or N 8.3 Phosphorus 27 Sulphur 25.6 Rings: 3-membered -6. 4-membered -8.5 5-membered -11.5 6-membered -15. naphtalene -30. anthracene -47.5

4 - 40

Catalyst deactivation

Introduction

CATALYST DEACTIVATION

5.1 Introduction
Obviously, a good catalyst has to show high activity. A high activity allows the use of relatively small reactors and operation at mild conditions. If thermodynamics favour low temperature (exothermal reactions) the activity of the catalyst usually is decisive for the success of a process. However, activity is not the only crucial property of a catalyst. A high selectivity for the desired products is often of even greater importance. Furthermore, it is important that a catalyst should retain its activity and selectivity for some time. Depending on the process used the catalyst useful life may vary from a few seconds, as in Fluid Catalytic Cracking (FCC) to several years, as in for instance ammonia synthesis. Causes for catalyst deactivation are various, and often more than one phenomenon occurs at the same time. The three main reasons for deactivation are poisoning, fouling, and thermal degradation (sintering). In the first two cases the catalyst might be regenerated in some way, while in the latter case the deactivation is generally irreversible and the catalyst has to be replaced. Catalyst attrition, which occurs in slurry- and fluidized-bed operation, is not deactivation in the true sense, but nevertheless this mechanical process also necessitates replacement of the catalyst.

5.2

Catalyst Deactivation Qualitative Description

According to the classical definition a catalyst does not change during reaction. In practice this is not true: during operation a catalyst looses activity and often also selectivity and mechanical strength. Catalyst deactivation is a common phenomenon rather than an exception, and is an important aspect for reactor design and operation.

5.2.1 Time Scale of Deactivation


The time scale of deactivation has profound consequences for process design. Figure 5.1 shows the deactivation time for several processes in the refining and petrochemical industry. This deactivation time refers to the time after which a catalyst has lost so much of its initial activity that it must be regenerated or replaced. As can be seen from Figure 5.1, the deactivation time varies greatly for the different processes. At the lower end we find the Fluid Catalytic Cracking (FCC) process. The FCC catalyst deactivates in a matter of seconds. Deactivation of a hydrodesulfurization (HDS) catalyst is much slower, depending on the feed, in the order of months or a year.

5-1

Catalyst deactivation
10 -1 10 0 10 1 10 2 10 3 10 4 10 5 10 6 10 7 10 8

Qualitative description

C3 dehydrogenation

HDS Hydrocracking Reforming FCC EO Fat hardening MA Aldehydes Hydrogenations Acetylene Oxychlorination SCR Formaldehyde NH3 oxidation TWC
101 102 103 104 105 106 107 108

Most bulk processes 0.1-10 year

Batch processes hrs-days

Time / seconds
10-1 100

1 hour 1 day

1 year

Figure 5.1 Time scale of deactivation of various processes. Figure 5.2 shows a generalized plot of activity (in terms of reaction rate constant or conversion) versus process time for various deactivation patterns.
conversion or kobs

process time

Figure 5.2 General deactivation behaviour High conversions are often achieved on the fresh catalyst, but these conversions drop rather rapidly as the catalyst adjusts to the operating conditions. Subsequent deactivation is usually much slower. In practice, the process conditions will be adapted so that a stationary level of activity is reached. Usually, this is done by increasing the operating temperature gradually to compensate for activity loss, often at the expense of selectivity. The higher temperature will also result in accelerated deactivation. At a certain time-on-stream (varying from seconds to years), deactivation can not be compensated for any more, resulting in a fast decline (not shown in Figure 5.2). Sometimes it is possible to regenerate the catalyst continuously (FCC, catalytic reforming), but often the process must be shut down for catalyst regeneration or for reloading the reactor with fresh catalyst. Whenever possible, regeneration and/or reloading are incorporated in the periodical (partial) revision of the plant. It will be clear that the time scale of deactivation influences the choice of reactor. Table 5.1 gives a global indication. Exceptions always exist, like the fixed-bed dehydrogenation of propene with a deactivation time in the order of 10 minutes.

5-2

Catalyst deactivation

Qualitative description

Table 5.1 Relationship between time scale of deactivation and reactor type. Time scale of deactivation Typical reactor type Years fixed-bed reactor Months moving-bed reactor, fixed-bed reactors in swing mode Days fluidized-bed reactor, slurry reactor; continuous regeneration Seconds entrained-flow reactor (riser) with continuous regeneration

5.2.2 Mechanisms of Deactivation


Catalyst deactivation can occur through: Fouling: secondary reactions of reactants or products, coke formation Poisoning: strong chemisorption of impurity in feed Aging or sintering: structural changes (loss of surface area) Redispersion and evaporation of catalyst components Mechanical deactivation: attrition and carbon filament growth

Figure 5.3 shows a schematic representation of the deactivation phenomena inside a catalyst particle.
Surface poisoning
S S

Selective poisoning Fouling


S S

Cl Cl

Cl Cl

= active site

Redispersion & evaporation

Cl Cl Cl Cl

Cl Cl Cl Cl

Sintering

Figure 5.3 Deactivation types. Fouling and poisoning are also referred to as chemical deactivation. Fouling is often caused by the deposition of carbonaceous material, coke, on the catalyst. Poisoning occurs by strong adsorption of feed impurities on the active catalyst sites. Poisoning as defined here refers to an irreversible process. A milder form of poisoning, also called inhibition, is the reversible adsorption of impurities or products. A better term is decreased activity due to competitive adsorption, and in principle this has been accounted for in the rate expression.

5-3

Catalyst deactivation

Qualitative description

Aging or sintering is also often referred to as thermal deactivation, because it usually occurs at high temperatures. However, in the sintering process the chemical environment also plays a role. For instance, a steam-containing atmosphere often accelerates sintering, especially of supports. Chlorine, on the other hand, may induce redispersion of the active pahse, like in Pt and Pd catalysts.

5.2.2.1 Fouling
Fouling covers all phenomena where a surface is covered with a deposit. Its origin may be dust, e.g., from combustion residues like ash or soot, or the undesired byproduct of a catalytic treatment. The injection of ammonia in selective catalytic reduction (SCR) of flue gases may lead to ammonium(bi)sulfate formation due to the presence of sulfur in the fuel. But more generally the deposit is a highly condensed carbonaceous material with a low H/C ratio, referred to as coke. This coke is produced by unwanted polymerization and dehydrogenation (condensation) of organic molecules present in the feed or formed as product. These reactions leave a layer of highly hydrogen deficient carbonaceous material on the catalyst surface, making the active sites inaccessible. This coke formation may even continue until complete pore blocking occurs. Another type of carbon formation is observed on metal catalysts (see Figure 5.4). Under some conditions hydrocarbons decompose at the metal surface of small crystallites and the carbon diffuses through the metal and forms a carbon nanotube at the other side of the crystallite. The crystallite is even lifted from the surface and becomes the head of a growing filament. Sometimes the crystallites split up and more filaments are formed. Under hydrogen rich conditions this process may be reversible.

C H2 CnHm H H C

Metal crystallite

Support

Growing fillament

Figure 5.4 Carbon formation on supported metal catalyst. CH4 decomposition at 873 K on Ni/CaO [1] Continued growth can lead to complete destruction of the catalyst particle. This phenomenon can also occur at metal surfaces and can lead to destruction of piping.

5-4

Catalyst deactivation

Qualitative description

5.2.2.2 Poisoning
When impurities in the feed interact more strongly with the active sites than reactant molecules, the activity of the catalyst decreases due to poisoning. Poisoning can be either temporary (reversible) or permanent (irreversible). In the former situation, the poison can be removed, whereas this is not the case in the latter situation. The distinction between temporary and permanent poisoning is not always this clear: poisons that are strongly bonded to the active sites at low temperature can sometimes be removed at a higher temperature and, hence, could be referred to as temporary poisons. Poisoning can be selective or nonselective (see also Section 5.3.2). The occurrence of selective poisoning can have several explanations: The impurity preferentially adsorbs on the catalytically active sites, which make up only a small fraction of the catalyst. Poisoning molecules interact with each other. More than one active site is poisoned by one poisoning molecule (at low poison coverage) Poisoning is diffusion-limited so that only the outer catalyst shell is poisoned: pore mouth poisoning.

On the other hand, nonselective poisoning results in coverage of the whole catalyst surface and the catalyst activity depends linearly on the number of unpoisoned active sites. If by adsorption of a molecule from the feed on an active site the activity increases, one speaks of a promoter. For instance, addition of a small amount of an alkali hydroxide to the reactant stream in the production of amines from nitriles by hydrogenation, improves the activity of the metal catalyst and the selectivity to the desired primary amine [2]. More commonly the promoter is incorporated into the catalyst as it is prepared (e.g., addition of a small amount of alumina to the conventional iron catalyst for ammonia synthesis [3a]. A molecule could also adsorb selectively on sites that catalyze undesired reactions, thereby increasing the selectivity for the desired reaction. In that case the term modifier is used. Catalyst poisons exist in various forms, as shown in Table 5.2. Metals or compounds reducible to metals under reaction conditions may alloy with the active catalyst metal and reduce its effectiveness [3b]. Examples are Cu in Ni and Ni or Fe in Pt. In the hydrogenation of the product from thermal cracking of ethane (see [4a]), in order to remove acetylene, adding Cu to the Ni catalyst results in selective hydrogenation of acetylene, while leaving ethene unaffected [3b]. Hence, in this case the poison can be termed a modifier: the rate of the undesired reaction (hydrogenation of ethene) is decreased. Similarly, on a palladium catalyst the additition of a small amount of Co can improve the selectivity of acetylene hydrogenation relative to ethene [3b]. Acetylenes and dienes readily polymerize to form carbonaceous deposits on catalysts. Higher temperatures may help to remove these deposits by depolymerization.

5-5

Catalyst deactivation

Qualitative description

Addition of a poison gives the option to enhance the selectivity, although the activity usually decreases. Table 5.2 Examples of catalyst poisons. Type of Poison Surface active metal or ion

High molecular weight product producer

Strong chemisorber Sintering accelerator

Examples Cu in Ni Ni in Pt Fe in Pt Pb or Ca in Co3O4 Pb in Fe3O4 Fe (from pipes) on Cu or Si-Al Acetylenes Dienes H2S NH3 H2O on Al2O3 Cl2 on Cu

Poisoning by sulfur compounds such as H2S is encountered in many large-scale processes using metal catalysts, such as methanation, steam reforming, hydrogenations, and even in hydrodesulfurization (HDS). Chloride on copper catalysts accelerates sintering via the formation of volatile copper oxychlorides or may lead to catalyst loss through oxychlorination.

5.2.2.3 Sintering
Sintering is the loss of catalyst active surface due to crystallite growth of either the support material or the active phase. It can occur due to local heating during preparation (calcination), reduction (fresh or passivated catalyst), reaction (hot spots, maldistribution), or regeneration (coke burn-off). Metal catalysts often consist of small metal crystallites on a porous support material. An important deactivation phenomenon is sintering of these small metal crystallites into larger ones. Of course, this leads to loss of activity. Figure 5.5 shows a schematic representation of this type of deactivation. Initially, the atoms are supposed to be present as small clusters of atoms (or small metal particles), termed a monomer dispersion. Surface diffusion of the atoms will lead to two-dimensional clusters, and upon further diffusion three-dimensional particles will be formed (Fig. 5.5 a). These particles can grow into larger ones through several mechanisms. Particles might move and coalesce (Fig. 5.5 b) or atoms move from one particle to another, either by volatilization or by surface migration (Fig. 5.5 c). The most important mechanism for sintering of small particles often is the movement of atoms rather than particles; depending on the conditions (temperature, type of surface) atoms will move via a hopping mechanism or via surface diffusion. Sintering much resembles crystallization: larger particles 5-6

Catalyst deactivation

Qualitative description

grow at the expense of smaller ones. The position of the particle contributes to sintering. A valley position is stable, while an on-top position is highly unstable.
(a)

monomer dispersion

2-D cluster

3-D particle

(b)

(c)
vapor

particles migrate
migrating

surface

coalesce interparticle transport


metastable

stable

Figure 5.5 Schematic of the various stages in the formation and growth of particles from a monomer dispersion. Sintering is a function of time and temperature. Empirical correlations have been proposed for the catalyst activity as a function of both:

da = ka m dt
E k = k 0 exp a RT
in which a is the catalyst activity. The exponent m in equation (1) often has a value of 2.

(1) (2)

The chemical environment, promoters, and melting point of the active phase or support also play a role in that they change the activation energy. Figure 5.6 shows the effect of sintering on n-heptane reforming (Note the different time scales of the two graphs).

5-7

Catalyst deactivation
m2/g Pt
300

Qualitative description
dp Pt /nm
50 40
60 50

hydrocracking dehydrocyclization isomerization

%
40 30

200

30 20 10

100

20 10

0 10 20 30 40 50 60 70 80

n-C7
0 0 5 10 15 20

time @780 C ( h)

time @780 C (h)

Figure 5.6 Effect of increasing particle size and decreasing surface area with time on the conversion during various reactions in n-heptane reforming; T = 780 oC. Exposure of the catalyst to a temperature of 780 oC for some time increases the size of the platinum crystals gradually. The active surface area, on the other hand, decreases much more rapidly. During reforming of n-heptane various reactions play a role. It can be seen that sintering has a different effect on each of these reactions. For instance, hydrocracking is hardly affected, while dehydrocylization decreases and isomerization increases. Furthermore, the reactivity does not decrease linearly with the loss of metal surface area (decrease in n-heptane conversion from about 93 to 90% with a 10-fold decrease in Pt surface from about 240 to 20 m2/g). Explanations are that the reactions are structure-sensitive, i.e., certain crystal planes are active for certain reactions, so only a small part of the surface area played a role in the first place.

5.2.3 How Deactivation Mechanisms Influence the Reaction Rate Constant


The three main modes of deactivation (fouling, poisoning, and sintering) can be discussed in terms of a generalized reaction rate constant:

k obs = N T k intr

(3)

Here, kintr is the intrinsic rate constant for the reaction (per active site), NT is the total number of active sites, and is the effectiveness factor. (For the following we will assume that = i, i.e., there are no external limitations. Furthermore, isothermal operation is assumed.) The intrinsic rate constant, although presented as a constant, may change as a result of catalyst poisoning; When an impurity in the feed interacts with the catalyst, the chemical nature of the catalyst changes, and thus, the rate constant also changes. An example is the interaction of H2S with transition metals leading to the formation of surface sulfides with a different activity than the original metal

5-8

Catalyst deactivation

Qualitative description

species. The activity of the catalytic site can also completely destroyed (kintr becomes 0 for this site), which results in reduction of the number of active sites. The catalyst effectiveness is influenced by poisoning, whereas the number of available active sites may or may not change; the activity of the sites may change or be completely destroyed. In general, fouling and sintering do not affect the intrinsic rate constant. Fouling results in physical blocking of the active sites and catalyst pores (e.g., by carbon or dust), and hence, affects both the number of active sites and the catalyst effectiveness. Sintering will also affect these variables. Figure 5.7 summarizes the phenomena responsible for deactivation, their influence on kobs.
heat Sintering loss surface area gradual or catastrophic irreversible - nonregenerable process conditions Fouling physical blocking surface by carbon or dust usually regenerable feed & process conditions

kobs
feed conditions Poisoning chemisorption on active sites reversible or irreversible feed & process conditions Selectivity poisons Modifiers block side reactions inhibit consecutive reactions (kinetics)

Figure 5.7 Deactivation phenomena, their causes and effects. Selectivity poisons or modifiers are often discovered by chance and kept as a secret.

5-9

Catalyst deactivation

Poisoning and fouling

5.3 Poisoning and Fouling Quantitative Description


5.3.1 Global View
Reactions leading to catalyst poisoning can be of a several types. Figure 5.8 summarizes the most important main reactions and reactions leading to poisoning.
Fouling or self-poisoning Parallel reaction Series self-poisoning (FCC, HDM) Triangular reaction (FCC) Poisoning Impurity poisoning (TWC)

B A C A A C A P B blocking B B C

Figure 5.8 Reactions in catalyst poisoning; A = reactant; B = product; C = poison, P = impurity. Self-poisoning or fouling occurs when the reactant itself is the poison precursor, either through a parallel reaction (reactant gives product and poison), a series reaction (reactant gives product which gives poison), or a triangular reaction (reactant gives product and both reactant and product give poison). Carbon formation during steam reforming is an example of parallel reactions leading to the formation of the product and a foulant. Steam reforming is a catalytic process in which hydrocarbons such as methane and naphtha react with steam to produce synthesis gas (H2 and CO). The process operates at high temperature, which also favours the decomposition of hydrocarbons into carbon (coke) and hydrogen. Coke formation in FCC can occur both as a result of series reactions and triangular reactions. In FCC a heavy hydrocarbon feedstock (vacuum gas oil) is cracked to produce gasoline using acidic catalysts (zeolites). Polymerization of alkenes produced by cracking of the feedstock alkanes is suspected to be the main mechanism for coke formation during FCC (series reaction). The favourite explanation for coke formation is the condensation of aromatics (both present in the feedstock and formed during cracking) with alkenes [5] (triangular reaction). Figure 5.9 shows a simple triangular reaction scheme and a more complex scheme involving many series and parallel reactions. The schemes shown are so-called lumped kinetic models. The three-lump model considers a gas oil feed (F), a gasoline (G) and a light gas + coke lump (C). In the ten-lump model the gas oil lump is divided in a heavy and a light gas oil (subscripts h and l), which each contain four lumps, i.e., alkanes (P), naphthenes (N), aromatic substituent groups (A) and 5 - 10

Catalyst deactivation

Poisoning and fouling

aromatic rings (CA). The two remaining lumps again are the gasoline (G) and the light gas + coke lump (C). The arrows represent the conversion of one lump into the other with the corresponding rate constants.
F C Ah G

Nh

Nl Ph Pl G

Al CAh CAl

Figure 5.9 Lumped kinetic schemes to describe the FCC process; Three-lump model [6] and ten-lump model [7]; explanation of symbols is given in the text.

The deposition of vanadium and nickel in hydrodemetallization (HDM) of vacuum residues also is an example of a deliberate series self-poisoning reaction. Vanadium is present as a vanadyl-porphyrin (see [4b], [8]), which by consecutive hydrogenations and ring cleavage yields smaller hydrocarbons and vanadium deposits. In HDM the catalyst pore structure and the reactor type (moving bed) are tailored to cope with this. It is also possible that the feed contains an impurity that adsorbs on the catalyst leading to blocking of the catalyst pores. An example is the early three-way-catalyst (in a packed bed of pellets instead of the present monolithic reactors) for the conversion of hydrocarbons, carbon monoxide and nitric oxides in the exhaust of car engines into harmless compounds. Impurities such as lead (TEL, anti-knock) and phosphorus (from lubricant oil) are deposited on the catalyst surface and hence deactivate the precious metals (rhodium, platinum) that catalyze the reactions. This was the major reason to ban lead from gasoline. The concentration profile of the poison in a catalyst particle (see Figure 5.10) as well as in an entire reactor depends on the poisoning mechanism, the reaction kinetics, and mass transport phenomena. In the case of a poison that is deposited on the catalyst independent of the main reaction, at the two limits one finds uniform (or homogeneous) poisoning and pore-mouth poisoning, depending on the diffusion rate of the poison relative to its deposition rate. If diffusion is fast relative to deposition of the poison, a uniform concentration profile is obtained. With lower diffusion rates (or faster reaction rates) the poison molecules can penetrate the interior of the catalyst less deeply before being deposited 5 - 11

Catalyst deactivation

Poisoning and fouling

(diffusion limited poisoning or shell-progressive poisoning). When diffusion is very slow or deposition very fast, only the outer shell of the catalyst is poisoned (pore-mouth poisoning).
increasing poisoning rate

uniform or homogeneous poisoning

diffusion limited poisoning

pore mouth poisoning

center poisoning

series poisoning

Figure 5.10 Poison concentration profile in catalyst particle as a function of poisoning rate. In parallel self-poisoning, the same poisoning profiles are found as in independent poisoning. The only difference is that in this case the profile depends on the relative reaction and diffusion rate of the reactant. Series self-poisoning can result in the poison being concentrated in the interior of the catalyst, if the diffusion rate of species A is high. This mechanism is termed center poisoning or core poisoning. An example is the HDM of vanadyl porphyrins on catalysts with wide pores [8]. Because the concentration of B (the poison precursor) is higher in the center of the catalyst particle (more A converted), more vanadium will be deposited there. It should now be clear that the Thiele modulus concept can be used to quantify the effect of poisoning. QUESTION: The profile of deposited vanadium in catalyst particles depends on the reaction coordinate, i.e., the position of the particle in the axial reactor direction. Explain this. How would the profiles change with the axial coordinate?

5.3.2 Poisoning by Feed-Impurity


In catalyst deactivation by poisoning by an impurity in the feed, the main reaction and the poisoning reaction are independent. Therefore, the deposition of the poison as a function of time can be uncoupled from the reaction kinetics of the main reaction. However, poisoning of the catalyst surface does influence the effectiveness of the catalyst particle. An effectiveness factor for the deactivated catalyst can be obtained in a similar way as for a fresh catalyst particle. The activity of the deactivated catalyst can then be related to the activity of the fresh catalyst at various degrees and types of poisoning. The next two sections will discuss uniform and pore-mouth poisoning, respectively. In uniform poisoning the activity (rate constant) of the pore wall is affected, while in pore-mouth poisoning the available surface area of the pore is reduced. Below, some principles and definitions concerning catalyst effectiveness for diffusion-limited reactions that are not affected by deactivation are summarized. 5 - 12

Catalyst deactivation

Poisoning and fouling

Summary internal diffusion limitations catalyst effectiveness (Chapter 4).


1st order, irreversible, slab:
1.0

C*
0.8 0.6 0.4

1.0

0.1

Effectiveness factor:

i =

tanh

2.0
0.2 0.0 1.0

cylinder
0.4 0.2 0.0

10.0
0.8 0.6

sphere
1

slab

x*

Thiele modulus:

=L

kv Deff

1 L= = Ap a'

Vp

0.1 0.1

10

5.3.2.1 Homogeneous Poisoning Homogeneous poisoning, in which the poison is located uniformly throughout the catalyst particle, is the simplest type of poisoning to model. The activity of the catalyst decreases linearly with the fraction of sites that are poisoned, (i.e., with time). That is, if diffusion limitations are absent for the main reaction. When reactions are affected by both deactivation and diffusion the differential equation that describes the relationship between internal diffusion and reaction rate must solved, corrected for deactivation. For a first order irreversible reaction in a slab-like catalyst particle the equation becomes:

Deff

d 2c = (1 )k v c dx 2

in which k v = N T k intr

(4)

in which is the fraction of poisoned catalyst sites ( = f(t, cP, NT)). cP is the concentration of the poison in the feed. Analogous to the case where no deactivation takes place (Eq. 23, Chapter 4), the solution to the concentration profile inside the catalyst particle becomes (with substitution of kv(1 - ) for kv in Thiele modulus).

x cosh( 1 ) L c = cs cosh( 1 )
in which is the Thiele modulus for an unpoisoned catalyst.

(5)

5 - 13

Catalyst deactivation

Poisoning and fouling

From Eq. 5 the internal effectiveness factor for a deactivated catalyst i,deac can be derived analogously to the internal effectiveness factor without deactivation (see Chapter 4):

i ,deac

dc dx x = L tanh( 1 ) = = (1 )k v c s L 1 Deff

(6)

For the fraction F of observed initial activity the following equation can now be derived:

F=

rpoisoned runpoisoned

i ,deac tanh( 1 ) (1 ) = 1 i tanh

(7)

Figure 5.11 shows the activity decrease as a function of fractional poisoning for various values of the Thiele modulus.
1.0

Fraction of initial activity F

0.8

large 1

0.6

0.4

small
0.2

0.0 0.0

0.2

0.4

0.6

0.8

1.0

Fraction poisoned

Figure 5.11 Fraction of observed initial activity versus fraction of poisoned sites in homogeneous poisoning. For uniform deactivation, is the same everywhere in the catalyst particle. Therefore, for a diffusionfree main reaction, i.e. a small value of , the fraction of initial activity F remaining is simply that given by:

F = 1
This is often referred to as non-selective poisoning.

(9)

For strongly diffusion-limited reactions, i.e., reactions governed by large , the fraction of the observed initial activity becomes:

5 - 14

Catalyst deactivation

Poisoning and fouling (10)

F = 1

This type of poisoning might be called anti-selective. The physical interpretation is that the reaction gradually uses more of the internal surface of the less active poisoned area, i.e., the slower reaction on a poisoned catalyst penetrates deeper into the catalyst particle, because of the decreased value of the Thiele modulus ( 1 ) [9]. So, the activity decrease is less than might be expected. QUESTION: Derive equations 9 and 10.

5.3.2.2 Modeling Pore-Mouth Poisoning


When poisons selectively adsorb on the outer periphery of the catalyst particle this is called poremouth poisoning. Deactivation is much faster than would be expected from the fraction of catalyst covered by the poison. This is due the presence of the completely deactivated outer shell as shown in Figure 5.12 for a slab-like catalyst particle. Poisons showing this behaviour have a large diffusion resistance (high Thiele modulus value poisons). It is assumed that the interface between the poisoned Fraction poisoned outer shell and the clean core is sharp, so that this type of poisoning can be modeled as follows: First, the reactant cd cs diffuses through the completely deactivated layer L, followed by diffusion and reaction in the (1 - )L layer. The rate of the main reaction within the fresh inner core is set equal to the transport rate through the deactivated outer shell at steady state (linear concentration gradient c/L):

0
(1-)L

xd L

Deff (c s c d )

= {A (1 ) L} k v c d i ,in

(11)

Figure 5.12 Pore-mouth poisoning.

in which A is the diffusion area. The effectiveness factor for the fresh inner core i,in (based on the conditions at x = xd), after elimination of cd, is given by:

i ,in ==
QUESTION:

tanh{(1 ) } (1 )

(12)

Why (1-) instead of (1-)? Hint: what is the length of the slab available for reaction in uniform and in pore-mouth poisoning?

The fraction of initial activity is defined as:

5 - 15

Catalyst deactivation

Poisoning and fouling

F=

rpoisoned runpoisoned

i ,in c d (1 ) i c s

(13)

Rearranging equation (11), eliminating cd, and substituting Eqs. (11) and (12) in (13) yields:

F=

tanh{(1 ) } 1 1 + tanh{(1 ) } tanh ( )

(14)

In the case of negligible diffusion resistance for the main reaction ( small), equation (14) reduces to:

F = 1
while equation (16) is obtained when diffusion limitations are important ( large):

(15)

F=

1 1 +
Prove that the limits in equations (15) and (16) are obtained.

(16)

QUESTION:

Figure 5.13 shows the fraction of initial activity as a function of the fraction poisoned for different values of .
1.0

Fraction of initial activity F

0.8

0.01
0.6

3
0.4

10
0.2

100
0.0 0.0 0.2 0.4 0.6 0.8 1.0

Fraction poisoned

Figure 5.13 Fraction of initial activity versus fraction of poisoned sites in pore-mouth poisoning. When comparing Figures 5.12 and 5.13 the difference between the two is obvious. For diffusionlimited reactions in the case of pore-mouth poisoning rapid deactivation takes place already when a relatively small fraction of active sites is poisoned. This is often referred to as selective poisoning. The physical explanation for the large drop in activity is that on the unpoisoned catalyst, the reactions took only place in the outer catalyst shell (outer 1% for = 100). Now, poisoning the outer 10% of the catalyst surface has destroyed the original sites of reaction. Therefore, the reactants are forced to 5 - 16

Catalyst deactivation

Poisoning and fouling

diffuse deeper into the particle before they reach the active catalyst parts. Since diffusion over this longer distance is slow, the observed rate drops severely.

5.3.2.3 Non-uniform Catalyst Distribution for increased Catalyst Life


In the absence of diffusion limitations usually a catalyst with a uniform activity profile is aimed for, i.e., the active phase should be homogeneously distributed over the catalyst particle. This way, the catalytic activity per unit volume of catalyst bed is maximized. In Chapter 4 we have implicitly assumed a uniform activity profile. However, if the main reaction is accompanied by strong diffusion limitations, the reactants will never reach the interior of the catalyst particle, leading to a low catalyst effectiveness. Obviously, catalytic material (which is often expensive) may be saved by only covering an outer shell with it (the egg-shell catalyst, see Figure 5.14). Even more spectacular advantages of non-uniform activity distributions may be gained when an irreversible poisoning reaction accompanies the main reaction. The main factors in choosing the best activity distribution are the Thiele moduli of the main reaction and the poisoning reaction. Figure 5.14 presents a catalyst selection chart on this basis, to obtain the longest catalyst life.
10

Egg shell
3 m = R km Deff ,m
1

Egg white

Uniform
0.1 0.1 1 10 40

Egg yolk
3 p = R kp Deff ,p

uinform poisoning

pore mouth poisoning

Figure 5.14 Catalyst selection for a reaction accompanied by a poisoning reaction. Subscript m refers to main reaction; p refers to poisoning reaction. Criterion: longest catalytic life with effectiveness of 0.4 or higher. Adapted from [10]. Four limiting situations can exist: 1. Poisoning is uniform and the main reaction is diffusion-free. In this case the uniform activity distribution is the best choice. 2. Poisoning is uniform and the main reaction is strongly diffusion limited. An egg-shell catalyst is preferred. 3. Pore-mouth poisoning occurs and the main reaction is diffusion-free. It can be advantageous to concentrate the active material in the interior of the catalyst (egg-yolk catalyst). When the support material is capable of adsorbing the poison, the active catalyst components are thus protected from it [11]. 5 - 17

Catalyst deactivation

Poisoning and fouling

4. Pore-mouth poisoning occurs and the main reaction is strongly diffusion limited. In this case an egg-white catalyst, in which the active material is located in the middle with an inert core and shell, might be the best alternative. Since the main reaction is diffusion-limited, the reactants will not reach the very interior of the catalyst, making an egg-yolk type unsuitable. On the other hand, the poison will be concentrated in the outer shell. This leaves the middle part of the catalyst the location of action for the main reaction. Although the most suitable activity distribution can increase the productivity per unit volume of reactor, and hence decrease costs, one must also consider the catalyst production costs. Obviously, it is more difficult and costly to produce a catalyst with an activity profile other than uniform.

5.3.3 Modeling Coke formation


Self-poisoning often involves the formation and deposition of coke on the catalyst. Coke consists of carbonaceous material, usually with high molecular weight and low H/C ratio. Coke poisoning generally is reversible because regeneration is possible by oxidative treatment with air, oxygen, steam, or carbon dioxide. Self-poisoning or fouling is much more complex than the poisoning by feed impurities treated in the previous sections. In the considerations presented above it was assumed that the pore dimensions remain unaltered. However, upon deposition of significant amounts of material diffusion becomes hindered and eventually pore blocking may occur. The fouling problem requires the simultaneous solution of diffusion and reaction equations. Several authors have provided numerical as well as analytical solutions to the self-poisoning equations for fouling through series and parallel reactions (see [12]). Even a remotely detailed treatment is beyond the scope of this book, but some general remarks can be made. For parallel fouling Figure 5.15 shows a logarithmic plot of the internal effectiveness factor (which is a measure for activity) as a function of time and Thiele modulus. At low values of the Thiele modulus the initial activity is high. With increasing Thiele modulus there is a continuing decrease in activity of the fresh catalyst. However, the curves cross each other after some time. Hence, the residual activities of catalysts with higher Thiele moduli is higher after some time. Thus a catalyst with some diffusional resistance appears to be preferred in view of stability over a long process period [12]. In series fouling the extent of fouling increases continuously with increasing Thiele modulus for all values of process time. Therefore, a catalyst with the least diffusion resistance is preferred [12].

5 - 18

Catalyst deactivation
increasing

Poisoning and fouling

ln i,deac

ln t

Figure 5.15 Internal effectiveness factor (catalyst activity) as a function of time and Thiele modulus for a particle subjected to parallel fouling. Coke can also be formed by a precursor in the feed that is not involved in the main reaction. Then, the deactivation process can be treated similar to poisoning by impurities.

5.3.3.1 Kinetics of Independent Coke Formation When coke deposition takes place uniformly throughout the catalyst particle and independently of the main reaction, e.g., due the presence of impurities in the feed that lead to the formation of coke, the reaction rate of the main reaction can be expressed as follows:

r = C r 0

(17)

The reaction rate is thus separated into a part that is dependent on the reaction kinetics of the main reaction r0 and a factorc correcting for the deactivation caused by coke formation. c can be expressed as either a function of coke concentration, precursor concentration, or time-on-stream. The latter is usually most convenient, since it provides a relation for the time-on-stream activity development. Voorhies [13] proposed the following empirical relationship between coke formation on the catalyst and time-on-stream:

cC = t

with : 0.5 (0.3 - 1)

(18)

in which cc is the concentration of coke on the catalyst (kg/kg) and and are constants. This expression is a classical one, since it was one of the first attempts to describe coke deposition. Various empirical functionalities between the activity and the amount of coke on the catalyst have been observed. One commonly used form is:

5 - 19

Catalyst deactivation

Poisoning and fouling

C =

1 1 + cC

(19a)

or in terms of time:

C =

1 1 + 't

(19b)

in which and (= ) are constants. Table 5.3 summarizes the most frequently used empirical deactivation functions. Note that time-onstream is used as the variable. Table 5.3 Examples of deactivation functions [14]; , , and n are constants. C = 1 t C = exp( t ) 1 C = 1+ t C = t

C = (1 + t ) n

QUESTION:

Check the limits of the functions in Table 5.3 for t 0 and t . Do you notice anything strange?

For practical purposes expressions such as in Table 5.3 are most convenient. Intellectually, however, they are less satisfactory, as they are empirical rather than based on theory. Moreover, in practice their usefulness is limited because of lack of predictive value. In literature many attempts have been described to derive theoretical models for coke deposition. A logical step is to correlate C to the kinetics of coke formation:

rC =

d C dt

(20)

in which rC is the mass of coke formed per mass of catalyst per unit time. The coke deposition kinetics can be determined experimentally.

5 - 20

Catalyst deactivation 5.3.3.2 Kinetics of Dependent Coke Formation Self-Poisoning

Poisoning and fouling

Coke formation usually is a result of reactions of the reactants and/or products of the desired reactions (fouling). These reactions (e.g., in hydrodesulfurization, steam reforming, etc.) typically follow Langmuir-Hinshelwood-Hougen-Watson (LHHW) kinetics (see Chapter 3). The kinetic rate equation without deactivation has to be corrected for coke deposition (Eq. (17)). For reaction of A to produce B and C (coke), the rate of disappearance of A is given by equation (21):

rA =

0 kA N T K A C p A 1 + K A p A + K B pB

(21)

QUESTION:

Which step is the rate-determining step (adsorption of A, reaction, or desorption of B?)

Parallel coking A C and series coking (A B C) yield equations (22) and (23), respectively, for the rate of coke formation, rC.

A B

rC =

0 k AC N T K A C p A 1 + K A p A + K B pB

(22)

rC =

0 k BC N T K B C p B 1 + K A p A + K B pB

(23)

In the case of series coking, relating C to the time by substituting Eq. (23) in Eq. (20) and integrating, yields the coke deactivation function:

C = exp

0 k BC NT K B pB dt 0 1 + K A p A + K B pB t

(24)

From equation (24) it is clear that C is a function of time, concentrations of A and B, and reactor coordinate (concentrations change along the reactor). If a function from Table 5.3 had been used, only the time dependence would have been accounted for. This assumes deactivation that is uniform in a catalyst pellet and in the reactor and independent on the local concentration in pellet and reactor. The variables are then lumped in one constant ( in the exponential function of Table 5.3). The deactivation as a function of time would then be valid for the conditions at which it was measured only, and should not be used for the prediction of deactivation at other conditions.

5 - 21

Catalyst deactivation

Poisoning and fouling

Although using only the time dependence in modeling of reactions involving coke deposition may produce serious errors in predictions applying to other situations, it is the most convenient method to apply.

5 - 22

Catalyst deactivation

Notation

Notation
a a Ap c Deff Ea F k kintr kv K L n NT p rv rv, chem rv, obs r0 R R t T x Greek activity specific surface area of catalyst particle external surface area of particle concentration effective diffusivity in particle activation energy fraction of initial activity rate constant for catalyst aging intrinsic rate constant rate constant per unit particle volume adsorption equilibrium constant characteristic catalyst dimension reaction order total number of sites pressure reaction rate per unit particle volume intrinsic reaction rate per unit particle volume observed reaction rate per unit particle volume reaction rate of main reaction universal gas constant particle radius time temperature coordinate m-1 m2 mol m-3 m2 s-1 J mol-1 s-1 (m3 mol-1)n-1 s-1 (mp3 mol-1)n-1 s-1 atm-1 m Pa mol s-1 mp-3 mol s-1 mp-3 mol s-1 mp-3 mol s-1 mp-3 J mol-1 K-1 m s K m

fraction of poisoned catalyst sites or constant in Table 5.3 constant constant in Table 5.3 overall effectiveness factor internal effectiveness factor Thiele modulus decay function

i
C

i ,deac internal effectiveness factor of deactivating catalyst

Subscripts app b apparent in bulk phase 5 - 23

Catalyst deactivation C d chem eff i in m obs p s 0 carbon at surface of deactivation chemically controlled effective intraparticle, internal inner core main reaction observed particle or poison at external particle surface at reactor entrance

Notation

Superscripts 0 without deactivation

Literature
1. P. Chen et al., Growth of carbon nanotubes, Carbon 35(10-11), 1495 (1997). 2. C. de Bellefon and P. Fouilloux, Homogeneous and heterogeneous hydrogenation of nitriles in a liquid phase: chemical, mechanistic, and catalytic aspects, Catal. Rev. Sci. Eng. 36(3), 459 (1994). 3. C.S., Satterfield, Heterogeneous catalysis in industrial practice, 2nd ed., McGraw-Hill, USA, 1991. a) Chapter 4; b) Chapter 6. 4. J.A. Moulijn, M. Makkee and A.E. van Diepen, Process Technology, lecture notes TU Delft, 1998.a) Chapter 3; b) Chapter 2. 5. D.L. Trimm, Deactivation and regeneration, in: Handbook of heterogeneous catalysis, vol 3, G. Ertl, H. Knzinger and J. Weitkamp (eds.), VCH, Weinheim, 1997, p.1263. 6. V.W., Weekman, AIChE Journal 16(3), 397 (1970). 7. S.M. Jacob, B. Gross, S.E. Voltz and V.W. Weekman, AIChE Journal 22(4), 701 (1976). 8. J.P., Janssens, R.M. de Deugd, A.D. van Langeveld, S.T. Sie and J.A. Moulijn On the metal deposition process during the hydrodemetallation of vanadyl-tetraphenylporphyrin, in: Catalyst deactivation 1997, C.H. Bartholomew and G.A. Fuentes (Eds.), Stud. Surf. Sci. Cat. 111, 283 (1997). 9. A. Wheeler, Reaction rates and selectivity in catalyst pores, Advances in Catalysis III, 249 (1951). 10. E.R. Becker and J. Wei, Nonuniform distribution of catalysts on supports, J. Catal. 46, 372-381 (1977). 11. S.T. Sie, M.M.G. Sender and H.M.H. Wechem, Catal. Today. 8(1), 371 (1990). 12. L.K. Doraiswamy and M.M. Sharma, 'Heterogeneous Reactions: Analysis, Examples and Reactor Design, Vol. 1: Gas-Solid and Solid-Solid Reactions', Chapter 8, John Wiley and Sons, New York, 1984. 5 - 24

Catalyst deactivation

Literature

13. A. Voorhies, Ind.Eng.Chem. 37, 318 (1945). 14. G.F. Froment and K.B. Bischoff, Chemical Reactor Analysis and Design, John Wiley, New York, 2nd ed., 1990.

5 - 25

Selectivity

Introduction

SELECTIVITY

6.1 Introduction
The selectivity of a catalyst towards a desired product is of major interest when multiple products can be formed thermodynamically. A striking example is synthesis gas technology (C1 chemistry). Depending on the catalyst used different products can be synthesized (see Figure 6.1).
CH4 Ni H2 / CO Fe or Co CnH2n+2 CnH2n Cu + Co CnH2n+1OH (n = 1 - 6) CH3OH Cu or Pd

Figure 6.1 Product selectivity control by catalyst in synthesis gas applications. Over Ni only methane is formed. Long-chain hydrocarbons are produced in the Fischer-Tropsch synthesis. This process is mostly used for the production of middle distillates from synthesis gas, a novel process, in which linear alkanes in the diesel range are produced [1]. The reaction selectivity is dominated by the nature of the active catalyst sites and the reaction kinetics. Therefore, reaction conditions are very important. Often the pore structure and pore size of the catalyst may introduce geometric constraints and diffusion limitations that affect selectivity. From the previous chapter it will be obvious that catalyst deactivation may also influence selectivity. Diffusion limitations (as discussed in Chapter 4) are often disadvantageous, for instance in series reactions in which the intermediate is the desired product. However, sometimes diffusion limitations are introduced deliberately. Shape-selectivity in zeolitic catalysts is a well-known example.

6.2

Reaction Selectivity

Consider a reactant A that is converted to the desired product B and an undesired product C. The products can be formed through parallel or series reactions (or both):

B
1. Parallel reactions

A C

2. Series reactions

6-1

Selectivity

Reaction selectivity

Examples of situation 1 are the conversion of methane, which may yield various products, and the parallel oxidation of ethene to ethene oxide and carbon dioxide plus water. Situation 2 is very common and includes partial hydrogenation and oxidation reactions, in which the intermediate product is the desired one. Ethene oxidation to form ethene oxide is also part of this class, because ethene oxide can be further oxidized. Ethene oxidation in fact is an example of a triangular scheme:

B
3. Triangular reactions

A C

Selectivity is also affected when different reactants react on the same catalyst:
4. Independent reactions

A P

B Q

Situation 4 occurs when it is desired to convert one type of compounds from a mixture, while leaving the others unaffected. An example is a selective catalyst that will hydrogenate alkenes in the presence of aromatics, while leaving the aromatics unchanged. In all cases the selectivity for product B can be controlled by the intrinsic kinetics of the reaction or by mass transport phenomena. In the next sections, we will assume that external mass transport limitations are absent (cs = cb). Only internal diffusion limitations will be treated. Furthermore, temperature gradients will be neglected.

6.2.1 Parallel Reactions


For irreversible parallel reactions with a common reactant:

kvB A B kvC A C

n rvB = k vB c A m rvC = k vC c A

the reaction selectivity is :

rvB nm = S cA rvC

(1)

S = kvB / kvC is the kinetic selectivity factor. This selectivity factor is based on the intrinsic rate constants and is determined by the temperature through the activation energies of the reactions. For reactions of different order the concentration of the reactant also affects the selectivity. The reaction of highest order is favoured with increasing concentration. When pore diffusion limitations of A in a catalyst particle are an issue, we have:

6-2

Selectivity

Reaction selectivity

rvB ,obs rvC ,obs

JB JC

(2)
x=L

in which JB and JC represent the fluxes of B and C out of the catalyst particle. Slow diffusion of A will lower the concentration inside the particle, which will only have effect for reactions of different order. The reaction with the lowest order is favoured by diffusion limitations. Hence, whether diffusion limitations are desired or undesired depends on whether the desired reaction has the lower or the higher order in A. QUESTIONs: Proof that the selectivity is not influenced in case of equal reaction orders for strong diffusional limitations ( large, see also Chapter 4). Explain the physical background of the phenomenon that higher order reactions are affected more by diffusion limitations.

6.2.2 Series Reactions


Series reactions in which the intermediate product is the desired one, are very common in the chemical industry. Think, for example, of partial hydrogenations and oxidations. In the latter it is often difficult to prevent part of the product being completely oxidized. 6.2.2.1 Modeling Consider the consecutive reaction:

kvB kvC A B C
with the following first-order rate expressions:

rvA = k vB c A rvB = k vB c A + k vC c B
Assuming equal diffusivities, the pore diffusion equations are written as follows:

(3) (4)

Deff Deff

d 2cA = k vB c A dx 2 d 2 cB = k vC c B k vB c A dx 2

(5) (6)

Solving equation (5) and (6) for the concentrations inside a slab-like catalyst particle yields:

6-3

Selectivity

Reaction selectivity

c A = c As

cosh( B x / L) cosh( B )

(7)

k vB cosh( C x / L) cosh( B x / L) cosh( C x / L) c B = c As + c Bs cosh( B ) cosh( C ) k vB k vC cosh( C )

(8)

in which cAs and cBs are the surface concentrations of A and B, respectively (equal to the local gasphase concentrations, no external limitations play a role). From equations (7) and (8) the fluxes of A into and B out of the external slab surface can be derived (differentiate with respect to x). The yield of B for the slab then equals the ratio of the fluxes of A and B (see [2] for a more detailed treatment):

JB JA

=
x= L

dc B k vB C tanh( C ) c Bs C tanh( C ) = 1 dc A k vB k vC B tanh( B ) c As B tanh( B )

(9)

For the limits of kinetic control (large pores) and diffusion control (small pores) equation (9) can be simplified to equation (10) and (11), respectively. These equations represent the local or instantaneous selectivity in a reactor.

JB JA JB JA

=
x=L

dc B 1 c Bs = 1 dc A S c As dc B S 1 c Bs = dc A 1 + S S c As

(10)

=
x=L

(11)

with S = kvB / kvC. QUESTION: Derive equations (9), (10), and (11) from equations (7) and (8).

Integration of equations (10) and (11) using the integrating factor gives the yield on B at the end of a reactor bed (cB0 = 0, cA0 = 1 are concentrations at zero conversion). For the two limiting cases of chemical control and diffusion control the solution, i.e., the yield on B as a function of the concentration of A, is represented by equations (12) and (13), respectively.

c Bs S c As = c A0 S 1 c A0

c As c A0
c As c A0

1 S S

1
1

(12)

c Bs S c As = c A0 S 1 c A0
6-4

1 S S

(13)

Selectivity

Reaction selectivity

Figure 6.2 shows the yield on B as a function of the fraction of A converted for these two limiting cases (S = 4).
0.7

Fraction A converted to B

0.6

Chemical control

0.5

0.4

0.3

0.2

0.1

Diffusion control

0 0 0.2 0.4 0.6 0.8 1

Fraction A converted

Figure 6.2 Effect of diffusion limitiations (pore size) on yield of intermediate B in consecutive reaction; S = 4. Chemical control (or kinetic control) applies to catalyst with large pores or to catalysts that are nonporous, while diffusion controls in catalysts with small pores. QUESTION: Calculate the maximum yield on B and the conversion of A at which this yield is attained for both cases.

6.2.2.2 Example: Butene Dehydrogenation A technically important example of a consecutive reaction in which the intermediate is the desired product is the dehydrogenation of butene to form butadiene:
k1 k2 = == C4 C4 ' coke'

Butadiene is highly unstable and can polymerize and crack to coke and other undesirable by-products. Figure 6.3 shows the selectivity towards butadiene at a butene conversion of 35% as a function of the particle diameter.

6-5

Selectivity
Selectivity butadiene at 35% conversion
100 90 80 70 60 50

Reaction selectivity

10

particle diameter (mm)

Figure 6.3 Butadiene selectivity versus particle diameter. Adapted from [3] At small particle diameters the selectivity remains constant, while with particle diameters of over about 3 mm the selectivity decreases. QUESTION: Does the result of Figure 6.3 agree with theory? Explain.

6.2.3 Independent Reactions


For independent first-order reactions with B the desired product:

kvB A B kvQ P Q

rvB = k vB c A rvC = k vQ c P

the diffusion model for a slab-like catalyst particle assuming equal diffusivities yields the concentrations of A and P inside the pores:

c A = c As

cosh( B x / L) cosh( B )

(14)

c P = c Ps

cosh( Q x / L) cosh( Q )

(15)

where cAs and cPs are the concentrations of A and P at the entrance of the pores (surface concentration = bulk concentration). The reaction selectivity can be written as: 6-6

Selectivity

Reaction selectivity

rB ,obs A k vB c As = rQ ,obs P k vQ c Ps

j =

tanh j

(16)

with the limiting cases of small j (kinetic control):

rB ,obs rQ ,obs

k vB c As c = S As k vQ c Ps c Ps

(17)

and large j (diffusion control):

rB ,obs rQ ,obs

k vB c As k vQ c Ps

= S

c As c Ps

(18)

QUESTION:

Prove these limits.

Despite the independence of the desired and undesired reaction, diffusion limitations always disfavour the desired reaction (presuming that its rate constant without diffusion limitations is generally larger than that of the undesired reaction). The reason is that in this situation, only a relatively small fraction of the surface is used for the faster of the two reactions, while a larger fraction is available to the slower reaction. Hence, the fast reaction is retarded more than the slow reaction.

6.2.4 Temperature Effects


The reaction selectivity in parallel and independent reactions is proportional to the selectivity factor S or its square root. This selectivity factor is defined as the ratio of the rate constants of the desired and undesired reaction (kdes / kundes). Obviously, with increasing ratio of the rate constants, and hence the selectivity factor, the selectivity towards the desired reaction increases. In series reactions, the selectivity towards the desired intermediate product is a somewhat more complex function of the selectivity factor, but for such reactions it is also true that with increasing selectivity factor, the selectivity for the desired reaction increases. The effect of temperature is to increase or decrease both the rate constant for the desired reaction and that for the undesired reaction. The influence on the selectivity depends on the activation energies (Ea) for the competing reactions. Expressing S in terms of activation energies and temperature shows that the largest activation energy dominates:

S=

Ea ,undes E k des exp a ,des k undes RT

(19)

The selectivity factor changes with temperature depending on which activation energy is the largest. When the temperature rises: 6-7

Selectivity

Reaction selectivity

S increases if Ea,des > Ea,undes S decreases if Ea,des < Ea,undes Thus the reaction with the largest activation energy is more temperature-sensitive. A general rule can be deduced from the foregoing for the influence of temperature on competing reactions, viz., a high temperature favours the reaction of highest activation energy, while a low temperature favours the reaction of lowest activation energy. QUESTION: Apply the above principle to the competing reactions discussed in the previous sections.

6.3 Shape-Selectivity
An interesting type of selectivity occurs when the catalyst pores are of molecular dimensions. Then so-called shape-selectivity is possible. Shape-selectivity can be distinguished in reactant selectivity, transition-state selectivity, and product selectivity. In general it can be said that there is a large penalty on branching or bulkyness. Even small changes in molecular dimensions can alter the diffusivity enormously.

6.3.1 Reactant Selectivity


An example of reactant selectivity (size exclusion) is found in the selective cracking in a ZSM-5 zeolite of normal alkanes and alkenes from a gasoline to produce mostly propane (Selectoforming process [4]). This simultaneously removes the components with the lowest octane number and produces LPG as a valuable by-product [5].

Figure 6.4 Reactant selectivity in Selectoforming process.

6.3.2 Transition-state Selectivity

6-8

Selectivity

Shape selectivity

In transition-state selectivity, a product is selectively produced because the transition-state leading to the other possible product(s) does not fit in the catalyst pores. An example is the production of cis-4tert-butylcyclohexanol by the reaction of 4-tert-butylcyclohexanone with isopropanol in zeolite BETA (Figure 6.5) [6].
4-tert-butylcyclohexanone + isopropanol cis-4-tert-butylcyclohexanol

O + CH3 H C OH CH3

BETA + 350 K CH3 C O CH3

OH

Figure 6.5 Selective production of cis-isomer. The cis-isomer is formed preferentially because the transition state leading to the trans-isomer does not fit inside the catalyst pores. For a more detailed treatment of this reaction, see Chapter 8.

6.3.3 Product Selectivity


The disproportionation of toluene to selectively form para-xylene is an example of productselectivity. Figure 6.6 shows the principle.

Deff,rel. >10000 + >10000


H3C CH3 CH3 CH3 H3C CH3

Deff,rel. 10000

>10000 ZSM-5

CH3

1
CH3

Figure 6.6 Shape selectivity in the disproportionation of toluene. Adapted from [5] The selective formation of p-xylene from toluene depends on the differences in diffusion rates of the isomeric xylenes in the ZSM-5 zeolite. Without diffusion control, the xylenes would be produced according to their thermodynamic equilibrium. However, p-xylene diffuses much faster than the two 6-9

Selectivity

Shape selectivity

others, and therefore becomes the preferred product when the catalyst pores are narrow and diffusion becomes limiting. An unusual example of product selectivity, one that does not involve a zeolite catalyst, is the production of methylamines by reaction of ammonia with methanol over an amorphous Si-Al catalyst. The problem with the conventional process is that the least desirable product (trimethylamine, TMA) is also the thermodynamically most favoured product. The selectivity towards monomethylamine (MMA) and especially dimethylamine (DMA) can be increased by making use of the different sizes and thus diffusivities of the three amines (see Figure 6.7).
ordinary Si-Al catalyst

MeOH + NH3 MeOH + MMA MeOH + DMA

MMA (0.4 nm) DMA (0.45 nm) TMA (0.5 nm)

Carbon molecular sieve layer (~ 0.5 nm pores)

Figure 6.7 Principle and reactions of selective MMA and DMA production. Coating the standard Si-Al catalyst with a carbon molecular sieve layer with pores of about 0.5 nm (roughly the size of TMA) produces a composite catalyst that allows higher than normal selectivities to MMA and DMA [7]. Chapter 8 gives a more detailed treatment of the processes described above and also some other examples of product selectivity.

6.4 Catalyst Modification for improved Selectivity


A slight improvement in selectivity to the desired product(s) often is of great economic significance. Myriad possibilities exist to improve the selectivity. First of all, improvement of the selectivity can be due to an increase of the intrinsic (kinetic) selectivity, by improving adsorption and/or desorption characteristics or the surface reaction (see Chapter 3). Improvement of the adsorption/desorption characteristics of the catalyst can be achieved, e.g., by modification of the active phase (addition of Ga to isomerization catalyst to facilitate desorption of H2) or the support material (in deep hydrodesulfurization) or by using a completely different active phase or support (methane reforming). In addition, the catalytic surface can be modified (for better selectivity to one the enantiomers), a different fluid phase or solvent (in 6 - 10

Selectivity

Catalyst modification

alkylation, nitrile hydrogenation) can be used or a fluid phase added (production of cyclohexene). Finally, shape selectivity (by means of pore architecture) can be applied. The production of cyclohexene by selective hydrogenation of benzene illustrates the applicability of the addition of an extra fluid phase to enhance the selectivity. Cyclohexene is the intermediate product in a series hydrogenation reaction:
+ 2 H2 + H2

Conducting this reaction in a single organics phase (either gas or liquid) on typical hydrogenation catalysts (e.g., Co, Pt, Pd, Rh, Ru) produces low yields of cyclohexene [8], because either the conversion (activity) or the selectivity is too low. Ruthenium catalysts are active hydrogenation catalysts, but also for this catalyst selectivity is a problem. The use of a two-liquid system in which one of the fluid phases contains the hydrocarbons and the other is an aqueous solution of a salt like iron or zinc sulfate (see Figure 6.8) clearly improves the selectivity [8].
Ru Salt-water

Organics

Figure 6.8 Reaction system for partial benzene hydrogenation. The main effect of the salt solution is that it provides a diffusion barrier through which benzene diffuses faster than cyclohexene as a result of the lower solubility of cyclohexene in water. Readsorption of cyclohexene on the catalyst once it has been formed is made slow, resulting in low hydrogenation rate of cyclohexene and hence a high selectivity [8]. The selectivity of the surface reaction can be affected by coupling of reactions (see Chapter 3 Case studies) and transient operation. Another way to enhance the selectivity is to introduce shape selectivity by creating pores of molecular sizes, either inside the catalyst or around it. Examples are so-called transition-state confinement (e.g., in acylation catalysts), the use of membrane coatings, and the deliberate introduction of diffusion limitations as discussed in Section 6.3. The distribution of the active catalyst ingredient over the catalyst particle can also affect selectivity. For instance, it appears that in series reactions (A B C) in which B is the desired product, a catalyst with an activity that decreases towards the center of the catalyst particle is optimal [9]. The extreme of this distribution is the egg-shell catalyst in which catalytically active material only is present in the outer shell of the catalyst particle (see Section 5.3). The same is valid for independent reactions when the desired reaction has the highest rate constant. 6 - 11

Selectivity

Catalyst modification

QUESTION:

Explain why an egg-shell catalyst is preferred for a series reaction with kB >> kC and equal diffusivities of all reaction components.

6.5 Effect of Catalyst Deactivation on Selectivity


Catalyst deactivation is not only responsible for decreasing activity, but also can cause lower selectivities. First of all, often the operating temperature is increased during time-on-stream to compensate for activity loss. If the undesired reactions have a higher activation energy than the desired reactions (which is very common) the selectivity will decrease. Selectivity changes as a direct result of catalyst deactivation are also encountered frequently. The three most important mechanisms of deactivation are aging, poisoning, and fouling (see Chapter 5).

6.5.1 Effect of Aging on Selectivity


An example of the loss of selectivity due to aging is the catalytic treatment of exhaust gases in a threeway catalyst. One of the aims of exhaust gas treatment is the reduction of the amount of NO. Conversion of NO can be by the desired reaction: 2 NO + 2 CO N2 + 2 CO2 or by the undesired reaction: NO + CO N2O + CO2 in which N2O, also a harmful component, is formed. This component can also be reduced by CO: N2O + CO N2 + CO2 Figure 6.9 shows the effect of aging on the selectivity for the undesired N2O forming reaction.

NO conversion to N2O

Aging

~400

~700

Temperature (K)

Figure 6.9 Effect of aging on NO conversion to N2O 6 - 12

Selectivity

Deactivation

The undesired reaction leading to N2O production has the highest apparent activation energy and starts at higher temperature (ca. 300 K on fresh catalyst) than the desired reaction. At low temperature (ca. 300 400 K), the reduction of the formed N2O by CO is inhibited due to stronger adsorption of NO, but at higher temperature reduction does take place, resulting in the profile around 400 K of Figure 6.9. Because of activity loss due to aging, the temperature in the reactor has to be increased with time-onstream in order to maintain a sufficient activity (rate constants increase). However, this results in a penalty on the selectivity, and more N2O is formed. The window of N2O formation is broadened and shifted to higher temperature. Nearly quantitative conversion to N2O results at high operating temperature.

6.5.2 Effect of Poisoning and Fouling on Selectivity


Two extreme types of poisoning are uniform poisoning, in which the poison is distributed homogeneously throughout the catalyst, and pore-mouth poisoning, where the poison is concentrated in the outer catalyst shell. Uniform poisoning does not affect the selectivity when both the desired and the undesired reaction take place on the same catalytic site. However, if the reactions require different sites or on ensembles of different size, and the poison selectively adsorbs on one of these sites, the selectivity might be influenced in a negative or positive way (modifier, see Section 5.2.2.2). When pore-mouth poisoning occurs, reactions with the highest Thiele modulus (fast reactions) are affected most. Pore-mouth poisoning can be disastrous for selectivity in series as well as independent parallel reactions, since the desired reaction generally has the largest rate constant. Fouling or self-poisoning decreases the selectivity in its own right because the foulant, usually coke, is either produced from the reactant to be converted to a desired product or from the desired product itself.

Notation
c Deff Ea Ji kv L m n concentration effective diffusivity in particle activation energy flux of component i rate constant per unit particle volume characteristic catalyst dimension reaction order reaction order mol m-3 m2 s-1 J mol-1 mol m-2 s-1 (mp3 mol-1)n-1s-1 m -

6 - 13

Selectivity rv rv, chem rv, obs R S T x reaction rate per unit particle volume intrinsic reaction rate per unit particle volume observed reaction rate per unit particle volume universal gas constant selectivity factor temperature coordinate mol s-1 mp-3 mol s-1 mp-3 mol s-1 mp-3 J mol-1 K-1 K m

Notation

Greek

= i, overall effectiveness factor without external limitations Thiele modulus

Subscripts b chem des eff i j obs p s undes 0 in bulk phase chemically controlled desired effective intraparticle, internal applying to component j observed particle at external particle surface undesired at reactor entrance

Literature
1. S.T. Sie, M.M.G. Senden and H.M.H. Wechem, Catal. Today. 8(1), 371 (1990). 2. A. Wheeler, Reaction rates and selectivity in catalyst pores, Advances in Catalysis III, 249 (1951). 3. H.H. Voge and C.Z. Morgan, Effect of catalyst particle size on selectivity in butene dehydrogenation, Ind. Eng. Chem. Process Des. Develop. 11(3), 454 (1972). 4. N.Y. Chen, J. Maziuk,, A.B. Schwartz and P.B. Weisz, Selectoforming, Hydroc. Proc. 49(9) 192 (1970). 5. P.B. Weisz, Molecular shape selective catalysis, Pure & Appl. Chem. 52, 2091 (1980). 6. E.J. Creyghton, S.D. Ganeshie, R.S. Downing and H. van Bekkum, Stereoselective MeerweinPondorf-Verley and Oppenauer reactions catalysed by zeolite Beta, J. Mol. Catal. A-Chem. 115(3), 457 (1997). 7. H.C. Foley, D.S. Lafyatis, R.K. Mariwala, G.D. Sonnichsen, and L.D. Brake, Shape selective methylamines synthesis Chem. Eng. Sci. 49(24A), 4771 (1994). 6 - 14

Selectivity

Literature

8. M. Soede, The partial hydrogenation of aromatics, Thesis, TU Delft, 1996. 9. L.K. Doraiswamy and M.M. Sharma, 'Heterogeneous Reactions: Analysis, Examples and Reactor Design, Vol. 1: Gas-Solid and Solid-Solid Reactions', Chapter 5, John Wiley and Sons, New York, 1984.

6 - 15

Multiphase reactors

STRATEGIES FOR MULTIPHASE REACTOR SELECTION

7-1

Pergamon

Chemical Engineering Science, Vol. 49, No. 24A, pp. 4029 4065, 1994 Copyright 1995 Elsevier Science Ltd Pnnted in Great Britain. All rights reserved 0(~9-2509/94 $7.00+0.00

0009-2509(94)00262-2

STRATEGIES FOR MULTIPHASE REACTOR SELECTION


R. K R I S H N A * Department of Chemical Engineering, University of Amsterdam, Nieuwe Achtergracht 166, 1018 WV Amsterdam, The Netherlands and S. T. SIE Faculty of Chemical Technology and Materials Science, Delft University of Technology, Julianalaan 136, 2628 BL Delft, The Netherlands

(Received 16 June 1994; accepted for publication 10 August 1994)


Abstract--The central theme addressed in this paper is: how do we arrive at the "ideal" reactor configuration meeting most closely with the process requirements? The problem of reactor selection is analyzed at three strategy levels_ Decisions are made at each strategy level using the reactor "wish" list. Combination of the individual decisions yields the final, ideal, reactor configuration. The three strategy levels are: Strategy leoel I: "Catalyst" design strategy. At this strategy level the ideal catalyst particle size, shape, porous structure and distribution of active material are determined. For gas-liquid systems, the appropriate decision concerns the choice of gas-dispersed or liquid-dispersed systems, and the provision of the appropriate ratio between liquid-phase bulk volume and volume of liquid-phase diffusion layer. Strategy level I1: Injection and dispersion strategies. (a) Reactant and energy injection strategy: injection strategies examined include one-shot (batch), continuous, pulsed injection, reversed flow operation, and staged injection (in time or space), and the use of dispersionless contacting by keeping the reactants separated by a barrier (membrane). (b) Choice of the optimum state of mixedness for concentration and temperature: the proper choice of state of mixedness can influence selectivity and product properties. (c) Separation of product or energy in situ: product removal in situ helps to increase conversion by driving the reaction to the right and preventing undesirable side reactions_ Removal of energy in situ by use of evaporating solvents has the function of a thermal flywheel. (d) Contacting flow pattern: here there is a choice between co-, counter- and cross-current contacting of phases. Strategy leoel Ill. Choice of hydrodynamic flow regime. Here the choice between the various "fluidization" regimes, e.g. dispersed bubbly flow, slug flow, churn-turbulent flow, dense-phase transport, dilute-phase transport, is made on the basis of the interphase mass transfer characteristics, heat transfer, mixing, etc. Combination of the decisions reached at the three strategy levels will yield the most suitable reactor configuration. In this paper it is argued that a systematic approach to reactor selection may lead to novel and innovative reactor configurations with a potential edge over existing and conventional technologies.

INTRODUCTION F o r carrying out multiphase reactions (gas-solid, gas-liquid, gas-liquid-solid, liquid-liquid, g a s - l i q u i d solid, l i q u i d - l i q u i d - s o l i d , . . . ) , the n u m b e r of reactor configurations that are possible is extremely large. Therefore, there is a need to give careful consideration to the choice of the " i d e a l " reactor configuration t h a t meets fully with all the process " m u s t s " and, to the m a x i m u m possible extent, the process " w a n t s " . The process " m u s t s " could be: operability within technology feasible reaction coordinates of temperature, pressure and residence time, safe o p e r a t i o n and freedom from runaways, e n v i r o n m e n t a l acceptability, and

feasibility of scale-up to economically justified size_ The process " w a n t s " could be the following: m a x i m u m possible conversion of the feedstocks, m a x i m u m selectivity of r e a c t i o n to desired products, easy operability, lowest capital and operating costs, s t e m m i n g from, e.g. - - l o w pressure drop, --effective utilization of reactor space, - - s i m p l e constructions, etc. This paper considers both two- a n d t h r e e - p h a s e reactor selection strategies (see Fig. 1). Typically, in the petroleum and petrochemical industry, even small percentage improvements, say of the order of 0 . 5 ~ with respect to selectivity can be

*To whom correspondence should be addressed.

4029

4030

tnt21
f.maximum selectlvlty~ maximum conversion

R. KRISHNAand S. T. SlE

Economics

d products ired products verted reactants

Process requirements

ease of scale-up high throughput low pressure drop

Environment & Safety

Fig. I. The problem of reactor selection to meet the desired process requirements and constraints. The reactor has to meet with several "musts" and "wants". G extremely significant. For example, improvement of gasoline selectivity in Fluid Catalytic Cracking (FCC) operations by 0.5~ would mean an increased revenue of US$2.5 million per day on a global basis. Improved yield and selectivity can be crucial for process licensers. A 1~ selectivity advantage in the manufacture of ethene oxide (obtained by air oxidation of ethene) could be significant enough for a process licenser to gain a marketing edge over a competitor. For the ethene oxide producer, a 1~ selectivity improvement signifies an increased revenue of the order of US$1 million per year. While the major process improvements will no doubt stem from improved reaction "chemistry" and catalyst design, there is further scope for effecting improved performance by clever choice of the reactor configuration. In the FCC riser reactor, improved feed atomization, better gas-solids contacting at inlet (e.g. by pre-fluidization with steam), and closer approach to plug flow of gas and catalyst phases are known to lead to great economic benefits. For ethene oxide manufacture, if it is possible to develop a packed bed reactor that operates under substantially isothermal conditions, the result will be significant selectivity advantages due to suppression of the undesirable combustion reaction. In the Shell Middle Distillates Synthesis (SMDS) process which converts natural gas to synthetic hydrocarbons via advanced Fischer-Tropsch synthesis, the synthesis reactor configuration chosen for the first commercial unit in Malaysia, started up in 1993, was the multi-tubular downflow trickle bed with catalyst inside the tubes (Sic et al., 1991); see Fig. 2(a)_ Because of the enormous exothermicity of the synthesis reaction and the relatively poor heat transfer, a very large heat transfer area is required. The reactor volume and weight are largely governed by the installable heat transfer area in a vessel of given volume. Use of the bubble column slurry reactor [see Fig. 2(b)'l provides much better heat transfer

GUIIL
(a) (b)
Fig. 2. Reactor types used for Fischer-Tropsch synthesis of hydrocarbons. (a) Multitubular trickle bed reactor, (b) bubble column slurry reactor.

characteristics (an improvement of a factor of about five over fixed bed units) and could lead to considerably lower reactor volumes. However, the anticipated scale-up problems with bubble column slurry operation were of overriding concern for Shell, who decided to adopt the fixed bed technology mainly due to a quicker development scenario which allowed them to meet the time plans or'the business. The lead time for development of processes in the petroleum and petrochemical industries is usually of the order of a decade and for first-of-a-kind-technology such as the SMDS process there is an incentive to adopt a sure, safe and quicker process development route. It is interesting to note that other companies, e.g. Sasol and Exxon, have more recently opted for the slurry reactor configuration. These companies apparently did not consider the long lead time for development of the bubble column slurry reactor to be an insurmountable problem. The above discussions serve to underline the importance not only of choosing the reactor with the promise of the best performance but also for the need to anticipate scale-up difficulties. The approach advocated in this article is to attack the problem of reactor selection in a systematic structured manner.

Strategies for multiphase reactor selection

4031

Level ~~o~ l
~ " " ~ ~ .~;i~i~

"catalyst" design ~ Reactan| andEnergy Strategies


Injection

Level
[

.................................~, ..........= ........~ ' i m ~ n e

...........

uuu ....................................................................... ~......................................................................... i i................................................................................... !(b) ~ , ~ / i(c) ,'~'\ i(d) ~ n 1-hlxn

II01 " ~

Level

Hydrodynamic Flow Regimes

l [~ :i :~ ~"~1 ~
t!,:~

i A + B ~ C) i ~i

~'@ U~

~ ~1 ~"i~ Iiiill

Fig. 3. Three-level strategy for multiphase reactor selection. The reactor selection problem is analyzed at three strategy levels as pictured in Fig. 3, and at each of these levels "decisions" are made on reactor sub-sets or attributes. At Strategy level I, the optimal "catalyst" design is developed. This involves determining the particle size, shape, porous texture, etc. For a gas-liquid system this would involve choosing an appropriate value of the parameter fl -- (bulk liquid volume)/(volume of liquid diffusion layer). At Strategy level II, a decision is made on the ideal manner in which reactants and energy are injected into the reactor, the ideal state of mixedness with respect to concentration and temperature, whether to remove products in situ, and the desirability of employing dispersionless contacting by separating the phases by a barrier. The choice of the hydrodynamic flow regime is considered at Strategy level III. A combination of the choices made at each strategy level will lead to the final reactor configuration. Compromises need to be made if incompatible sub-sets are chosen. It will be demonstrated here, with the help of several examples, that the structured, systematic approach may reveal several novel reactor configurations and operating strategies. Each of the three strategy levels will now be considered in detail. optimal way. The interplay between chemical kinetics and rates of diffusion of reactant or product molecules can be infuenced by variation of catalyst size, shape and porous texture. The distribution of active material inside the catalyst particles can be an additional variable in designing optimal catalysts, m particular when diffusion-limited poisoning reactions occur.

Effect of particle size. Intraparticle diffusion limitation occurring with relatively fast reactions and slowly diffusing reactant molecules may result in incomplete utilization of the catalyst as the interior is not reached by the reactant molecules. This effect has been the subject of many chemical engineering studies and is relatively well understood. Utilization is expressed as the effectiveness factor q, which for a given particle geometry is a function of the generalized Thiele modulus:

(~gen ~

S A ~[ D~ff

Gas-solid (catalyst) systems

STRATEGY LEVEL

t--"CATALYST" DESIGN

Whilst the choice of"fluidization" regime (Strategy level III) sets limits for suitable sizes and shapes of catalyst particles via demands on fluidizability, suspensibility, pressure drop, flowability, etc., at this first strategy level the possibility of varying the catalyst morphology will be considered, so as to meet process demands dictated by reaction and diffusion in an

in which (V/SA) is the volume/external surface area ratio of the catalyst. For a given conversion rate, the external surface determines the flux density for diffusion of reactants to the catalytic surface inside the catalyst volume. The ratio (VISA), which has the dimension of length, can be considered to be a measure of the average diffusion distance necessary to penetrate the particle. The relation between r/and 4~o, is shown in Fig. 4_ As a general rule, the particle size choice should aim at high effectiveness, so as to make good use of catalyst materials and reactor volume, and for improved activity. There are exceptions to this general rule, however, as will be discussed later. Pore diffusion limitation is not uncommon in industrial fixed bed catalytic processes, particularly in classical processes which use relatively large catalyst

4032

R. KRISHNAand S. T. StE surface diffusion, and hence depends on the gas pressure and the pore texture. In catalytic processes where liquid is present, the catalyst pores are likely to be filled with liquid and the low diffusivity in the liquid phase may increase the likelihood of diffusion limitation. Figure 6(a) shows the effect of particle size on the performance of a C o / M o / a l u m i n a catalyst for hydrodesulfurization of a gas oil in trickle flow, demonstrating the occurrence of diffusion limitation. Figure 6(b) s h o w s effectiveness factors of a series of catalysts of different particle sizes having the same composition. It can be seen that utilization of the catalyst is reasonably complete only with particles having a diameter of less than 1 mm. Catalyst utilization is far from complete for catalysts of relatively large size (pellets of 5-15 mm diameter and length) as used in classical hydrogenation processes before, and shortly after, World War II. Currently, extrudates of about 1.5 mm size are used for hydroprocessing, achieving effectiveness factor values close to unity. Another example of the occurrence of diffusion limitation caused by the presence of liquid in the catalyst pores is depicted in Fig. 7. This figure shows the effectiveness factor as a function of the Thiele modulus for a number of Fischer-Tropsch catalysts operated under different conditions. The experimental results are in very good agreement with the theoretical curve calculated for the case that the reaction rate obeys first-order kinetics in hydrogen (as has been verified experimentally), while the limiting factor is the rate of diffusion of hydrogen as reactant in the liquid hydrocarbon product which fills the pores. The excellent fit demonstrates that as far as catalyst effectivity is concerned, its dependence on catalyst morphology may be established on the basis of first

Effectiveness0. factor, rt

0.0

.)0 Generalized Thiele Modulus =,,,~

Fig. 4. Effectiveness factor as a function of the generalized Thiele modulus for a first-order reaction. particles (pellets of 5 - 1 5 m m diameter). Simple calculations comparing the time needed for diffusion over a given distance inside the particle with the maximum t=me available for the reaction, show that for practical reaction rates as apply to most processes, and typical diffusivities in gas- or liquid-filled pores of catalysts, diffusion limitation will generally occur with particles having a diameter of a few millimeters_ In the case of very fast reactions or very slow diffusion, as in the case of so-called conformational-type diffusion inside zeolite pores, pore diffusion limitation will occur with particles of even much smaller dimensions. Figure 5 shows effectiveness factors of some catalysts in gas-phase reactions as a function of catalyst size. It can be seen that in these processes catalyst utilization is far from complete for fixed bed catalysts of practical sizes, i_e. above 3 mm_ In such gas-phase reactions the diffusivity inside the catalyst particles is usually a Knudsen-type diffusion or a

ammon!a
0.8

0.4

\ ~ Dehydrogenation Cumene'~ '~.~cyclohexane cracking D""B.~__ffi


I I I I 4 I I I 8 I I I I.

(~ 0

12

dp/ [mm]
Fig. 5. Effectiveness (actors in some gas-phase fixed bed processes. Data from: Corrigan and Garver (1953), Weisz and Schwegler (1955) and Bokhoven and Van Raayen (1954).

Strategies for multiphase reactor selection

4033

100

% Desulphurization

_ m
5x5mm " ~
5(
I I |

2 3 4 E :)ace Velocity/[l~ kg rrrz h"]

dpl [mm]
(b)

(a)

Fig. 6.(a) Effect of catalyst particle size in trickle-flow hydrodesulfurization of Kuwait straight run gas oil over a Co/Mo/alumina catalyst. (b) Utilization of catalyst as a function of particle size in hydrodesulfurization of Kuwait straight run gas oil over a Co/Mo/alumina catalyst. Adapted from Le Nobel and Choufour (1959).

Effectivenes,, factor, q

particle diameter = 0.38 - 2.6 mm pore diameter = 10 - 65 nm

0.1 ~

Diversecatalysts Thiele Modulus, 10

0.1

Fig. 7. Effectiveness factors as a function of the Thiele modulus for Co-type Fischer-Tropsch Catalysts_ Adapted from Post et al. (1989).

principles if the reaction kinetics and the mechanism of diffusion are known.

Geometry of catalyst particles. The geometric factor which governs the utilization of catalyst particles from a given catalytic material is in first approximation the ratio between external surface area and volume of the particle, (SAllY, cf. expression of the generalized Thiele modulus). For geometrically similar shapes the (SA/V) is inversely proportional to a characteristic dimension, e.g_ the diameter dp. Hence, for such similar shapes the (SA/V) multiplied with the diameter dp is a dimensionless constant which is characteristic for the shape. Figure 8 compares the (SA/V) of some catalyst shapes currently used in fixed bed reactors in a plot of the above dimensionless constant against the length over diameter ratio (L/dp). It can be seen that rings, hollow extrudates and more sophisticated shapes such as "wagon wheels" and "miniliths" offer a greater surface-to-volume ratio than the more c o m m o n cylindrical tablets and cylindrical extrudates of the same outside diameter and length. Since pressure drop and effective radial heat conductivity in packings are mainly determined by the outside

diameter of the particles, these shapes are advantageously used in gas-phase operated adiabatic fixed bed reactors or non-adiabatic tubular reactors. For liquid-phase processes they are less advantageous since the hollow space is likely to be filled with relatively stagnant liquid. Extrudates with a clover leaf cross section, namely trilobe or quadrulobe extrudates, offer a greater surface-to-volume ratio than cylindrical extrudates of the same maximal outside diameter, and retain their advantage also in liquid-phase operation. It can be seen that round, diskshaped tablets of a length-to-diameter ratio below 1, have a greater surface area than cylindrical extrudates of the same diameter with a ratio above 1. Notwithstanding this, extrudates have largely superseded tablets in most fixed bed processes. This is because reduction of the diameter of pressed tablets to values much lower than 5-15 mm, as used in classical fixed bed processes, results in rapidly escalating manufacturing costs (the number of particles for a given weight of catalyst increases with the third power of the linear dimension). Extrudates, however, can be produced at little extra cost at diameters below 1 mm.

4034

R. KRISHNAand S. T. SIE

50

-~-.x dp 10 '

hollowextrul rings

tr"ob ql
~1~ spheres
2 0 I I L / dp

v
quadrulobes cylindrical extrudates

~~'~"~-~i~ :liscs --J 10

Fig. 8. Surface area over volume ratio of shaped catalyst particles.

Pressure drop, reactor productivity and strength of catalyst particles. As discussed above, the problem of
diffusion limitation can be alleviated by choosing smaller catalyst particles of a shape that has a high (SA/F) ratio_ The main limiting factor for reduction of particle size in a fixed bed reactor is generally the pressure drop over the bed. Figure 9 shows the general

maximum
Pressur Am edro/ p,
i:~Opsab.~

/ /
r~Rtm

!~?.m

~ reaction rate Overallreaction rate per kgcalalyst


Fig. 9. Relation between reactor productivity and pressure drop for some types of fixed bed catalysts.

acceptable

relation between pressure drop and reactor productivity for extrudates of different shape and size_ This figure shows trends rather than quantitative data, since the latter depends on several factors such as properties of reactant fluids, aspect ratio of the bed, axial versus radial flow, etc. The demand for a minimum desired reaction rate within a maximum allowable pressure drop determines suitable sizes and shapes of catalyst particles for a given case. For instance, for the case depicted, it can be seen that cylindrical extrudates need not be considered, but that wagon wheels in a certain range of sizes can be used. Trilobe extrudates of a specific size are marginally applicable for this case. The superiority of wagon wheel and trilobe extrudates over cylindrical extrudates with respect to the reaction rate-pressure drop relation is caused by the higher surface area to volume ratio of the former types of extrudates. However, these types of extrudates also have their drawbacks so that they may not be the preferred type under all circumstances_ Figure 10 shows the relationship between manufacturing costs and surface-to-volume ratio for the

~ manufacturing ........................
/

/ i/
/ .:

............... ~

maximum allowable
cost

SA IV
Fig. tO. Relation between reactor manufacturing cost and surface/volume ratio of some fixed bed catalysts.

Strategies for multiphase reactor selection

4O35

amply fulfilled by trilobes and cylinders, this is only marginally so in the case of wagon wheels_

Strength

. . . . . . . . . . . . . . . .

SA/Vratlo
SA/V

~ ~

Fig. I1. Strength vcrsus surface-to-volume ratio of some fixed bed catalysts. different types of extrudates. It can be seen that, in general, the simple cylindrical extrudates are the cheapest to manufacture. The figure shows that to satisfy the specifications on minimum ( S A / V ) ratio and maximum allowable costs, as indicated in this figure for a hypothetical case, both trilobe and cylindrical extrudates may be considered. Of these two, cylindrical extrudates present the more economic solution, while trilobe extrudates are merely acceptable. Strength of catalyst particles is in most cases an important factor in fixed bed processes, Particles should be able to withstand the forces exerted on them by the weight of the bed above, and by the pressure drop. This strength is generally measured either on a single catalyst particle, giving the so-called sidecrushing strength (SCS), or measured by exerting a pressure on a small bed of particles, giving the so-called bulk-crushing strength (BCS). Figure 11 shows the relationship between catalyst strength (either SCS or BCS) and ( S A / V ) ratio for extrudates of some different shapes and different sizes. It can be seen that extrudates of the more sophisticated shapes are, in general, less strong. Minimum requirements on strength and ( S A / V ) ratio as indicated in this figure for a hypothetical case, can be satisfied by all three types of extrudates. Whereas the requirements are

Effect of diffusion limitation on selectivity. Diffusion limitation not only affects the apparent activity of a catalyst, but can also affect selectivity. The best-known effect of diffusion limitation is a negative one. This applies to a case of consecutive reactions Am ~ A 2 ---, A3, where A 2 is the desired product. Diffusion limitation reduces the chances of the intermediate product A 2 escaping from the catalyst particle so that the selectivity of its production will be lowered due to increased chances of A 2 being further converted to unwanted product A 3. A good example of the influence of intraparticle diffusion on selectivity is the partial hydrogenation of edible oils to modify their melting behavior and their taste stability. Edible oils consist of esters of mainly unsaturated acids and glycerol. Hydrogenation of oils is the reaction of hydrogen with unsaturated triacylglycerol molecules over a nickel-based catalyst. For example, the hydrogenation of trioleoylglycerol (O13) to tristearoylglycerol (St3) occurs according to the following consecutive steps: O13 ---, StOl 2 St2OI---, St 3 where StOI 2 and St2Ol represent the partially hydrogenated products. With respect to product properties like melting behavior, selective hydrogenation giving the maximum concentration of intermediates StOl 2 and St2OI is required. The selectivity can be influenced by proper choice of the pore size and particle size of the catalyst; see Fig. 12 (Colen et al,, 1988). In contrast to the above case, diffusion limitation can also have positive consequences. An example of the use of diffusion limitation as a means of steering the formation of products towards a desired molecule is the selective production of p-xylene by disproportionation of toluene. The principle is shown schematically in Fig. 13. The selective formation of p-xylene from toluene depends on the differences in diffusion rates between the isomeric xylenes in ZSM-5 zeolite, p-Xylene diffuses much faster than the two other xylenes, while the feed molecule toluene and the

I OI3 --) StOI2 ~> St2OI --) St3 I

increasing catalyst melting point of hydrogenated product decreasing particle size

30 F
% St3
lo I0.1
I

10

10

Thiele modulus

pore radius/[nm]

Fig. 12. Influence of intraparticle diffusion on selectivity of intermediate product in a consecutive reaction sequence. The reaction considered is the hydrogenation of trioleoylglycerol (O13) to tristearoylglycerol (Sta), which proceeds through the partially hydrogenated intermediates StO,, and St2OI. By tailoring the catalyst with respect to particle size and pore size the diffusional influence can be altered to obtain a product with the desired property, such as melting point. Adapted from Colen et al. (1988).

4036

R. KRISHNAand S. T. Sm

Dtelaliva

Drelative

I--

10000

I
I
I
Fig. 13. Schematic representation of the role of diffusion in production of p-xylene by disproportionation of toluene. Adapted from Weisz (1980). by-product benzene diffuse even faster. Without diffusional control, the relative proportions of the xylene isomers will correspond to the thermodynamic equilibrium between them. As the diffusion of the xylene molecules becomes more limiting, the differences in diffusivity of the isomeric xylenes gain importance, resulting in increased selectivity to the para-isomer.
60

Vmax/ [% wt on fresh catalyst]

Effect of diffusion limitation on a catalyst deactivation. If the catalyst deactivates as a consequence of the occurrence of a reaction which is subject to diffusion limitation, regulated access of the catalyst particle to the poison-generating molecules by tailoring catalyst morphology offers a means to control catalyst life. This principle forms the basis for catalyst design for hydroconversion of metal-containing residual oils (Sie, 1980). In this process, the deposition of contaminant metals (Ni, V) as sulfides on the catalyst surface deactivates it for the main reactions, the hydrodesulfurization and hydrocracking of the oil molecules_ The metals originate from large molecular complexes (asphaltenes) which because of their size diffuse very slowly into the catalyst particles. Since the breakdown of these molecules which releases the metals is pore diffusion limited, deposition of the metal sulfides occurs in an outer zone of the particle, leaving the inner core unpoisoned. The penetration depth of metal-containing asphaltenic species into the catalyst particle determines the size of the unpoisoned inner core of the catalyst particle, and thereby its residual activity for the desired reactions, hydrodesulfurization and hydrocracking of the non-asphaltenic oil molecules_ Equally important is the effect of penetration depth on catalyst life: in the course of time, metal sulfides continue to be deposited in the peripheral zone and eventually fill the pore volume in this area completely, which means the end of catalyst life. As the zone is broader and involves a larger proportion of the particle, it takes more time to completely deactivate the catalyst. Figure 14 shows that for different catalysts used in hydrodesulfurization of a given residual oil feed, the

0! 0

I 0.5

F xPV I [mUg]

Fig. 14. Correlation between vanadium uptake capacity with effective pore volume in residue hydroprocessing. F is the effectiveness factor for storage of metals deposits in the pore volume of the catalyst particle. Adapted from Sic (1993).

maximum metal uptake capacity is proportional to the effectively used pore volume for metal deposition (Sie, 1993). By variation of the catalyst morphology, i.e. changing catalyst size and shape or the porous texture, one can precisely regulate the relative depth of metal penetration and thereby control the life of the catalyst and its activity for the main reactions. This is illustrated in Fig. 15 which shows the effect of average pore diameter on the performance of a series of similarly shaped catalysts for hydrodesulfurization of a Middle East atmospheric residue. Catalyst activity in this plot has been defined as the second-order rate constant for sulfur removal at half of the useful catalyst life. It can be seen that activity decreases as accessibility for asphaltenes increases with larger pore diameters. This is because the active inner core of the catalyst becomes smaller. Catalyst life, which is also shown in this plot, becomes longer as accessibility increases, because a larger proportion of the catalyst pore volume becomes available for storage of metal deposits. From Fig. 15, it can be seen that even relatively small changes in average pore diameter (1 nm) can have significant effects. Control of average pore size of the catalyst support (which can be done,

Strategies for multiphase reactor selection

4037

b..9
0

0"-.. -. ..41..............'.,,...

Activity

.......Q
.... . D

~
.................

Useful
life

1 nm I I

pore diameter/[nm]
Fig. 15. Effect of average pore diameter on the performance of catalysts for the hydrodesulfurization of atmospheric residue of a Middle East crude oil. Catalysts studied belong to a series of Co/Mo/alumina catalysts of identical shape and size having sharp unimodal pore size distributions. Adapted from Sie (1993).

I PROCESS SPECIFICATIONS Conversions (% HDS, %HDM, % Cracking) Run Length

FEEDSTOCK PROPERTIES Metals content ~._. , 1~,~__.CATALYST PROPERTIES Asphaltenes content ...-.__~ I CATALYST PERFORMANCE Porous texture ~"~'1 -Activity Particle morphology I ,Stability Chemical composition PROCESS CONDITIONS ,Temperature -Pressure CATALYST CONFIGURATION (types, amounts and sequence) Process Line-Up Operations Strategy
Fig. 16. Relationships between process, feedstock and catalyst parameters determining the choice of hydroprocessing catalysts. Adapted from Oelderik et al. (1989). for instance, by hydrothermal treatment) thus offers a means of adjusting catalyst performance for a given duty. In a fixed bed process for hydroprocessing of residues, the removal of metals from the oil stream will cause a decrease in their concentration in the direction of the stream. Therefore, the balance between activity and tolerance for metals deposition will not be the same in different parts of the reactor or reactor train. Thus, it is advantageous to use a combination of different catalysts rather than just one: a highly metals-tolerant catalyst upstream (which can cope with the relatively high concentration of metal contaminants in the feed) and a highly active catalyst in the downstream part. The latter catalyst will have a low tolerance for metals deposition, but this tolerance suffices since metals concentration in the oil has already been lowered. Ideally, a large number of catalysts designed for a gradually sliding balance between activity and metals tolerance should be placed in series_ With such a multicatalyst combination, definition of an optimal catalyst is not straightforward since the desired activity and life of a catalyst depend on its position in the reactor train, on feedstock characteristics, on target conversion level and desired onstream time, etc., as is shown in Fig. 16. However, when a process model is available which has been validated with sufficient practical data, it becomes possible to predefine catalyst parameters for a desired duty_ An example of catalyst design assisted by a computerized process model is shown in Fig. 17. It

4038

R. KRISHNAand S. T. SIE

8000

Predicted catalyst life/[h] Uniform


0 ' '

Outer Egg Shell

t --I

-3 Base 3 Average pore diameter/[nm] Fig. 17. "Computer-aided design" of catalyst porous texture. Predicted life as a function of the average pore diameter of a tail-end catalyst in the hydroconversion of an Arabian Heavy vacuum residue at 60'~,~conversion. Adapted from Oelderik et al. (1989). can be seen that catalyst life passes through a maximum at a certain value of the average pore diameter. At smaller diameters, the catalyst deactivates faster because the relatively shallow outer layer available for deposition of metal sulfides becomes more rapidly filled. At larger diameters, the activity of the catalyst is relatively low (as a consequence of the relatively small unpoisoned inner core) so that higher temperatures have to be applied to reach the process specifications. The smaller span of operating temperatures thus shortens the operating cycle. It can be seen that, in this example, the experimentally tested base case is not far from the calculated optimum situation, but this is of course not always true. This possibility of designing catalysts is especially important in residue hydroprocessing since actual life tests require runs of several thousands of hours, so that catalyst optimization by way of empirical testing is not likely to achieve an optimum result within a reasonable amount of time and effort.

Middle Egg White

Inner Egg Yolk

Fig. 18. Different activity distributions in a catalyst pellet.

accompanied by an irreversible poisoning reaction (Becker and Wei, 1977). In the case of a diffusionlimited poisoning reaction, where the poisoning occurs in an outer shell (pore-mouth poisoning) it can be advantageous to concentrate the active component in a layer inside the pellet. When the bare support is capable of adsorbing the poison, the active components inside are thus protected from it. The main factors determining the best type of activity distribution are the Thiele moduli for the main reaction and the poisoning reaction. Figure 19 presents a catalyst selection chart on this basis.

Spatial distribution of activity within catalyst pellets.


In the absence of diffusion limitation, it is a c o m m o n policy to uniformly load the support pellet with catalytically active species, in order to maximize the catalytic activity per unit volume of catalyst bed. If the effectiveness factor is significantly lower than unity, it is obvious that the catalytic agent may be saved by leaving the inner core free from it, without losing overall activity. However, non-uniform distributions of activity within a catalyst pellet can have additional advantages besides savings in active material. Different possible activity distributions are shown in Fig. 18. The problem of optimal activity distributions has been studied theoretically by several authors for various cases of single, parallel and consecutive reactions, in the presence or absence of extraparticle mass transfer limitation and for nonisothermal cases; see, e.g. the recent review of Gavriilidis and Varma (1993). Non-uniform activity distributions can have particular advantages when the main reaction is

Intraparticle heat effects. Intraparticle heat transport can also have an effect on selectivity when the selectivity of the desired reaction is strongly dependent upon temperature and when significant amounts of heat are generated by the reaction itself or by accompanying reactions. A temperature profile over the catalyst particle with a peak at the interior, as a result of the difficulty of removing locally generated heat, may give rise to deterioration of selectivity_ In this case, one could opt for shortening the path of intraparticle heat conduction by choosing small particles or shapes such as rings, miniliths, etc., depending opon the allowable pressure drop. Another solution is to locate the catalytic active material at the outside of the particles, i.e. apply an egg-shell-type catalyst. The latter approach is illustrated in Fig. 20 for the oxidation of ethene to ethene oxide: a strongly exothermic reaction which is accompanied by an even more exothermic unwanted reaction, namely total combustion to carbon dioxide and water. The latter reaction, which lowers the selectivity to desired ethene oxide, has a higher activation energy than the desired one. Therefore, intraparticle temperature gradients are detrimental to selectivity_ Mechanical transport properties of catalyst particles.
Aside from intraparticle mass and heat transport effects as discussed above, the mechanical transport

Strategies for multiphase reactor selection 10


diffusion-limited main reaction with uniform poisoning diffusion-limited main reaction with pore mouth poisoning

4039

o
Egg
Shell , = Rk~km~ 1
"

o
~gg
/ ~ , ~
White

/
"

Beckerand Wei' 1977 I Egg Yolk diffusion-free m~np~era:~nu.th ~is~Jo;ing 10 40

Uniform/ ,/~ /~l iren:c~een 0.1 ~ ~ s o n i n g . ' ' ' , 0.1 1

Fig. 19. Catalyst selection chart for four catalyst designs for a catalytic reaction accompanied by a poisoning reaction. Criterion: longest catalytic life with effectiveness of 0.4 or higher. Adapted from Becker and Wei (1977).

80

activelayerlocated 0.00 mmfromsurface '-....,..


"--., ,..,.. ~===.. '-....,

% selectivity
;', ~ .28 0.64
b I

%conversion

60

Fig. 20. Selectivity to ethene oxide in partial oxidation of ethene as a function of the location of active layer. Adapted from Gavriilidis and Varma (1992). properties desired for a chosen reactor type are factors in catalyst particle design. In fluidized bed processes the desire to operate at reasonably low gas velocities usually restricts the diameters to below a few tenths of a millimeter. For good fluidization behavior in a bubble fluidized bed, catalyst particles in the form of microspheres with a certain size distribution, e.g. in the range of 30-150/am, are preferred. Microspheroidal catalyst particles in this size range can be prepared by spray drying of a hydrogel or slurry containing the catalytic material. Fluidized bed reactors are very often applied when there is a need to cycle catalyst particles between the reactor and a regenerator (in case of fast catalyst deactivation of the catalyst during the reaction) or a heater (in case heat has to be supplied). Rapid transport of catalyst between vessels implies subjecting the particles to strong attrition forces and possible thermal shocks; the attrition resistance is therefore an important criterion. The microspheroidal form, as already mentioned, is a favorable shape to reduce mechanical attrition.

However, when the catalytic material is inherently weak, special means are required to bring attrition resistance to the required level for a practical process. A way in which this can be achieved is illustrated by the development of the catalyst for the recently commercialized Du Pont fluid bed butane oxidation process to produce maleic anhydride. The vanadium phosphate (VPO) used as catalyst is too weak to withstand the forces associated with the cycling between the fluidized bed regenerator and the dense phase riser reactor where gas velocities up to 1 m s prevail_ Attrition resistance is imparted to the VPO catalyst by spray drying this material together with a silica hydrogel under conditions which allow silica to migrate to the outer regions, thus encapsulating the active VPO catalyst into a porous silica shell whose pore openings allow reactant and product molecules to diffuse into and out of the particle; see Fig. 21. In slurry processes the particle size is often determined by the desired catalyst concentration to obtain acceptable reaction rates on the one hand, and maximum allowable slurry viscosity on the other. At the same solids concentration, finer particles give rise to higher slurry viscosities. Other factors which have an impact on the choice of particle size are the settling velocity (one may either want to avoid settling at the velocities applied or to have rapid settling for slurry thickening), filterability of the slurry, or separation possibility by hydrocyclones, sieve bends, or other separating equipment. Separation of solid catalyst from liquid by the separation methods mentioned above is difficult or impossible for very fine catalyst particles. Such catalyst particles, e.g. smaller than 1/am, may be desirable for reasons of catalyst effectiveness with fast reactions. Another advantage of such fine particles is that due to the fast reaction, and their presence in

4040 colloidal

R. KRISHNAand S. T. SIE

50% Catalyst Particles silica particl@ @

10% PSA

microspheres 45-150 p.m Conventional Attrition resistant catalvst VPO catalvst with silica shell
Porous Fig. 21. Attrition resistant VPO catalyst with silica shell as used in the Du Pont butane oxidation process.
After Contractor et al. (1987).

gas Electromagneticproduct separator


'i
QQO QO D~ Do

catalyst particles smaller than 1 p.m can be used reasons of effectiveness

for

t=L
gas ~ i q u i d

magnetised catalyst particles

fine particles in diffusion film may enhance mass transfer

Fig. 22_ Conceptual reactor system using extremely fine particles for fast reactions.

or near the diffusion film, they may significantly enhance the gas-liquid mass transfer. A possible way to use ultrafine particles in a slurry bubble column and yet avoid the solid-liquid separation problem is to make the particles magnetizable, so that they can be removed from the liquid with an electromagnetic separator; see Fig. 22_ Another conceptual way of improving the separation possibilities between catalyst and liquid is to encapsulate the catalyst by a porous shell rather similar to the butane oxidation catalyst discussed above; see Fig. 23. This technique may even be applicable to homogeneous catalysts: if the metal organic complex synthesized from components small enough to pass through the pores of the shell is too bulky to diffuse out of the shell, a situation is obtained akin to the well-known ship-in-a-bottle. In this concept, there is not only a diffusion barrier for the active catalyst, but there may also be a diffusion limitation. In the case of a hydroformylation reaction, which is of a negative order in CO, such a diffusion barrier may be beneficial. In moving bed processes, smooth movement of the catalyst poses demands on flowability and additional requirements have to be met on attrition resistance

and crush strength. Uniformity of size may also be important for moving bed catalysts. Since the particles are free to move, undesired segregation may occur when particles appreciably different in size are present (note that the mass of a particle is proportional to the third power of its diameter). As a result of movement, smaller particles may tend to occupy the space left by the larger ones, thus decreasing the permeability of the packing.
Structured packings and monoliths. In the catalytic treatment of flue gases to reduce atmospheric pollutions, e.g. of sulfur and nitrous oxides, very large volumes of gas at relatively low (near atmospheric) pressure must be handled, which calls for reactor designs with very low pressure drop. In addition, the chance of plugging by particulates (soot or ash) in the gas should be minimal. Figure 24 shows various catalyst configurations meeting with these requirements. In the parallel-passage reactor, particles of a catalyst or a regenerable adsorbent are enclosed in wire screen envelopes which are mounted in a parallel fashion. Gas flows in the empty passages between

Strategies for multiphase reactor selection

4041

olefin, CO. H2

Hydroformylation reaction (negative first order in CO)

an be enhanced by crate introduction of diffusion limitation

reactants to make RI complex ac from outsid homogene, catalyst syrlu lu~,~uu 11 ,-~,tu

rous ca Ilow ._.rosphere r roduct

Fig. 23. Concept of an encapsulated (homogeneous) catalyst. The homogeneous catalyst is encapsulated inside a porous hollow microsphere. The complex is synthesized in situ and cannot diffuse out due to the small pore size, say or the order of 1 nm. These spheres are dispersed inside the reactor. Deliberate introduction of diffusional limitation for CO to the ligand has potential beneficial effects.

monolith wash coat parallel passage reactor 1 mm I D

lateral flow reactor

beads-on-astring reactor

Fig. 24. Catalyst configurations for low-pressure drop applications. adjacent envelopes. The straight, unobstructed gas channels give rise to a low-pressure drop and good tolerance to plugging by dust. Pollutant molecules are transported to the catalyst or adsorbent by lateral diffusion. This type of reactor has been applied in the Shell Flue Gas Desulphurization (SFGD) process (Dautzenberg et al., 1971) which is based on the use of a regenerable copper adsorbent. A commercial SFGD unit for cleaning flue gas from refinery furnaces has been built and operated in Japan. A more recent application of the parallel-passage reactor is the selective catalytic reduction of NO x with ammonia to N2 with a vanadium-on-silica-typecatalyst (Goudriaan et al_, 1989). In applications where flue gases are relatively free of particulates, such as flue gases from gas-heated furnaces, dust tolerance requirements are less and the lateral-flow reactor is an option_ In this reactor, which is in essence a fixed bed reactor with a very large cross section over volume ratio, gas flows through the catalyst layers rather than alongside. This type of reactor has been applied also for selective catalytic reduction of NO~ (Samson et al., 1990). Monoliths are also applied in selective catalytic reduction of NO~ because of their low pressure drop and tolerance to plugging by dust, due to the presence of straight parallel channels through which the gas flows in the laminar regime. The inner wails of the channels are provided with a layer of catalytically active material (wash coat). Because the diameter of the channels is rather small, mass transfer by radial diffusion in the gas to the gas-catalyst interface can be sufficiently fast. These monoliths form the elements

4042

R. KRISHNAand S. T. SXE The hydrodynamic and mass transfer characteristics for any system are reflected by the parameter ,8 which is the ratio of the liquid-phase volume to the volume of the diffusion layer. The first major decision for a gas-liquid system is the choice for this parameter ,8; this choice is analogous to the particle size decision for the catalyst. The value of/3 lies in the range 10-40 for liquid sprays and thin liquid films, whereas /~ = 103-104 for gas bubbles in liquid. The choice with regard to ,8 depends on the relative rates of chemical reaction and mass transfer within the liquid phase, portrayed by the Hatta number. The choices for/3 are summarized in the Enhancement factor-Hatta number diagram of Fig. 26, which is equivalent to the

Effectiveness factor-Thiele modulus diagram of Fig. 4. The overall aim is to choose the value of ,8 such that the reactor volume is effectively utilized. Thus, for slow liquid-phase r~actions the aim should Gas-liquid systems For a gas-liquid system with reaction within the be to increase the bulk liquid volume at the expense liquid phase there are fundamentally three different of interfacial area. A high value of ,8 is achieved by modes of gas-liquid contact: (1) gas bubbles dispersed dispersing the gas in the form of bubbles (e.g. bubble in liquid (as encountered in bubble columns), (2) liquid columns and tray columns operating in the froth droplets dispersed in gas (e.g. tray operating in the regime). To give an example, air oxidation of cyclospray regime), and (3) a thin flowing liquid film in hexane (in the liquid phase) is a slow reaction usually contact with a gas (e.g. gas-liquid contacting in a carried out in bubble contactors. packed column or wetted-wall column); see Fig. 25. In the fast pseudo-first-order reaction regime, the reaction occurs predominantly in the diffusion film close to the gas-hquid interface and a contactor with Dispersion Modes a low value of ,8 should be chosen (e.g. spray towers 4 and packed columns). Furthermore, in the fast gas bubbles liquid drops gas and pseudo-first-order reaction regime, the rate of transfer dispersed in dispersed in liquid in is independent of the liquid-phase hydrodynamics; liquid gas film contact there is no need to spend energy for increasing turbulence in the liquid phase. An example of process operating in the fast pseudo-first-order reaction regime is absorption of carbon dioxide in aqueous caustic solutions; this is usually carried out in packed columns_ The liquid phase flows down the column in thin liquid rivulets. If the gas-liquid reaction correfl = 103-104 /3 = 1 0 - 40 sponds to the instantaneous reaction rate regime, once Fig. 25. Three fundamental procedures for contacting gases again efforts should be made to maximize the and liquids, fl is the ratio of the liquid-phase volume to the interracial area at the expense of bulk liquid volume. volume of the diffusion layer within the liquid phase. In contrast to the fast pseudo-first-order reaction

of a commercial reactor, which consists of a stack of monolith blocks. The features of low pressure drop and dust tolerance are also of importance in the application of monoliths in automotive exhaust catalysis. For this application, monoliths offer the additional advantages of minimal volume for the required catalytic conversion (which gives flexibility in mounting a catalytic exhaust converter in a car) and minimal mass. The latter is important since it minimizes the heating-up time and thus the extra emissions during the cold-start period when the temperature is still too low for the catalyst to be active. The perspectives for use of monoliths in heterogeneous catalysis are analyzed by Cybulski and Moulijn (1994).

1000'

100 Enhancement factor, E 10

slow reaction; choose high value of I~ (e.g. bubble columns)

Fast pseudo first order reaction; choose low I~ (e.g. packed columns, spray towers) instantaneous reaction; choose low [~with high degree of gas phase turbulence 1000

0.1

10 100 Hatta number, Ha

Fig. 26. Enhancement factor for gas-liquid reactions as a function of the Hatta number; adapted from Trambouze et al_ (1988). This diagram is also valid for liquid-liquid contacting.

Strategies for multiphase reactor selection regime, it generally pays to attempt to enhance the degree of turbulence in both the liquid and gas phases. Contactors that satisfy these requirements include tray columns operating in the spray regime and venturi scrubbers. The sulfonation of aromatics using gaseous sulfur trioxide is an instantaneous reaction and is controlled by gas-phase mass transfer. In the commercially used thin-film sulfonator, the liquid reactant flows down a tube as a thin film (low /~), in contact with a highly turbulent gas stream (high kg). A thin-film contactor is chosen in place of a liquid droplet system due to the desire to remove heat from the liquid phase; this heat is generated due to the highly exothermic sulfonation reaction. Chlorination of p-cresol follows a consecutive reaction sequence and the selectivity towards the intermediate monochloro product is influenced by mass transfer limitations within the liquid phase. Increasing the liquidphase mass transfer coefficient, by increasing the bubble size (Teramoto et al., 1970) or the stirrer speed in a stirred vessel (Pangarkar and Sharma, 1974), increases the selectivity towards the intermediate product. Considerations of intrinsic process safety often dictate the choice of fl; it is often desired to minimize the hold-up of one of the hazardous reacting components in the reactor.
STRATEGY LEVEL

4043 II--INJECTION AND

DISPERSION STRATEGIES

Four different sub-levels of this strategy are considered.

ll(a) Reactant and energy injection and phase dispersion strategy


For each reactant to the reactor there are several injection strategies to choose from: (1) one-shot (i.e_ batch), (2) step function (continuous feed), (3) pulsed feed injection, e.g. using square wave functions, and (4) staged injection of the feed, in the sense of time or space_

Gas-liquid-solid systems
Here, both the solid particle size and the ratio fl need to be chosen. The considerations leading to the choices for these parameters are the same as above. If, from a transport-reaction analysis, particle sizes smaller than 1 mm are chosen, slurry operations would need to be considered. On the other hand, if particle sizes larger than say 2 mm are allowable there is extra flexibility in choosing fixed bed (e.g. trickle beds) or three-phase fluidized operations. The choice between fixed beds and three-phase fluid bed operations can he further narrowed down by additional analysis at Strategy levels II and III.

Liquid-liquid systems
Assume that the reaction takes place in one of the phases, say L2. The ideal choice for the parameter ilL2 -~ L2/6L2 a is dictated by the same considerations as for gas-liquid systems; cf. Fig. 26. To achieve high values of ilL2 we should disperse phase L1 in the form of drops in the continuous phase L2. Low values of ilL2 could be achieved by dispersing L2 in the form of drops in the continuous phase L I. Thin liquid film flow, as encountered in gas-liquid systems (cf. Fig. 25), though not impossible, is unusual in liquid-liquid systems. Sometimes, practical considerations override the decisions arrived at from a transport-reaction analysis. It is thus advisable to disperse corrosive liquid so as to reduce contact with the reactor walls. Hazardous liquid mixtures are usually dispersed so as to reduce their hold-up, even if this is contrary to conclusions reached from a transport-reaction analysis.

Batch vs continuous operation. One of the most important decisions the chemical technologist has to make is whether to use continuous or batch processing (Gonzalez and Larder, 1984). The parameters that determine this choice include scale of operation, supply and availability of raw materials, capital and operating costs, maintenance and labor requirements, environmental problems, product quality, operability and safety. Since the continuous processes were developed at the beginning of the century, conventional wisdom has held that batch manufacture is better for high-value products with low outputs (such as fine chemicals), whereas continuous processing is more advantageous for high-volume products of low value (such as petrochemicals). The desire to operate under intrinsically safe conditions requires that the inventory (hold-up) of reactants in the reactor be minimized. Wiederkehr (1988) presents an example of the production of "Carbol", a key product in the synthesis of a vitamin. In view of the reaction conditions (long reaction times, heterogeneous nature of reaction medium, reaction under pressure), a continuous reaction seemed to be feasible only with great difficulty, but Hoffman-La Roche nevertheless decided to go for a continuous reaction, mainly for reasons of safety. Another example is the production of peroxy esters (e.g_ tert-butyl peroxy 2-ethyl hexanoate), based on the reaction between the corresponding acid chloride and the hydroperoxide in the presence of NaOH or KOH, which are highly temperature sensitive and violently unstable. Batch operations to produce even 1000 t per annum are found to be unsafe. A continuous reactor was found to overcome most of the problems and claims have been made for producing purer chemicals at lower capital and running costs (Kohn, 1978). The continuous reactors produced 7-10 times more material per unit volume compared to batch processes, and since the amount of hazardous product present in the unit at any given time is so small, protective barrier walls were unnecessary. Pulsed reactant injection. Continuous processes have been typically operated at steady state. This has been viewed as desirable due to the relative ease of this type of operation. Despite its ease, steady-state operation is not necessarily the optimum mode of

4044

R. KRISHNA and S. T. SIE

operation and varying one or more variables periodically can significantly enhance the process performance (e.g. Douglas, 1967; Horn and Lin, 1967; Ray, 1968; Bailey et al., 1971; Rigopolous et al., 1988; Hatzimanikatis et al., 1993). The simplest mode of periodic operation consists of a pulsed operation since it simply amounts to switching a process variable between two values. When the reaction kinetic is non-linear or when the reaction sequence involves series or parallel reactions proceeding with different reaction orders, the use of pulsed concentrations can lead to conversion or selectivity advantages. One example of the use of the pulsed injection strategy is in the production of diamines by hydrogenation of nitriles, e.g. succinonitrile_ Hydrogenation of nitriles to diamines proceeds via the intermediate imines, which are extremely reactive. Under hydrogen depletion conditions inside the catalyst, these imines will react to produce undesirable by-products. Manna et al. (1992) showed that by using asymmetric oscillations or pulsations at very low frequencies of the inlet concentration of the dinitrile, a significant improvement of yield and selectivity to diamines can be obtained.
Periodic flow reversal. Recently, stemming from the pioneering work of Matros and co-workers (Boreskov and Matros, 1983; Matros, 1985) there has been considerable interest in the cyclic operation of catalytic reactors with periodic flow reversal; see Fig. 27. In fixed bed operation with periodic flow reversal, the fore and aft parts of the catalyst bed act as regenerative heat exchangers for feed and emuent, allowing weakly exothermic reactions to be operated autothermally at high reaction temperatures. The catalytic combustion of undesired components, e.g. volatile organic compounds, in air, are well suited for this way of operation (Eigenberger and Nieken, 1988; Vanin et al., 1993). Reverse flow operation has also been successfully used, on an industrial scale, for oxidation of SO2 (Matros, 1985; Silveston et al., 1994) and, on a pilot scale, for production of synthesis gas from the catalytic partial oxidation of natural gas with air (Blanks et al., 1990). Reversed flow operation for

methanol synthesis has been extensively investigated by Froment and co-workers (Van den Bussche et al., 1993).
Controlled (staged) injection o f reactant. Consider -the co-dimerization of propene and butene to produce heptenes. This reaction is accompanied by two competing, undesirable reactions: dimerization of propene to hexene, and dimerization of butene to octene. The propene dimerization reaction proceeds extremely rapidly and in order to suppress the formation of hexenes, a staged injection of propene into the reactor should be adopted, with all the butenes at the beginning of the operation (Trambouze et al., 1988). In the chlorination of propene to allyl chloride the undesirable side reactions to 1,3-dichloropropane (DCP) and 1,3-dichloropropene (DCP =) can be reduced by employing staged injection of chlorine (see Fig. 28), a strategy which has surprisingly not been suggested in the literature (Biegler and Hughes, 1983; Seider et al., 1990). In the sulfuric acid alkylation of isobutane, staged injection of the olefin feed to the cascade of stirred reactors is employed because of selectivity considerations and improved octane number of the product (Lerner and Citarella, 1991); see Fig. 29. For polymerization of ethene in a simple tubular reactor, ethene conversion of 20-25% is normally achieved. However, on subsequent addition of cold gaseous monomer and initiator, and applying tubes with increasing diameters, conversion of up to 35% is possible (Gerrens, 1982); see Fig. 30. In the hydroprocessing ofoils in trickle bed reactors, staged injection of cold hydrogen gas between the catalyst beds not only compensates for hydrogen consumption but also serves to limit the temperature rise in the reactor. Controlled addition of one of the reactants through a membrane maintains its low concentration and thus limits side reactions or subsequent reactions of the product. Controlled addition can thus also be used to prevent catalyst deactivation by the latter reasons; see Fig. 31(a). Another possibility is the addition of one reactant through a membrane so that a higher

time z

7"

~
7-

Fig. 27. Reverse flow strategy for catalytic combustion of volatile organic compounds in a fixed bed reactor.

Strategies for multiphase reactor selection

4045

propene + CI a DCP

+ CI a .~ allyl chloride

+ CI a .~

DCP =

propene

propene

CI 2

AC, DCP, DCP =


Fig. 28. Staged injection strategy for chlorination of propene to allyl chloride.

refrigerant isobutane

vapors to ~ ~compressor

debutanizer feed

, Gll l

,I,I,) F
recycle acid

Fig. 29. Staged injection ofolefins improves selectivityin the alkylation and Citarella (1991).

ofisobutane. Adapted from Lerner et al.

T/[oC] 300~ 200]:: elhene ~, initiator (a) simple tube reactor

3ooF T/[C] 2ool-

(b) Tube reactor with distributed feed


Fig_ 30. Staged injection of ethene gas and initiator in polymerization. Adapted from Gerrens (1982)_

tubular reactor. A membrane can also be used to prevent a poison from reaching the catalyst site [Fig. 31(c)]. Falconer (1993) list several examples where controlled addition of reactant is beneficial. In the examples considered above, one of the reactants was introduced in a progressive, staged manner. Pursuing this line of attack, the benefits in keeping the two reactants completely segregated from each other and allowing them to meet only on active catalytic sites should be examined. The active components of the catalyst could be incorporated within a non-permselective ceramic membrane with the reactants on either side. Figure 32 shows a schematic diagram of such a catalytic membrane reactor for carrying out the Claus reaction (Sloot 1990)

et al.,

2H2S + SO 2 ~-, asSs + 2H20. This novel reactor type has specific advantages for chemical processes requiring strict adherence of the feed rates to the reaction stoichiometry. The reaction plane within the catalyst membrane would shift in such a manner that the molar fluxes of the reactants across the membrane are always in the stoichiometric ratio; this allows greater flexibility of the reactor to

concentration of the reactant can be obtained on the catalyst surface [Fig. 31(b)]. This may yield higher conversions when the surface concentration of the reactant is limited because of competitive adsorption. In addition, a more uniform concentration can be obtained on the catalyst surface than in a standard

4046

R. KRISHNAand S. T. SIE

Feeq A (a)
Feel B Fee A (b) Fee, B
D

A sorbed on catalyst surface

Feed

(c)

Fig. 31. Dispersionless contacting strategies. (a) Progressive addition of A through a porous membrane wall. (b) Injection of reactants A and B on opposite sides of a catalytic membrane wall with reaction within the catalyst plane. (c) Selectiveexclusion of a poison from the reaction zone by selectivepermeation through a membrane. Adapted from Falconer et al. (1993).

'A"-~'~ 2 H2S + SO~ <->3S 8 + 2H~O "

~ ,,LA A ~

/ ~C C

I I'~1~"1~ A+C ,~, ~

Fig. 32. Catalytic membrane reactor for carrying out the Claus reaction: 2H2S + SO2 ~ 3/8Ss + 2H20. Adapted from Sloot et al. (1990). feed rates of hydrogen sulfide and sulfur dioxide. The practical feasibility of this novel concept has been demonstrated by Sloot et al. (1990). This concept has also been suggested by Van Swaaij and co-workers (see Sloot et al., 1990) for catalytic reduction of nitric oxide with ammonia in flue gas; by keeping the reactants separated and allowing reaction only within the membrane it will be possible to cope with varying ratios of the concentration of nitric oxide and ammonia, while ensuring stoichiometry at the reaction side and avoiding significant slip of reactants in the treated flue gas without the need for accurate dosage of ammonia_ Another class of processes where it is advantageous to keep the reactants separated from each other, except on catalytic sites, is partial oxidation of light hydrocarbons (e.g. methane, ethene, propene, or butene). Flammability considerations usually restrict the feed mixture composition. By adopting the concept of a multitubular cooled catalytic membrane reactor, with the reactants kept separate, it should be possible to avoid any flammability constraint; see the schematic in Fig. 33_ This concept has been demonstrated by Veldsink et aL (1992). Lafarga et al. (1994) used a similar reactor concept for methane oxidative coupling_ Oxygen supply is by permeation through the walls and this helps to maintain the oxygen concentration in the reactor at a low level_ Improved hydrocarbon selectivity is obtained in this membrane reactor compared to a conventional packed bed reactor. The problem of incomplete wetting efficiency of a trickle bed reactor can be solved by the cross flow catalyst contactor concept of Sch66n and co-workers (De Vos et al., 1982) in which the gas and liquid phases are made to flow in separate channels, separated by flat catalyst "plates"; see Fig_ 34. In order to obtain sufficient reaction capacity, the catalyst may contain a great number of parallel plates. In the study by Sch66n, the plates were separated from each other by corrugated planes, forming a system of parallel

Strategies for multiphase reactor selection

4047

Feed C2H4

I keep ethene and oxygen separate and I allow contact only at active site I
catalytic

me

Feed 02

coolant i

out !H40

Fig. 33. Catalytic membrane concept for oxidation of ethene to ethene oxide. Adapted from Veldsink et aL (1992).

Catalyst
heet

LRqaiCdant ~ g e r |
Fig. 34. Cross-current dispersionless contacting of gas (hydrogen) and liquid (oil) in a monolith reactor. Adapted from De Vos et aL (1982). channels. The pressure drop of flow in the channels is negligible and the specific gas-liquid contact area is large.
Staged energy supply and removal. There is a complete analogy between staged reactant injection and staged energy supply, or removal, strategies_ In a batch reactor, for example, the energy removal (supply) strategy can be programmed to keep the temperature constant. The energy removal can be by means of an evaporating solvent or the product itself_ In a continuous flow reactor the temperature profile can be optimized by programmed energy removal by installation of, say, interstage cooling (heating). Commonly, heat is removed (supplied) by indirect heat exchange through installed surfaces within the reactor. Such heat transfer surfaces are conceptually similar to dispersionless contacting in the membrane systems discussed above. If(b) Choice o f the optimum state o f mixedness o f concentration and temperature

The optimum states of mixedness of concentrations and temperature within the reactor have to be decided upon separately. For an isothermal reaction within a

single phase this decision is often governed by the desire to reduce the reactor volume required for achieving a specified conversion level_ If it is desired to maximize the intermediate A 2 in a consecutive reaction scheme AL ~ A2 ~ A3 within a phase, plug flow conditions should be aimed for, i_e. "unmixed" concentration conditions within the reactor; this conclusion is analogous to the avoidance of intraparticle diffusion resistances (see Fig. 12)_ For highly exothermic reactions the dilemma is usually: to mix or not to mix ~ From a concentration viewpoint it is usually preferable to approach plug flow conditions, i.e. concentrations should not be mixed along the reactor. But from the point of view of temperatures, a thermally well-mixed system would be preferred. Consider the specific example of the oxidation of ethene to produce ethene oxide. This highly exothermic reaction is conventionally carried out in a cooled multitubular packed bed reactor. Close to the inlet of the reactor there is a temperature peak (hot spot)_ At increasing temperatures there is loss of selectivity to ethene oxide because of the parallel reaction to combustion products_ Figure 35 shows the simulated temperature and selectivity profile of an ethene oxide. From the point of view of maximum selectivity, it would be preferable to have a reactor concept in which the temperature is completely backmixed but the concentrations are unmixed (plug flow). The peak temperature can be "shaved" by a programmed energy removal strategy or a profiled axial catalyst activity; see Fig. 36. The energy flux removal near the inlet, "hot spot" zone needs to be higher than towards the exit of the reactor; this strategy is approached by using co-current singlephase coolant_ Using lower activity catalyst in the initial hot spot zone will also lead to a selectivity improvement. Figure 36 shows that either of the above strategies leads to selectivity improvements of

4048

R. KRISFINAand S_ T. S]E

0.8

.,. ...... ....

....

525

Selectivity to EO

"'"'"'"'"'"'"'"'"".......,......,

...... ..,

T/[K]

0.74 0

'

'

'

'

'

Reactor Length/[m]

500 12

Fig. 35. Temperature and selectivity profiles for the exothermic reaction of oxidation of etfiene to ethene oxide in a multitubular packed bed reactor. The simulations for the reactor temperature and selectivity were carried out using the kinetic data of Westerterp and Ptasinski (1984). The coolant temperature was constant along the length or the solvent; this was achieved by the use of evaporating water as a cooling medium.

programmed energy removal


.,..,.., .-"" "'-....

_i
r heat flux ~'tocoolant 08

axially profiled catalyst activity 1.0


.....

0.5 /

~co-current i'~;. N~oolant

to

EO
0.74 0 = ,

'~'~'~'L':"'.':
= , i I ~ , Reactor Length/[m] , , 12

Fig. 36. Selectivity profiles for base case with constant coolant temperature, co-current coolant strategy and axially profiled catalyst activity strategy. The base case chosen in the calculations is one in which the coolant temperature is constant and the activity profile along the length of the reactor is at unity level.
the order of 0.5-1%, which can be commercially significant. Another strategy for operating a packed bed reactor under near isothermal conditions is the use of the pancake reactor (Fig. 37), wherein the contact time is kept short by high severity operation. Heat removal at high temperatures is by use of molten salt (Blumenberg, 1992). A well-mixed temperature in partial oxidation reactions can be achieved by circulating the catalyst and removing the heat externally. The Du Pont process for production of maleic anhydride by oxidation of butane uses the circulating fluid bed reactor concept (Contractor and Sleight, 1988); see also discussions below on Strategy level III. As discussed earlier, catalyst circulation is possible only if the attrition resistance is improved. Circulating dense-phase fluidized bed reactors have good potential for carrying out exothermic gas-solid reactions, especially with deactivating catalysts (see Gianetto et al., 1990); this potential is as yet largely untapped In the Exxon process for partial oxidation of methane to produce syngas, well-mixed temperature conditions are realized by opting for a bubbling fluid bed operation (Eisenberg et al., 1994). The degree of backmixing has a significant impact on product properties in a polymerization process For polymerization of methyl methacrylate the standard deviation in diameter of the particles of the latex produced in batch, pulsed tubular and CSTR are compared in Fig. 38. A pulsed tubular reactor is to be favored over a CSTR for a controlled, narrow

Strategies for multiphase reactor selection

4049

crystalline

bn e --S t
mol

salt

e -~ - r f

silver catalyst d -- 5 mm deep

I formaldehyde
Fig. 37. Cooled pancake reactor concept of BASF for maintaining isothermal conditions in a multitubular reactor. Adapted from Blumenberg (1992).

36

CSTR

Standard deviation in particle diameter/[nm]


~

I averagedp= 100 nm ]
Pulsed Tubular

v
I I I I I

Batch
I

45

averageresidencetime/[min]
Fig. 38. Particle size distributions in batch, pulsed tubular and CSTR reactors for polymerization of methyl methacrylate. Adapted from Paquet and Ray (1994)_

size distribution of polymer product (Paquet and Ray, 1994). Debling et al. (1994) demonstrated the influence of the residence time distribution within different olefin polymerization reactor configurations on the grade transition performance of a process. A choice of a reactor configuration that appproximates plug flow is shown to achieve efficient transitions between grades. ll(c) Separation of product(s) in situ from the reactor zone, supply of reactants in situ, supply or removal of energy in the reactor in situ The main reasons for considering in situ removal of product(s) from the reaction zone are (1) to enhance conversion in equilibrium-limited reactions by shifting the equilibrium towards the right, (2) to prevent further, undesirable, reaction of products and consequently improved selectivity, and (3) as a remedy for product-inhibited reactions. The various techniques that can be considered for selective product removal are discussed below.

Extraction in situ, By deliberate addition of a second liquid phase containing a selective solvent, the desired product may be extracted from the reaction zone and further side reactions prevented (Br/indstr6m, 1983; Sharma, 1988). For example, in the bromination of dialcohols in aqueous phase, the problem is to prevent the second OH group from reacting with HBr to form the dibromide_ This can be solved by adding a hydrocarbon to the reaction mixture. The hydrocarbon extracts the monobromide, but not the dihydroxy compound or the HBr, from the reaction mixture. The monobromide is thus removed from the reaction mixture as soon as it is formed and thereby the action of HBr on it is prevented by phase separation (Br~indstr6m, 1983). In the Hofmann reaction of an amide with hypochlorite, the intermediate isocyanate can be extracted out to make isocyanates with high yields. In photochemical sulfoxidation of paraffins, water is used to extract sulfonic acid to prevent formation of polysulfonic acids. The epoxide product derived from 6-methylhept-5en-2-one is known to undergo very facile rearrangement to 1,3,3-trimethyl-2,7-dioxabicyclo[2,2,1]heptane. Under normal epoxidation conditions these two products, desired and undesired, are produced in about equal proportions. If the reaction is carried out in the presence of dichloromethane, the desired product is extracted to the hydrocarbon phase and a high yield (83-85%) of the epoxide is obtained with no detectable amounts of the rearranged product. The two-phase expoxidation strategy can be applied to a variety of olefins: cyclohexene, l-hexene and limonene (Anderson and Veysoglu, 1973). In the Ruhrchemie-Rh6ne Poulenc process for the production of butanal by hydroformylation of olefin, the in situ extraction of the desired product from the aqueous reaction phase prevents the formation of heavy ends; see Fig. 39 (Kuntz, 1987). Sharma (1988) considered several examples of reactions which would profit from introduction of an additional extractant phase. Reaction rate and selectivity in heterogeneously catalyzed liquid-phase hydrogenation can be influenced by the composition of the reaction medium in the vicinity of the liquid metal. This strategy can also be used for the selective hydrogenation of benzene to cyclohexene with powdered Pt or Ru catalyst. Introduction of an aqueous layer surrounding the catalyst particle is found to increase the selectivity towards cyclohexene (Wismeijer et al., 1986; Struijk, 1992); see Fig. 40. The desired product cyclohexene distributes preferentially into the organic phase and is thus prevented from further hydrogenation to cyclohexane. Asahi chemicals are believed to utilize such a concept in their commercial process. Supercritical extraction in situ. In equilibriumlimited biocatalyzed reactions the removal of the desired products, which are often thermally labile, by in situ supercritical extraction with carbon dioxide

4050

R. KRISHNAand S. T. SIE

Aqueous Phase CH2=CH-CHj + CO + H2--~Bt anal [ -~


H

heavyends

u-

Fig_ 39. In situ extraction of desired butanal in hydroformylation of olefins.

mml
Aqueous ZnSO 4 solution

lmm

nnll

thin aqueous film on catalyst contains hi~lh benzene concentration

0"-
Ru catalyst impregnated with aqueous ZnSO4solution
Fig_ 40. Use of aqueous layer on catalyst particles to enhance the selectivity towards cyclohexene during hydrogenation of benzene. Adapted from Wismeijer et al. (1986) and Struijk (1992). can lead to substantial benefits. In the lipase-catalyzed interesterification of triglycerides, for example: tricaprylin + methyl oleate ,--* 1-oleodicaprylin + 1,3dioleocaprylin + methyl caprylate, a high degree of incorporation of required fatty acids into triglyceride cannot be obtained because of its reverse reaction. Adschiri et al. (1992) applied supercritical carbon dioxide extraction to the removal of products from a liquid-phase reaction system as a means of solving the problem. Supercritical CO2 is non-toxic, and its critical temperature (304.2 K) is both sufficiently low for dealing with thermally labile materials and close to the optimal temperature for the enzymatic reaction. Significant improvement in the degree of incorporation of oleic acid into the triglyceride, above equilibrium values, was achieved in a batch reactor; see Fig. 41 (Adschiri et al., 1992); Aaltonen and Rantakylh (1991) listed the various advantages of using supercritical CO2 as a solvent in enzymatically catalyzed reactions, compared to aqueous media.
Distillation in situ. In situ product separation by distillation (Fig. 42) offers applications in esterification (e.g. for ethyl acetate), trans-esterification (e.g. for

butyl acetate), hydrolysis (e.g. for ethylene glycol, isopropyl alcohol), metathesis (e.g. for methyl oleate), etherification (e.g. for MTBE, ETBE, TAME) and alkylation reactions (e.g. for cumene). The most dramatic example of the benefits of in situ separation within the reactor is afforded by the Eastman Kodak process for methyl acetate where the conventionally used reactor followed by several distillation columns was replaced by one integral reactive distillation column with considerable economic advantages (Doherty and Buzad, 1992). The in situ distillation concept has gained considerable attention recently for carrying out catalyzed liquid-phase reactions such as etherification; the catalyst in this case is usually incorporated in the form of a structured packing (DeGarmo et al., 1992).

M e m b r a n e reactor f o r selective removal o f product.

Falconer et al. (1993) and Saracco and Specchia (1994) give several examples of in situ separations using membrane for the purpose of improving yield and selectivity. The basic concepts involved are shown schematically in Fig. 43. The majority of reactions

Strategies for multiphase reactor selection

4051
ole

cap + Me. ole

ole cap + cap 1II

cap + Me. cap ole

Extractive Reaction at 10 MPa


Batch Reaction

~'~-f conversion OFV.. . . . . . 0 time/[h]

at 0.1 MPa 14

Fig. 41. Supercritical extraction of product in situ. Adapted from Adschiri et al. (1992).

Recycle stream Inerts /_]._~ errs


Hectitlcation

zone Distilla ion Colbl in Inerts L~ I Reactants + Inerts "/.

Reactive
distillation zone

"--~ Product

zone "-~roduct

Stripping

Fig. 42. Comparison of (reactor-followed-by-distillation) concept with (reactor-with-in situ-distillation) concept. Adapted from DeGarmo et aL (1992).

(a)

(b)

CE$ 49:Z~,-D

Fig 43. Membrane concept for in situ product removal to (a) shift equilibrium or (b) prevent side reactions_ Adapted from Falconer et aL (1993).

4052

R. KRISHNA and S. T. SIE

studied in catalytic membrane reactors have taken advantage of the ability of membranes to selectively remove a product, usually hydrogen, through Pd alloys or by Knudsen diffusion through microporous glass or ceramic membranes. For example, for propane dehydrogenation, significant improvements over conventional packed bed reactors have been realized (Ziaka et al., 1993). Selectivity can also be enhanced when only the desired product of a set of equilibrium reactions is selectively removed through the membrane, or when the permeating component is an intermediate product of a set of consecutive reactions. Matsuda et al. (1993), during i s o b u t a n e dehydrogenation in a Pd membrane reactor, noticed that the selectivity to i s o b u t e n e was enhanced by in situ hydrogen removal through the membrane due to suppression of undesirable side reactions such as hydroisomerization and hydrogenolysis. Other reactions where the use of permselective membranes in catalytic reactors can be expected to lead to significant improvements include dehydrogenation of ethane, cyclohexane and ethylbenzene. For example, an experimental study by Becker et al. (1993) showed that use of this concept for dehydrogenation of ethyl benzene results in a 20~o increase of conversion over the conventional fixed bed operation. A survey of potential applications of inorganic membrane reactors is given by Falconer et al. (1993), Hsieh (1991) and Saracco and Specchia (1994). The Knudsen selectivity achieved in glass or ceramic membranes is limited; much higher selectivities can be obtained by making use of the principle of shape selectivity, possible with membranes made of microporous materials such as zeolites or carbon molecular sieves. For equilibrium-limited isomerization of paraffins, it is advantageous to remove the normal paraffins in situ in the reactor. Two membrane reactor

concepts are sketched in Fig. 44; these concepts have the potential of achieving near to complete conversion to either isoparaffins or n-paraffins from a mixed paraffin feed (Sie, 1994). Membrane reactors may well be the future for many -equilibrium reactions requiring effective removal of water--a small molecule that usually boils at higher temperatures than reactants, making it tough to remove by continuous distillation_ Polymeric membrane reactors have been used for esterification of oleic acid and acetic acid with ethanol in the presence of an asymmetric polyether imide membrane; in this case, selective permeation of water allows 100~ conversion levels to be achieved (Kita et al., 1987). Du Ponrs fluorinated sulfonic acid polymers have been found useful in the separation of carboxylic acids. Valeric, propionic and butyric acids all diffuse through Nation 125 in protonated forms with essentially identical rates whereas functional acids (levulinic, succinic, lactic) flow at a significantly lower rate (a half to one-fifth). In a process involving the reduction oflevulinic to valeric acid the product can be removed selectively (Chum et al., 1983). An increasing number of bioreactor studies have focused on immobilized cell systems because their high cell density leads to improved volumetric productivity. But sometimes cells reduce their activity in the presence of the product. Concurrent production and recovery help in such situations. One well-known system of feedback inhibition is the fermentation of glucose to ethanol by Saccharomyces cerevisiae. For this process Steinmeyer et al. (1988) report on the operation ofa multimembrane bioreactor system with an extraction solvent (see Fig. 45). Membranes in the reactor separate gas, cell, nutrient and solvent layers. The hydrophobic membrane that separates the gas phase from the cells allows for the removal of carbon

H2

Product: iso paraffins

isomel ization cataly,'

Feed:
iso + n

Feed:
iso + n

oduct: . paraffins
Fig. 44. Membrane reactor concepts for paraffin isomerization. Adapted from Sie (1994).

Strategies for multiphase reactor selection

4053

Hydrophobic membrane Hydrophilic membrane

Hydrophobic membrane

Fig. 45. Multilayer membrane reactor for biochemical application. Adapted from Steinmeyer et al. (1988).
Hz

H~

Dinitdle --> Intermediate -> D i a m i n e

Fig. 46. In situ supply of hydrogen by incorporation of metal hydrides within catalyst. Adapted from Snijder et aL (1992).

dioxide from the cell layer and allows for the entry of oxygen, which is needed for lipid synthesis in yeast. A hydrophilic membrane separates the cell layer from the nutrtent layer and allows the medium to readily pass through it. Nutrient supply and ethanol removal are achieved by a combination of diffusive and convective flows across this membrane. A hydrophobic membrane separates the nutrient and solvent layers. The solvent wets the membrane but does not pass through it to the nutrient side if a critical pressure differential is maintained. Successful compartmentalization of the solvent prevents emulsification and subsequent toxic effects on the cells. Solvent selectivity removes ethanol from the nutrient, thereby lowering ethanol inhibition of fermentation and increasing reaction rate and extent. In the fermentation process for acetic acid, the use of electrodialysis for separation of product overcomes product inhibition and improves the space time yield considerably (Von Eysmondt and Wandrey, 1990). Another example is presented by Tichy et al. (1990) where not only the product inhibition by amino acid

methionine is resolved by electrodialysis but where the losses of the substrate (in this case hydroxymethylthiobutyric acid) for the fermentation are reduced as well. The hydrogenation of dinitriles to diamines is of practical importance since the diamines are used as the monomer for the production of nylons. The hydrogenation of succinonitrile is usually thought to proceed via reactive intermediate imines (Mares e t a / . , 1988). Deficit of hydrogen within the catalyst particle can cause excessive by-product formation. Van Swaaij and co-workers (Snijder et al_, 1992) developed a novel method for carrying out catalytic hydrogenation reactions. They incorporated metal hydride particles within the catalyst particles which served as in situ suppliers of hydrogen; see Fig. 46. For succinonitrile hydrogenation the selectivity of the diamines can be expected to increase markedly in comparison with conventional catalyst systems without metal hydride incorporation. This selectivity improvement is particularly significant at lower operating pressures. Incorporation of metal hydrides can therefore provide

4054

R. KRISHNA and S. T. StE

a means for reducing the severity (pressure) of hydrogenation reactions in the fine chemicals industry.
Removal o f energy in situ. Removal of the heat of reaction in situ is often effected by use of an

molten phthalic ~hydride

evaporating component of the reaction medium. This can be a product (as in reactive distillation) or a solvent. For example, in the sulfuric acid alkylation process sketched in Fig. 29, the concept of autorefrigeration is used. Vaporization of a portion of the i s o b u t a n e and propane in the feed mixture is used to control the temperature of each stage. NIl l ( d ) Counter-, co-, or cross-current contacting

lalimide

The decision whether to adopt counter-, co-, or cross-current contacting of two phases is dictated by factors such as equilibrium limitations, flooding, pressure drop and degree of conversion required. For gas treatment applications such as absorption of CO2, H2S and COS using amines where high conversion levels are usually desired, it is common to adopt counter-current operation from considerations of phase and reaction equilibrium. For the equilibriumlimited hydration of butenes to produce tert-butyl alcohol, Fig. 47 illustrates the significant advantage of counter-current operation over the conventionally used co-current trickle flow reactor (Jansen et al., 1995). For the equilibrium-limited reaction of gaseous ammonia with phthalic anhydride to yield phthalimide, a continuously operated multistage counter-current reactor offers significant advantages over batch operation (Bartholom6 et al_, 1978) and a high purity product containing a negligible amount of unreacted anhydride is obtained; see Fig. 48.

Fig. 48. Continuous counter-current multistage contacting for production of phthalimide. Adapted from Bartholom~ et al. (1978). Co-current gas-liquid downflow trickle bed reactors are widely applied for hydroprocessing of heavy oils. This co-current mode of operation is disadvantageous in most hydroprocesses, and counter-current flow of gas and liquid would be much more desirable. This is because reactions such as hydrosulfurization and hydrogenation are inhibited by the hydrogen sulfide formed, even when using so-called sulfur-tolerant catalysts of the mixed sulfide type. The removal of sulfur from heavy oil generally follows second-order kinetics in sulfur concentration, which is a reflection of the presence of a variety of sulfur-containing compounds with different reactivities. The secondorder kinetics imply that a relatively large proportion of sulfur is removed in an early stage of the process (due to conversion of the bulk of reactive molecules) while removal of the remaining sulfur takes place

H=On n butene H=O

countercurrent

cocurrent TBA

T B A Y U butene

Conversion of
butene
awt a aS

t S

-j
., adiabatic

0 Catalyst Weight / [103 kg]

170

Fig. 47. Advantages of counter-current contacting over co-current contacting for hydration of butenes to tert-butyl alcohol. Adapted from Jansen et al. (1995).

Strategies for multiphase reactor selection


100

4055

2 ...............

% HDS

wt%S in oil

O0

% of catalyst bed
... ......

100

[bar]

t."
0 I0
,

cu.en,
, ~ : ,

% of catalyst bed

100

Fig. 49. Partial pressure profile of hydrogen sulfide in counter-current and co-current operation during hydrodesulfurization of gas oil. much more slowly in later stages. This means that the bulk of the H2S is generated in a small inlet part of the bed and that this H2S exerts its inhibiting influence in the remaining part of the bed. Figure 49 shows the profiles of sulfur in liquid and H2S in the gas phase for hydrodesulfurization. It can be seen that in co-current operation the larger part of the bed operates under a H2S-rich regime; see Fig. 49. The situation is clearly more favorable in the countercurrent mode of operation since in this case the major part of the bed operates in the H2S lean regime. The co-current mode of operation is particularly unfavorable since the inhibiting effect is strongest in the region where the refractory compounds have to be converted, which calls for the highest activity. A similar situation exists in hydrocracking The by-product of conversion of nitrogen-containing organic compounds, namely ammonia, is a very strong inhibitor for hydrogenation and particularly for hydrocracking reactions. For the hydrogenation of aromatics, too, the co-current operation is unfavorable_ This is not only so from a kinetic point of view (inhibition by H2S and NH3), but also because of thermodynamics (Trambouze, 1990). Deep removal of aromatics from an oil is generally limited by thermodynamic equilibria. In the co-current mode of operation the partial pressure of H 2 at the exit end of the reactor is lowest because of the combined effects of pressure drop, hydrogen consumption and build up of gaseous components other than H 2 (H2S, NH 3, H20, and light hydrocarbons). Problems with counter-current operation, however, are excessive pressure drop and flooding limitations. To overcome these problems larger sized (say 5 mm) "shaped" catalysts could be considered, in the form of Raschig rings (Trambouze, 1990). Note that in order to fulfill the requirements of Strategy level II the Strategy level I decision needs reconsideration. Alternative conceptual catalyst configurations for

counter-current G - L contacting are shown in Fig. 50. The philosophy behind these configurations is that typically in hydroprocessing of heavy oils the gasliquid mass transfer is not a limiting factor. By allowing gas and liquid to flow through separate channels, with the possibility of contacting each other at regular intervals, the momentum losses are reduced by sacrificing gas-liquid contacting efficiency. Figure 50(a) shows a possible construction in which a granular catalyst with a particle size around 1 mm is enclosed between horizontal screens. The screens are made from wire mesh or perforated plate. The upper screen is a flat one, while the lower screen has folds which serve as the vertical passages for gas. In the "tea bag" concept of Fig. 50(b), the filling consists of catalyst containers which have permeable walls that can be made from wire mesh or perforated sheet. These containers can have a spherical shape with diameters of the order of a few centimeters. The screens containing loose particles can be dispensed with if these particles are bonded together to form aggregates which can be considered as porous macroparticles. The techniques to form such agglomerates may involve sintering, cementing, superficial peptization of the particles, etc. The technique should not affect catalytic performance in a negative way and should also produce macroparticles of sufficient strength. Cross-current contacting of phases is the strategy to be adopted when the contact between the phases needs to be limited rather than maximized. Some processes such as the dehydrogenation of propane to propene and isobutane to isobutene require a short (typically less than 3 s) contact time between the gaseous reactants and a solid catalyst. Moreover, the residence time distribution of the gas phase should be narrow, i.e. the aim should be for plug flow in order to achieve high selectivity and yield. This requirement is due to consecutive reactions where the desired product is the intermediate in the reaction sequence. Furthermore, for such processes, continuous catalyst regeneration is necessary to remove coke deposited during the reaction. Wolffet al. (1995) describe a novel circulating cross-flow moving bed reactor system for such applications; see Fig. 51. In this concept the variable gas residence time is independent of the longer, solids residence time. The latter can be adjusted by the superficial gas velocity and reactor inclination angle_ STRATEGY LEVEL Ill---CHOICE OF
H Y D R O D Y N A M I C FLOW REGIMES

The choice of the "fluidization" regime of operation is to a large extent a logical consequence oftbe various decisions already reached at Strategy levels I and 1I with respect to particle sizes, phase hold-ups, desired backmixing characteristics of each of the phases, etc.
Gas-solid systems Basically, a choice has to be made between the following five modes of operation (see Fig. 52).

4056

R. KRISHNAand S. T. S~E

IL;
lit

I|1

l,;
i

iii

II

J'J'"l

"""

1""'"1

Fig. 50. Three-levcls-of-porosity concept for counter-current contacting of gas and liquid phases. (a) Catalyst enclosed in wire-screen structure, (b) the "tea bag" concept, (c) agglomerated catalyst.

Air-slide reactor

~1, i C4=, H2
I / I

. i"'""

Risel
N2

y
'N2

,-2

talyst I regenerator I

Fig. 51. Cross-current air-slide reactor concept for dehydrogenation of isobutane. Adapted from Wolff et al. (1995). (1) Packed bed regime (.fixed or moving bed operation). Here the particle hold-up is typically in the region 0.5-0.7. The particle size suitable in the packed bed regime is usually larger than 1 mm because smaller particle sizes result in unacceptably high pressure drops. The plug flow character of the gas phase is an important advantage of the packed bed reactor. (2) Bubblin.q bed operation. The particle hold-up in this regime is typically 0_4 0.5 and this regime is characterized by the presence of fast-moving bubbles that tend to churn the system, resulting in an almost completely backmixed solids phase. A part of the entering gas phase serves to keep the particles in suspension and this portion of the gas is also completely backmixed; see Fig. 53. If staging of the solids is desired then a multistage fluid bed concept can be employed, either disposed horizontally or vertically, as illustrated later for the oil shale reactor configuration selection. The heat transfer coefficient to installed heat exchange surfaces in bubbling fluid beds is about a factor of five higher than in packed bed operation and this is often an advantage for reactions involving high exothermicity_ (3) Turbulent bed and (4) dense-phase transport reactor. If the gas velocity is increased further beyond the bubbling fluidization regime the turbulent regime of fluidization is reached. In this regime the bubbles are of indistinguishable and ever-changing shape. The

Strategies for multiphase reactor selection

4057

Moving Bed Reactor

Bubbling Fluidized Bed Reactor

Turbulent Bed Reactor

Dense Phase Riser Reactor

Dilute Phase Riser Reactor

Fig. 52. Gas-solid reactors operating in various "fluidization" regimes.

product vapor product vapor

I .,~r~ . . . , . I

model
~ ~ ' !.
dilute phase in plug flow axially dispemed
dense

phase

U
feed vapor

(ufeed vapor

Fig. 53. Backmixing characteristics of dilute and dense phases in bubbling fluid bed. particle hold-up is typically in the range 0.3-0.45. There is heavy entrainment of the solids and bed inventory would be lost without solids recycle by means of a cyclone. If the gas velocity is increased still further, the bed can be transported and this mode of operation is commonly termed "fast" fluidization or dense-phase riser transport. To maintain bed inventory, the contents of the reactor have to be recirculated and the reactor is also called the circulating fluid bed reactor. The particle hold-ups in this regime are typically 0.1-0.2. The backmixing characteristics of the dense-phase transport reactor are shown in Fig. 54, The exchange between the gas and solid phase is much better than in the bubbling bed operation. (5) Dilute-phase riser transport. As the gas velocity is increased still further, the dilute-phase riser transport regime is reached. The particle hold-up in this regime is of the order of 0.05 or smaller. The gas and solid phases are virtually in co-current plug flow through the reactor (see Fig. 55). The first major decision is whether to keep the solids fixed (in a packed bed) or to move the solids in a moving bed or fluidized bed. This choice is largely dictated by catalyst deactivation kinetics and the time interval between successive regenerations. If this time interval is of the order of 1 year, fixed bed operation is usually preferred. If the time interval is of the order of 1 week, swing type operation using two beds is usually opted for; swing operation is, for example, used in the regenerative naphtha reforming technology_ If the time interval between successive regenerations is of the order of a day to a few hours, then moving bed operations can be considered such as in the CCR technology for naphtha reforming. If the time between successive regenerations is of the order of less than 1 h, then fluidized bed operation is called for_ The decision to transport the solids to and from the reactor is a crucial one because solids motion introduces several complications such as attrition and blockage. If the catalyst is expensive (e.g. Pt based) or environmental hazardous (e.g. Cr), it is usually not

4058
product

R. KR.ISI-INA and S. T. SIE vapor product ~ vapor U plug flow


.:.v..,

_t_
I IO0 ee

deactivated catalyst

of

".." 4". ..-.. -..'-" ";: :~_a:,

deactivated catalyst model

~apc

axially BB ~dispersed solids

cagaely:;ated ~fe regenerated I catalyst

ed

feed ~i~

Fig. 54. Backmixing characteristics of gas and solid phases in circulating fluid beds.

deactivated

catalyst'

:)roduct
~

product ~lX
va

~ deactivated
/ catalyst

vapor

000

'":
,::.~

.%.'.1
model
~-

Soo I

,..,
::i
' ~.=,

regenerated ::; catalyst :'.;

k3;

I~J e ed

feed

f regenerated | catalyst

Fig 55. Backmixing characteristics of gas and solid phases in dilute-phase risers. advisable to fluidize it because of inevitable losses through cyclones. Other reasons, distinct from catalyst deactivation, why fluidized bed operation is chosen could be the desire to use particles smaller than say 1 mm (such particles are usually not allowable in packed or moving bed operation due to excessive pressure drop). Another important reason to opt for fluidized beds may be the need to remove or supply large amounts of heat, taking advantage of the high heat transfer coefficients in fluid beds or the possibility to transport heat via flow of particles. Sometimes, the desire to have a completely thermally backmixed system (Strategy level II) would dictate the use of bubbling or circulating fluidized beds. This is the case for combustion of coke from deactivated FCC catalyst. Traditionally, FCC regenerator designs have adopted the bubbling fluid bed regime, because of the good backmixing of the emulsion phase. However, the major disadvantage of bubbling bed regenerator designs is oxygen slip due to poor mass transfer from bubbles; deep beds--typically 8 m in height--are required even for moderate conversion levels of about 90~o. The dense-phase riser transport regime, which is the current choice for FCC regenerator operation, has vastly superior gas-to-solid mass transfer characteristics and is the regime currently favored for newer designs. In Fig. 56, the gas-to-solid mass transfer characteristics of the bubbling bed, dense- and dilute-phase transport regimes are compared with the effective chemical reaction rate. It is clear that the optimum regime of operation is the dense-phase transport regime and indeed this is the regime of choice in modern FCC designs.

Strategies for multiphase reactor selection

4059

10

kinetic controlled

~ mass transfer ! controlled Van der Ham, Prins and Van Swaaij (1993)

krnPpF--p

[s'l]
kGa

I particle size
! phase riser
~?'4E

dp=60#m

[s-q
0.1

dilute phase riser

bubbling j bed

kGa

0.01 10 -3

10 ~ 10 -~ particle holdup,

Fig. 56. Gas-to-solid mass transfer characteristics of various regimes or operation for regeneration of coked catalyst.

fine particles

coarse particles

Gas-hquid systems
For upflow of gas through liquids in vertical columns, there is complete correspondence of flow regimes with gas-solid systems; see Fig. 58. The analogue of the homogeneous fluidization regime is the homogeneous bubbly flow regime. The bubbling fluid bed operation has a complete parallel in the churn-turbulent regime in gas-liquid systems (see Krishna, 1993). The choice between the various gas-liquid regimes, therefore, parallels the analysis for gas-solid systems; with increasing reaction rate the regime choice moves from left to right in Fig. 58. One possible starting point in the choice of the flow regime is consideration of the parameter #, already chosen in Strategy level I. Regimes to the left of the flow regime map of Fig. 58 correspond to high # (the choice for relatively slow liquid-phase reactions) whereas towards the right are low # values (a choice for high rates of liquid-phase reactions). To achieve low fl values it could be possible, for example, to operate in the spray regime in a tray column. The analogue of the turbulent or "fast" fluidization regime of gas-solids flow is the regime called turbulent bubbly flow in Fig. 58; this is the regime prevalent in air left fermentors, for example. In air left fermentors the aim is for good gas-liquid mass transfer to prevent oxygen depletion in the liquid phase. Both turbulent, or "fast", fluidization, and turbulent bubbly flow regimes are gaining in importance for similar reasons: they have superior mass transfer characteristics. A deeper appreciation of the analogies between gas-solid and gas-liquid systems will be helpful in reactor selection and could facilitate scale-up (Krishna, 1993). In the design of the Fischer-Tropsch bubble column slurry reactor, an important consideration is the choice between the homogeneous and heterogeneous

..

;.i)IA111:_',

10

kGa 1
[s'q0.1 0.01

i" . . . .
"

0.01

0.1
dp/[mm]

10

Fig. 57. Influence of particle size on bubble to dense-phase mass transfer in bubbling gas-solid fluid beds.

Another strategy to overcome the bubble to dense-phase mass transfer limitations of bubbling fluid bed operation is to resort to larger particle sizes because the interphase mass transfer coef~cient is roughly proportional to the particle size; see Fig. 57. To overcome the concomitant problem of intraparticle diffusion within these larger particles it may be necessary to resort to the use of egg shell catalyst (back to Strategy level I!).

4060
Fixed bed

R. KRISHNAand S. T. StE
Homogeneous

Heleroeneous

Slug
flow

Turbulent
[luidization

Dilute
phase dser

fluidization Gas-Solid Fluidization Regimes

fluidization

|::~1

Stagnant Homogliquid eneous bed bubblyflow Gas-Liquid Flow Regimes

Heteroeneous regime

Slug flow

Turbulent bed

Spray flow

Fig. 58. Gas-solid and gas-liquid flow regimes in vertical columns with up-flow of gas.

:1

id

as

]as

liquid homogeneous bubbly flow regime U-O.04 m / s

liquid heteroaeneous or churnturbulent flow regimo U=Q36 m/s

Fig. 59. Homogeneous and heterogeneous flow regimes in bubble column operation. flow regimes; see Fig. 59. The scale-up of a bubble column slurry reactor is limited to vessels of about 6 m diameter because of mechanical considerations. Operation in the homogeneous flow regime can be realized only at a superficial gas velocity U~ below about 0.04 m s -1 while in the heterogeneous flow regime, Ug can be as high as say 0.36 m s-~, which holds the promise of a factor of nine increase in the throughput per reactor of given diameter. The heterogeneous flow regime is therefore preferred in practice. gas. At temperatures below 400C the reaction is extremely slow and, at temperatures exceeding about 550C, excessive cracking of the oil vapor liberated during the reaction takes place. Burning off the coke from the spent shale in a combustor provides a source of energy for the endothermic pyrolysis reaction; see Fig. 60. There are numerous processes and reactor configurations that have been suggested for carrying out the thermal pyrolysis; these are sketched in Fig. 61, which is adapted from Levenspiel (1988). The various technologies have apparently little in common. For example, on the basis of particle size used in the processes (Strategy level I), there are: (a) large sized ( ~ 50 mm) particles, (b) medium sized ( ~ 5-10 ram) particles and (c) small sized ( < 3 ram) particles. From considerations at Strategy level II, the contacting pattern between the oil vapor ( + stripping gas) and the solid phase used in these technologies also varies widely: (1) counter-current in a-l, a-2 and c-2, (2) co-current in technologies b-1 and b-2 and

CASE STUDY OF OIL SHALE REACTOR SELECTION Having discussed the various strategies for selection of a multiphase reactor, these ideas and concepts can be applied to a specific example of the oil shale reactor selection_ Oil shale contains an aromatic component called kerogen, which on heating in the temperature range 400-550C decomposes to yield oil, coke and

Strategies for multiphase reactor selection

4061

[kerogen ] ~ { '~Heat

bitumen]-------~[Oil vapor+J ~ . . ~ '~Heat ["~'~ Oil vapor + gas


:leactor

Oil shale
I

Combustor coke I

Fig. 60. The oil shale reactor selection problem. Adapted from Krishna (1994).

(a) Large(about50 mm) particlesin packedmovingbedsG G

,1I

o1I,

oll s
a-3

slid/~l~l" as

/ ~

....... so.,.,.! .....................

.............................

I
" c-1
c-_~2

b-1

b.~lL,d~
2 ~

(b) Medium sized (5-10 mm) particles in moving beds

(c) Finesized ( < 3 mm) particles in fluidizedbeds

Fig. 61. The reactor choices available for the oil shale reactor. Adapted from Levenspiel(1988) and Krishna (1994). (3) cross-current in a-3. The solids phase is more or less well mixed in the SPHER process (c-I) whereas in the other technologies there is staging (i_e. plug flow) of the solids phase. The Unocal process involves another unique technology in which the shale particles are moved upwards, counter-current to the vapors, by means of a rock pump_ Analysis at Strategy level III, shows that both packed (moving) bed [Groups (a) and (b)] and fluidized operations [Group (c)] are encountered. For a reactor engineer involved in developing a shale oil recovery process the diversity of the reactor configurations shown in Fig. 61 is more than a little disconcerting. Poll et aL (1987) and Krishna (1994) have followed a systems approach to selection of the "ideal" oil reactor. First, a "wish" list is set up.
Wish I. The reactor must be capable of maximum recovery of oil. Oil recovery can be maximized by

making sure that high (say >/99%) conversion of kerogen is obtained and that the oil vapor once produced does not suffer further cracking and degradation to light gases.
Wish 2. The reactor design should allow scale-up to large scale units capable of handling say 50,000 t per day of oil shale in one processing train; this is important for economy of scale. Wish 3. In view of the large quantities of shale rock to be handled, of the order of 500 kg s- ~, there is need to restrict the reactor volumes in order to reduce the investment costs. Wish 4. During the grinding operation there is inevitable production of fines and the chosen reactor should be capable of handling these fines, both from economic as, well as environmental considerations.

4062

R. KRISHNAand S. T. SIE none ofthe "ideal ingredients" chosen at each strategy level is unique. Fine particles smaller than 3 mm are used in the SPHER [Fig. 61(c-1)] and Chevron [Fig. 61(c-2)] processes. Cross-current contacting is employed by Superior Oil [Fig. 61(a-3)]. The Chevron process uses a counter-current multistage fluidized bed. However, the combination of these ingredients, yielding the configuration shown in Fig. 62, was considered to hold sufficient promise of improved yield and economics to justify a substantial development effort (Poll et al., 1987). There were also no scale-up problems envisaged in the multicompartment fluidized bed approach; for scaling-up purposes it is sufficient to study in detail the hydrodynamics of one of the fluid bed compartments. In theory, to obtain sufficient solids-phase residence time, the number of compartments can be increased at will without running into any scale-up difficulties.
CLOSING REMARKS

On the basis of the wish list, one can proceed to make the choices at each strategy level.
Strategy level I For maximizing the yield ofoil from kerogen (Wish 1) the particle size has to be restricted to below 2 mm; this decision is made on the basis of a careful analysis of the kinetics of oil shale production and overcracking resulting from diffusion limitation within large sized particles. Strategy level II Strategy level lI(b). Since large throughputs have to be handled (Wish 3), conditions of plug flow of the solids has to be strictly adhered to because otherwise unacceptably large reactor volumes will be required. Strategy level II(c) and (d). The oil vapor, once formed, should be removed from the reaction zone and should not come into further contact with hot shale particles. Co- and counter-current contacting are undesirable and the ideal contacting pattern is cross-current contacting of gas and solid. Crosscurrent contacting also ensures in situ removal of oil vapor from the reaction zone. Strategy level III At Strategy level I a decision has been made to use particles smaller than 2 mm. Now, during the crushing operation, fine particles (much smaller than the specified top size of 2 mm) will also be produced. The "fluidization" regime should be capable of handling the produced fines (Wish 4) and this leads to the choice of a bubbling or turbulent fluidized bed operation. Combination of the decisions made at the various strategy levels leads to the cross-current multistage fluidized bed concept developed by Shell (Poll et al., 1987; Krishna, 1994); see Fig. 62_ It is interesting to note that this "ideal" concept did not correspond to any of the "known" reactor configurations shown in Fig. 61. A careful comparison of the Shell process with the existing technologies (Fig. 61) shows that

In this article the use of a systematic, structured, approach to reactor selection has been advocated. The reactor selection exercise is conveniently split into three, separate, sequential decisions on the following sub-sets, or attributes: Strategy level I. Strategy level II. Strategy level III. "Catalyst" design Reactant injection and dispersion strategy Multiphase hydrodynamic flow regimes.

The choice of each item at the various strategy levels discussed in the text is made on the basis of a wish list, set up right at the beginning of the reactor selection exercise. In the discussions, it has been stressed that a disciplined approach may unravel novel ways of improving reactor performance; this has been illustrated by means of several examples. The three-level strategy approach presented here has been mainly developed for application within the petroleum and petrochemical industries and, therefore, is restricted to continuous "steady-state" processing

OilVapor+Gas
I i .~

Raw Shale

s ., h~ . p S a . ~p e l n t.

, shale

Stripping Gas
Fig. 62. The cross-current multistage fluidized bed reactor concept developed by Shell (Poll et al., 1987). Adapted from Krishna (1994).

Strategies for multiphase reactor selection on a large scale. It is worthwhile examining the extension of these concepts to unsteady-state processing and batch operations in greater depth. Also, there may be other ways to structure the reactor selection problem to suit one's own work environment and technology culture. Acknowledgements--R.K. and S.T.S. acknowledge financial support from the Onderzoekschool Procestechnologie (the Dutch National Graduate School in Chemical Engineering). R.K. also acknowledges financial sponsorship of a research programme on reactor selection from the Netherlands Foundation for Chemical Research. Valuable suggestions and comments on the manuscript were received from A. Bliek, C. M. van den Bleek, A. N. R. Bos, A. A. H. Drinkenburg, J. M. H. Fortuin, G. J. Harmsen, M. Kuczynski, E. K. Poels and W. P. M. van Swaaij. W.P.M. van Swaaij provided several new ideas and suggestions while collaborating with R.K. in preparing a paper entitled "Potential of new reactor developments in fine chemicals manufacture" for presentation to Rh6ne-Poulenc, Lyon, France. M. M. Sharma made several valuable suggestions to R.K. on the subject of fine chemicals reactor engineering. NOTATION
a

4063

residence time, s &on generalized Thiele modulus diI = ~ ~/ D, rf

Subscripts eft effective parameter g referring to gas phase gen generalized parameter I referring to liquid phase L1, L2 referring to the two phases in liquid-liquid systems m referring to the main (desired) reaction p referring to particle; also referring to the poisoning reaction
REFERENCES

dp

D~rr
DI

Drelativ

E F

interracial area per unit reactor volume, m particle diameter, m effective diffusitivity within catalyst particle, m2s-i liquid-phase diffusivity, m 2 s - 1 relative value of diffusivity inside zeolite enhancement factor for gas-liquid reaction effectiveness factor for storage of metals deposits in the pore volume of the catalyst kx/~D~ kl gas-phase mass transfer coefficient, m s - I pseudo-first-order reaction rate constant for homogeneous liquid-phase reaction, s-L liquid-phase mass transfer coefficient, m s pseudo-first-order reaction rate constant for catalytic reaction, defined per kg of catalyst, m 3 kg - l s-1 characteristic length of particle, m pore volume, ml gas constant, 8.314 J m o l - ~ K - L external surface area of particle, m 2 temperature, K volume of particle, m 3 Hatta number Ha =

Ha
ko

kl
kl

ks

L PV R SA T V

Greek letters fl ratio of liquid-phase volume to diffusionlayer volume/~ = - 61 thickness of diffusion layer of liquid phase Di fractional hold-up of liquid phase particle hold-up effectiveness factor of catalyst particle particle density, kg m -3

et ~p r/ pp

Aaltonen, O. and Rantakyl/i, M., 1991, Biocatalysis in supercritical CO 2. Chemtech No. 4 (April), 240-248. Adschiri, T., Akiya, H. and Chin, L. C_, 1992, Lipasecatalyzed interestefitication of tfiglyceride with supercritical carbon dioxide extraction. J. Chem. Engng Japan 25, 104-105. Anderson, W. K. and Veysoglu, T., 1973, A simple procedure for the epoxidation of acid-sensitive olefinic compounds with m-chlorperbenzoic acid in an alkaline biphasic solvent system. J. Org. Chem. 38, 2267-2268. Bailey, J. E., Horn, F. J. H_ and Lin, R. C., 1971, Cyclic operation of reaction systems: effect of heat and mass transfer resistance. A.I.Ch.E.J. 17, 818-825. Bartholom6, E., Hetzel, E, Horn, H. C_, Molzahn, M., Rotermund, G. W. and Vogel, L., 1978, Chemical engineering problems in process development. Int. Chem. Engng 18, 381-386. Becker, E. R. and Wei, J., 1977, Nonuniform distribution of catalysts on supports. J_ Catal. 46, 363-381. Becker, Y_ L., Dixon, A_ G_, Moser, W. R. and Ma, Y_ H., 1993, Modelling of ethylbenzene dehydrogenation in a catalytic membrane reactor, J. Membrane Sci. 77, 233244. Biegler, L. T. and Hughes, R_ R., 1983, Process optimization: A comparative case study. Computers and Chemical Enong 7, 645-661. Blanks, R. F., Wittrig, T. S. and Peterson, D. A_, 1990, Bidirectional adiabatic synthesis gas generator. Chem. Engng Sci. 45, 2407-2413. Blumenberg, B., 1992, Chemical reaction engineering in today's industrial environment. Chem_ Engng Sci. 47, 2149-2162. Bokhoven, C. and Van Raayen, W., 1954, Diffusion and reaction rate in porous synthetic ammonia catalysts. J. Phys. Chem. 58, 471-476. Boreskov, G. K. and Matros, Y_ S_, 1983, Unsteady performance of heterogeneous catalytic reactions. Catal. Rev. Sci. Engng 25, 551-569. B#indstr6m, A., 1983, Chemical protection by phase separation and phase transfer catalysis. J. Mol. Catalysis 20, 93-103. Chum, H. L., Hauser, A. K. and Sopher, D. W_, 1983, New uses of Nation membranes in electroorganic synthesis and in organic acid separations_ J. Electrochem. Soc. 130, 2507-2509. Colen, G. C. M., Van Duijn, G. and Van Oosten, H. J., 1988, Effect of pore diffusion on the triacylglycerol distribution of partially hydrogenated trioleoylglycerol. Applied Catalysis 43, 339-350. Contractor, R. M. and Sleight, A. W., 1988, Sel~tive oxidation in riser reactor. Catalysis Today 3, 175-184. Contractor, R_ M., Bergema, H. E., Horowitz, H. S., Blackstone, C_ M., Malone, B., Toraridi, C. C., Griffiths, B., Chowdhury, U. and Sleight, A. W., 1987, Butane

4064

R. KRISHNA and S. T. SIE Krishna, R., 1994, A systems approach to multiphase reactor selection. Advances in Chemical Engineering, J. Wei (Editor-in-chief), Vol. 20, pp. 201-249, Academic Press, New York. Kuntz, E. G., 1987, Homogeneous catalysis in water. Chemtech No. 9 (September), 570-575. L.afarga, D., Santamaria, J. and Menrndez, M., 1994, Methane oxidative coupling using porous ceramic membrane reactors--I. Reactor development. Chem_ Engng Sci. 49, 2005-2013. Le Nobel, J. W. and Choufour, J. H., 1959, Development in treating processes for the petroleum industry, in Proceedings of the 5th World Petroleum Congress Sect. IlL 5th WPC Int., New York, pp. 233-243. Lerner, H. and Citarella, V. A., 1991, Improve alkylation efficiency. Hydrocarbon Processing November, 89-94. Levenspiel, O., 1988, Chemical engineering's grand adventure. Chem. Engng Sci. 43, 1427-1435. Manna, L., Sicardi, S., Baldi, G., Van Dierendonck, L. L., Smeets, T. and Stankiewicz, A., 1992, Influence of the unsteady state operation on yield and selectivity of a three phase catalytic reaction. Chem. Engng Sci. 47, 2418-2486. Mares, F., Galle, J. E., Diamond, S. E. and Regina, F_ J_, 1988, Preparation and characterization of a novel catalyst for the hydrogenation of dinitrils to aminonitrils. J. Catal. 112, 145-146. Matros, Y. S., 1985, Unsteady Processes in Catalytic Reactors. Elsevier, Amsterdam. Matsuda, T., Koike, I., Kubo, N. and Kikuchi, E., 1993, Dehydrogenation of isobutane to isobutene in a palladium membrane reactor. Appl. Catal. 96, 3-13. Oelderik, J. M., Sic, S T. and Bode, D., 1989, Progress in the catalysis of the upgrading of petroleum residue_ Appl. Catal., 47, 1-24. Pangarkar, V. G. and Sharma, M. M_, 1974, Consecutive reactions: role of mass transfer effects. Chem. Engng Sci. 29, 561-569. Paquet, D. A. and Ray, W. H., 1994, Tubular reactors for emulsion polymerization: I. Experimental investigation_ A.I.Ch.E.J. 40, 73-87. Poll, I., Krishna, R., Voetter, H. and Van Wechem, H. M. H., 1987, The basis of reactor selection for the Shell Shale retorting process, in Recent Trends in Chemical Reaction Engineering, Vol. II (Edited by B. D. Kulkarni, R. A. Mashelkar and M. M. Sharma). Wiley Eastern Limited, New Delhi. Post, M. F. M., Van't Hoog, A. C., Minderhoud, J. K_ and Sic, S. T. 1989, Diffusion limitations in Fischer Tropsch catalysts. A.I.Ch.E.J. 35, 1107-1114. Ray, W. H., 1968, Periodic operation of polymerization reactors, Ind. Engng Chem. Process, Des. Dev. 7, 422426. Rigopoulos, K., Shu, X_ and Cinar, A., 1988, Forced periodic control of an exothermic CSTR with multiple input oscillations. A.I.Ch.E.J. 34, 2041-2051. Samson, R., Goudriaan, F., Maaskant, O. and Gilmore, T_, 1990, The design and installation of a low temperature catalytic NO, reduction system for fired heaters and boilers. Paper presented at the Fall meeting of the Am. Flame Res., Comm., San Francisco, October 8-10. Saracco, G. and Specchia, V., 1994, Catalytic inorganic membrane reactors: Present experience and future opportunities. Catal. Rev.--Scl. Engng 36, 305-384. Sharma, M. M., 1988, Multiphase reactions in the manufacture of fine chemicals, Chem. Engng Sci. 43, 17401758. Sie, S. T., 1980, Catalyst deactivation by poisoning and pore plugging in petroleum processing, in Studies in Surface Science and Catalysis, Vol. 4: Catalyst Deactivation (Edited by B. Delmon and G. F. Froment), pp. 545-569. Elsevier, Amsterdam. Sie, S. T., 1993, Intraparticle diffusion and reaction kinetics as factors in catalyst particle design. Chem. Engng J. 53, 1-11.

oxidation in maleic anhydride over vanadium phosphate catalysts. Catalysis Today 1, 49-58. Corrigan, T. E. and Garret, J. C., 1953, Kinetics of catalytic cracking of cumene. Chem. Engng Progr. 49, 603-610. Cybulksi, A. and Moulijn, J. A., 1994, Monoliths in heterogeneous catalysis. Carol. Rev.--Sci. Engng 36, 179-270. Dautzenberg, F. M., Naber, J. E. and Van Ginneken, A. J., 1971, Shell's flue gas desulphurization process. Chem. Engng Prog. 67, 86-91. Debling, J. A., Han, G. C., Kuijpers, F., VerBurg, J., Zacca, J. and Ray, W. H., 1994, Dynamic modelling of product grade transitions for olefin polymerization processes_ A.I.Ch.E.J. 40, 506-520. DeGarmo, J. L., Parulekar, V. N. and Pinjala, V., 1992, Consider reactive distillation. Chem. Engng Prog. No. 3 (March), 43-50. Doherty, M. F. and Buzad, G., 1992, Reactive distillation by design. Trans. Inst. Chem. Engrs, Part A 70, 448-458. Douglas, J. M., 1967, Periodic reactor operation. Ind. Engng Chem. Process Des. Dev. 6, 43-48. Eigenberger, G. and Nieken, U., 1988, Catalytic combustion with periodic flow reversal, Chem. Engng Sci. 43, 21092115. Eisenberg, B., Ansell, L. L., Fiato, R. A. and Bauman, R. F., 1994, Advanced gas conversion technology for remote natural gas utilization, Paper presented at the 73rd Annual GPA Convention, March 7-9, 1994, New Orleans, Louisiana. Falconer, J. L., Noble, R. D. and Sperry, D. P., 1993, Catalytic membrane reactors, in Membrane Separations Technology." Principles and Applications (Edited by S. A. Stern and R. D. Noble), Elsevier, Amsterdam. Gavriilidis, A. and Varma, A., 1992, Optimal catalyst activity profiles in pellets: 9. Study of ethylene epoxidation. A.I.Ch.E.J. 38, 291-296. Gavriilidis, A. and Varma, A., 1993, Optimal distribution of catalyst in pellets. Catal. Rev.--Sci. Engng 35, 399-456. Gerrens, H., 1982, How to select polymerization reactors. Part II. Chemtech No. 7, 434-442. Gianetto, A., Pagliolico, S., Rovero, G. and Ruggeri, B., 1990, Theoretical and practical aspects of circulating fluidized bed reactors (CFBRs) for complex chemical systems. Chem. Engng Sci. 45, 2219-2225. Gonzalez, V_ L. and Larder, K., 1984, Batch or continuous processing: a case history. Chemtech No. 10 (October), 607-608. Goudriaan, F., Mesters, C. A. M. M. and Samson, R., 1989, Shell process for low temperature NO, control, in Proceedings of the 1989 EPA/EPRI Joint Symposium on Stationary Combustion NOt Control, Vol. 2, San Francisco, March 1989, pp. 8-39. Hatzimanikatis, V., Lyberatos, G., Pavlous, S. and Svoronos, S. A., 1993, A method for pulsed periodic optimization of chemical reaction systems. Chem. Engng Sci. 48, 789797. Horn, F. J. H. and Lin, R. C., 1967, Periodic processes: a variational approach. Ind. Engng Chem. Process Des. Dev. 6, 21-30. Hsieh, H. P., 1991, Inorganic membrane reactors. Catal. Ret,.--Sci. Engng 33, 1-70. Jansen, E., Jacobs, R. and Krishna, R., 1995, Effect orco- and countercurrent operation on the performance of a tricklebed reactor for the hydration of iso-butene to tert-butyl alcohol. Chem. Engny J. (submitted). Kita, H., Tanaka, K., Okamoto, K. and Yamamoto, M., 1987. The esterification of oleic acid with ethanol accompanied by membrane permeation. Chem. Lett. No. 10, 20532056. Kohn, P. M., 1978, Continuous peroxyester route to make its commercial debut. Chem. Engng No. 16, 88-89. Krishna, R., 1993, Analogies in multiphase reactor hydrodynamics, in Encyclopedia of Fluid Mechanics, Supplement 2, Advances in Multiphase Flow (Edited by N. P. Cheremisinoff), pp. 239-297. Gulf Publishing, Houston.

Strategies for multiphase reactor selection Sie, S. T., 1994, Past, present and future role of microporous catalysts in the petroleum industry, in Advanced Zeolite Science and Applications, Studies in Surface Science and Catalysis, Vol. 85 (Edited by J. C. Jansen, M. St6cker, H. G. Karge and J. Weitkamp). Elsevier, Amsterdam, pp. 587-631. Sie, S. T., Senden, M. M. G. and Van Wechem, H. M. H., 1991, Conversion of natural gas to transportation fuels via the Shell Middle Distillate Synthesis process (SMDS). Catalysis Today 8, 371-394. Seider, W. D., Brengel, D. D., Provost, A. M. and Widagdo, S., 1990. Nonlinear analysis in process design. Why overdesign to avoid complex nonlinearities? Ind. Engng Chem. Res. 29, 805-818. Silveston, P. L., Hudgins, R. R., Bogdashev, S., Vernijakovskaja, N. and Matros, Yu. Sh., 1994, Modelling of a periodically operating packed-bed SO2 oxidation reactor at high conversions. Chem. Engng Sci. 49, 335-341. Sloot, H. J., Versteeg, G. F. and Van Swaaij, W. P. M., 1990, A non-permselective membrane reactor for chemical processes normally requiring strict stoichiometric feed rate of reactants. Chem. Enon # Sci_ 45, 2415-2421. Snijder, E. D., Versteeg, G. F. and Van Swaaij, W. P. M., 1992, Theoretical study on hydrogenation catalysts containing a metal hydride as additional hydrogen supply. Chem. Engng Sci_ 47, 3809-3816. Steinmeyer, D., Cho, T., Efthymiou, G. and Shuler, M. L., 1988, Getting the product out. Chemtech. No. 11 (November), 680-685. Struijk, J., 1992, Catalytic routes to cyclohexene. Ph.D. thesis in Chemical Engineering, Delft University of Technology. Teramoto, M., Fujita, S., Kataoka, M., Hashimoto, K. and Nagata, S., 1970, Effect of bubble size on the selectivity of consecutive gas-liquid reactions. J. Chem. Engng Japan 3, 79-82. Tichy, S., Vasic-Racki, D. and Wandrey, C., 1990, Electrodialysis as an integrated downstream process in amino acid production. Chem. Biochem. Enong Quat, 4, 127-135. Trambouze, P_, Van Landeghem, H. and Wauquier, J.-P., 1988, Chemical Reactors. Design, Engineering and Operation. Editions Technip, Paris. Trambouze, P., 1990, Countercurrent two-phase flow fixed bed catalytic reactors, Chem. Engng Sci 45, 2269-2275. Van den Bussche, K. M., Neophytides, S. N., Zolotarskii, I. A. and Froment, G. F., 1993, Modelling and simulation

4065

of the reversed flow operation of a fixed-bed reactor for methanol synthesis. Chem. Engng Sci. 48, 3335-3345. Van der Ham, A. G. J., Prins, W. and Van Swaaij, W. P. M., 1993, Hydrodynamics of a pilot-plant scale regularly packed circulating fluidized bed, in Fluid-Particle Processes: Fundamentals and Applications (Edited by A. W. Weimer). AJ.Ch.E. Symposium Series, New York, pp. 5372. Vanin, G V., Noskov, A. S., Popova, G_ Ya., Andrushkevich, T. V. and Matros, Yu. Sh., 1993, The industrial plant for unsteady state purification of flue-gases from acrylonitrile and cyanic acid. Catalysis Today 17, 251-260_ Veldsink, J. W., Van Damme, R. M. J., Versteeg, G. F. and Van Swaaij, W. P. M., 1992, A catalytically active membrane reactor for fast, exothermic heterogeneous gas-solid reactions. Chem. Engn0 Sci. 47, 2939-2944. Von Eysmondt, J. and Wandrey, C., 1990, Integrierte Produktaufarbeitung mittels Elektrodialyse bei der mikrobiellen Produktion organischer S/iuren. Chem.-lno. Tech. 62, 134-135. De Vos, R., Hatziantonious, V. and Sch66n, N. H, 1982, The cross-flow catalyst reactor. An alternative for liquid phase hydrogenations. Chem. Engng Sci. 37, 1719-1726. Weisz, P. B., 1980, Molecular shape selective catalysis. Pure and Appl. Chem. 32, 2091-2103. Weisz, P. B. and Swegler, E. W., 1955, Effect ofintra-particle diffusion on the kinetics of catalytic dehydrogenation of cyclohexane. J. Phys. Chem_ 59, 823-826. Westerterp, K. R. and Ptasinski, K. J., 1984, Safe design of cooled tubular reactors for exothermic, multiple reactions; Parallel reactions--II. Chem. Engng Sci. 39, 245-252. Wiederkehr, H., 1988, Examples of process improvements in the fine chemicals industry. Chem. Engng Sci. 41, 17831791. Wismeijer, A. A., Kieboom, A. P. G. and Bekkum, H. van, 1986, Solvent-reactant-support interactions in liquidphase hydrogenation II. Supported metal catalysts modified by a liquid film. Rec. Tray. Chim. Pays-Bays 105. 129136. Wolff, E. H. P., Veenstra, P. and Chewter, L. A., 1995, A novel cross-flow moving bed reactor system for gas-solids contacting. Chem. Engng Sci. (submitted). Ziaka, Z. D., Minet, R. G. and Tsotsis, T. T., 1993. A high temperature catalytic membrane reactor for propane dehydrogenation, d. Membrane Sci. 77, 221-232_

Special Topics

Advantages

SPECIAL TOPICS THE DESIGN AND APPLICATION OF POROUS CATALYSTS

In the first part of this chapter the design of catalysts with porous structures, in particular zeolites and the like, and the additional advantages achieved with porous materials with respect to solid/nonporous catalysts will be highlighted. In the second part a few examples of large-scale industrial application of zeolites will be discussed.

8.1 General advantages


This section deals with general advantages of porous catalysts like zeolites in industrial processing of organic chemicals.

8.1.1 Catalytic Versatility


uniform micropores - potential control of molecular shape selectivity variety of pore size and shape - different product distributions

Table 8.1 Product distribution in Methanol to Olefins (MTO) process over different zeolites pore size in nm .45 .55 catalysts tested erionite ZSM-5 zeolite T chabazite ZK-5 ZSM-34 SAPO-34 typical yields wt% wt% 25 - 55 5 - 25 20 - 50 20 - 40 50 - 95 50 - 80 1 - 15 25 - 60 0.5 - 5 < 0.5

component C2= C3= total olefins C5+ coke

8-1

Special Topics

Advantages

diversity of micropore connectivity unique-/dual-/multifunctional catalytic activity - H+, Pt-Mordenite broad spectrum of catalytic activity by: - isomorphous substitution of Si by Al, Ga, or Fe resulting in Brnsted acidity with different strength in acidity (see Fig. 8.1)

hexane cracking activity Al3+ 1.5 Ga3+

Fe3+ 0.5 0 1 2 M3+ wt%

Figure 8.1 Relative hexane cracking activity of Al3+, Ga3+, and Fe3+. - variety in acid strength upon change of the Si/Al ratio low Si/Al results in weaker acid performance, and more sites available compared to high Si/Al with stronger acid sites but less in number - isomorphous substitution but unsaturated coordination of M3+ resulting in Lewis acidity - isomorphous substitution but saturated coordination of M4+ like Ti4+ resulting in oxidation catalyst performance such as hydroxylation and epoxidation - unsaturated coordination on/in the pore wall of V5+ to obtain oxidation catalyst activity - Mo such as Pt and Pd in the pore intersections - coordination compounds like Fe-phtalocyanine in the pore intersections to obtain oxidation catalyst

an

ability to sorb and concentrate small molecules post-synthetic modification - ion exchange - de-alumination resulting in mesoporous structure. This is a destructive method (see Fig. 8.2). More elegant and better controllable is a constructive method in which meso pores are created in direct synthesis (see Fig 8.3, Table 8.2).

8-2

Special Topics

Advantages

de-Al

HY NaOH 800 oC 4 h. pore

USY

pore diameter (nm)

zeolite
HY USY

pores in nm (vol%) micro meso macro


.9 (100) .9 (60) 2 (25) 6-20 (15)

Figure 8.2 Destructive method to obtain mesopores in microporous material.

dry precursor gel

aggregate of crystallites
Figure 8.3 Constructive method to obtain mesopores in microporous material.

8-3

Special Topics Table 8.2 Pore data on zeolite Beta (to be published, 1998) micro (cm3/g) meso(cm3/g) Si/Al BET(m2/g) 400 450 .19 .04 50 500 .18 .3 25 550 .16 .4 12.5 565 .15 .47 10 574 .13 .44 8 563 .11 .41 7 611 .10 .48

Advantages

8.1.2 Process Engineering Flexibility


Porous catalysts like zeolites are very flexible from a process engineering viewpoint: wide selection of reactor designs - fixed, fluid bed - CSTR, continuous slurry - structured bed - Sulzer packing - stainless steel sponge - ceramic monolith easy separation from liquids higher operational temperature than H2SO4 and ion-exchanged resins capability for multi-step conversion in single reactor zone (see Fig. 8.4)

methanol dimethyl ether olefins by disproportionation and polymerization H- ZSM-5 low mol. paraffins aromatics
Figure 8.4 Multi-step conversion of methanol into aromatics over H-ZSM-5. use of binders for attrition-resistance high internal surface area (economic throughputs & reactor size) High internal surface depends on: - Dp (see Fig. 8.5) - minimum mesopore contribution These factors also determine the compromise between accessibility and selectivity (Fig. 8.6).

8-4

Special Topics

Advantages

surface area (m2) * 10-0

10-5
o

* External (practice) Internal

10-10 External (theory)

10

-15

10-20

o*

10-1 10-2 10-3 10-4 10-5 10-6 10-7

Dp ( m)

Figure 8.5 Influence of Dp on the contribution of external surface catalytic activity compared to internal catalytic activity.

selectivity accessibility

Figure 8.6 The balance between accessibility and selectivity.

8.1.3 Environmental Aspects


Solid catalysts of the zeolitic type are also advantageous from in environmental respect: regenerable/ reusable Table 8.3 Maximum allowable temperatures of some zeolites zeolite Si/Al Max. Temp. oC A 1 600 Y 6 800 silicalite-1 infinite 1100 8-5

Special Topics

Advantages

Under dry conditions the stability of the framework is more preserved than under wet (steaming) conditions. safer handling than liquid mineral or solid Lewis acids non-toxic more selective/ tailor-made catalyst for specific reaction, thus less by-product

8.2 Optimization/Minimization of Factors affecting Conversion and Selectivity


The following factors affecting the conversion and selectivity of the reaction must be optimized/ minimized in the catalyst design.

8.2.1 Catalyst Structure


Si/Al ratio (see Fig. 8.7)
conversion

no acidity

weaker sites

stronger sites

7---------10

Si/Al

Figure 8.7 Influence of Si/Al ratio (acidity) on conversion. impurities - Ni and V in crude oil - water thermochemical activation history crystal size and morphology external surface activity by-product formation binder effects

8-6

Special Topics

Optimization

8.2.2 Reaction System


thermodynamics reactants and their ratios reactor design/ fluid/ catalyst contacting solvent effects fluid flow and mass transfer (see Fig. 8.8)

reactants macropores between bed particles catalyst bed

mesopores in pressed zeolite material products

micropores in zeolite particle


Figure 8.8 The tortuous path, comprising macro-, meso- and micropores in a particular distribution, which the reactant has to pass in the reactor.

molecular diffusion heat transfer adsorption/ desorption phenomena catalyst deactivation presence of poisons and inhibitors

Because of the multiparameter design of the porous catalyst the study of the product distribution can lead to dramatic differences. In case the product distribution is studied at one LHSV confusion in the interpretation of the catalyst performance might occur since at different LHSVs the product distribution is very different as depicted in Fig. 8.9.

8-7

Special Topics

Optimization

product distribution (wt%) 70 methanol 60 50 40 paraffins 30 aromatics 20 10 C2 - C5 olefins 0 10-4 10-3 10-2 10-1 1 10 Space time (1/LHSV)
Figure 8.9 Differences in product distribution, in the conversion of methanol to hydrocarbons with H-ZSM-5 at 630 K, as a function of the LHSV.

water dimethyl ether

8-8

Special Topics

Monolith reactors

9 SPECIAL TOPICS MONOLITH REACTORS 9.1 Why monoliths


In the previous chapters we have used mass- and heat balances to describe mass and heat transport limitations in catalytic reactors. In this process we have derived dimensionless numbers, such as the Thiele modulus and the Carberry number Ca. All these numbers indicate that in most cases (i.e. when one doesn't want transport limitations) it would be best to have the highest possible specific surface area, i.e. the shortest diffusional lengths. In other words, it is often favorable to have small particles. On the other hand, many of our calculations have focussed on particles sizes where mass or heat transfer limitations occur. The reason we often have such larger particles is that in reactor design also the momentum balance is considered. Solving the momentum balances for a packed bed leads to equations such as the Ergun equation for pressure drop in packed beds. These pressure drop correlations have in common that for increasing surface area the pressure drop will rise. In some applications high pressure drop is not a large issue. For instance, in a slow heterogeneously catalyzed process a large residence time will be required. The resulting low superficial velocities are not likely to pose a large pressure drop problem. On the other hand, fast reacting systems with high superficial velocities benefit most from high surface areas. Here pressure drop may cause a problem. It can be argued that the high pressure drop is not only caused by the size of the particles, but also by the irregularity of the packing, i.e. the random way in which they are ordered geometrically in the column. For instance, the friction factor for a regular structure like a pipe is significantly lower than for a packed bed. Moreover, there are more difficult phenomena in packed column design that are caused by this irregularity. - Maldistribution and incomplete wetting. Maldistribution in three phase reactors occurs when some areas in the column have more liquid holdup (and thus less gas holdup) than other areas. In a severe case of maldistribution, some catalyst particles are not wetted (completely) by the liquid. These particles do not convert as much as completely wetted particles. Wall effects are also a form of maldistribution. - Fouling and Attrition. Erosion can occur when particles are moving with respect to each other. All these considerations have triggered the development of structured catalysts. There are several types of structured catalysts, such as the Sulzer packing, the Three Levels of Porosity Reactor and the Bead String Reactor (see Figure 9.1). For more information on these types of reactors, as well as monolith reactors, the interested reader is referred to [1]. All these catalytic reactor concepts have in common that a higher level of structuring is found - relative to a packed bed - to overcome some of the aforementioned limitations. In some cases, the distinction between catalyst and reactor is lost, and the entire internal structure of the column is optimized to enhance mass and heat transfer and to lower energy consumption (pressure drop). The regular structure itself is made up entirely of catalytically active material or has a washcoat on its surface containing the active material. One type of common structured reactor, the monolith reactor, will be treated in more detail in this chapter. The monolith reactor was developed for the cleaning of exhaust gases from combustion 9-1

Special Topics

Monolith reactors

processes, both in cars and large power plants. For these processes, the monolith reactor offers an irresistible combination of low pressure drop (two to three orders of magnitude lower than in packed beds) and high surface area. Currently, an effort is being made, amongst others by Delft University, to use the monolith concept in other areas, including three phase processes and counter current flow operations.

(a) Sulzer packing

(b) Cross-flow reactor

(c) Monolith

(d) Parallel passage reactor

(e) Three levels of porosity

(f) Bead-string reactor

Figure 9.1 Structured catalytic reactors.

9.2 Geometry
The word monolith comes from Greek, and is interpreted here as made from one piece (of stone). Most monolith reactors consist of one piece of ceramic material. This ceramic block contains a large number of parallel channels extending over the entire length of the block, separated by thin walls. The channels usually have a small diameter, resulting in large specific surface areas, combined with the pressure drop of a (small) pipe. There are no open passages from one channel to neighboring ones, so the quality of distribution at the entrance of the block is maintained throughout the entire length of the block. The channels themselves can be square, hexagonal or triangular, etc. Another way of producing monoliths is to use corrugated plates of metals. Ceramic monoliths are extruded to a (currently) maximum length of one meter and a diameter of approximately 50 cm. Larger volumes of monolith structures are obtained by stacking smaller building blocks. Various channel geometries exist: square, hexagonal, triangular, finned channels, but also channel geometries produced by wrapping up flat sheets with corrugated sheets in between (see Figure 9.2). The monolith themselves have a square, oval, racetrack (cf. Indy 500 racetrack layout), triangular or circular geometry, depending on the application of the structure. 9-2

Special Topics

Monolith reactors

Figure 9.2 Monolith shapes.

9.3 Engineering Correlations - Single Phase


Engineering correlations for pipes are readily available in the literature. These can be used for monoliths if the feed of each channel is the same. The monolith can then be regarded as a large number of (parallel) tube reactors. For instance, the Fanning equation for pressure drop in pipes [2]

P =

L1 2 4 f d 2 16 f = laminar flow, circular channels Re 14.9 f = laminar flow, square channels Re

(1) (2) (3)

is used successfully to model the pressure drop in monoliths. Off course this simple equation can be improved to include some other phenomena. In the entrance region, the flow pattern develops from an initially flat profile to a parabolic one. (A similar phenomenon will be observed with mass and heat transfer later). This consumes energy, and is accounted for by a term with a vanishing contribution for long channels. Also, the in- and outlet effects can be included in a fashion, similar to the modeling of pipes

d 64 4f = 1 + 0.0445 Re Re L

0.5

+ Kw

d L

(4)

For mass and heat transfer, we can use the Chilton-Colburn analogies: for mass and heat transfer the equations will have the same form

L Nu = f Re, Pr , d

(5) 9-3

Special Topics

Monolith reactors

L Sh = f Re, Sc, d
In Perry and Green [3] we find for heat transfer for round tubes
0.8 L 0.17 RePr d Nu = 3.66 0.5 L RePr d 1 + 0.117

(6)

(7)

which closely resembles the Sherwood mass transfer correlation used for monoliths

d Sh = 3.661 + 0.095ReSc L

0.45

(8)

The expressions for Sh and Nu contain a term for infinitely long channels and a contribution for entrance effects. The length of the entrance effects depends on the flow properties (Re) and the fluid properties (Re, Sc or Pr). For a typical Re value of 100, a Sc value of 1 and 1000 for gases and liquids, respectively, and a channel diameter of 1mm the entrance length is 0.1 and 10 m, respectively. (Since mass transfer is enhanced by the entrance effects, the fact that monoliths are only available up to 1 m in length is less of a problem, if the entrance effects are repeated every once in a while.) So, we can conclude that well established solutions for the mass, heat and momentum balances can be used to create a reactor model for monolith reactors with single phase feed. And, indeed, these models have been applied successfully.

9.4 Example: DeNOx


Today, world-wide legislation (with local deviations) requires that the exhaust gases from power plants are cleaned, allowing a maximum of 100 ppm NOx. In order to comply with these requirements, one can distinguish primary measures, in which the combustion process itself is optimized for low NOx levels, and secondary measures, in which the effluent gas is cleaned. By primary measures alone, NOx levels of 1000 ppm are obtained, so the secondary measures are also necessary. In the so-called DeNOx process, ammonia is fed to the effluent gas to reduce the oxides to nitrogen. In this process, SO2 is also oxidized to SO3, which forms ammonium hydrates that damage process equipment and plug the catalyst pores. The chemistry of the process can be written as follows 4 NO + 4 NH3 + 2 O2 4 N2 + 6 H2O 2 SO2 + O2 2 SO3 9-4 desired undesired

Special Topics

Monolith reactors

QUESTION: The reaction of NO is fast relative to the reaction of SO2. Derive how the selectivity depends on the surface area and the catalyst volume. The fact that selectivity is a linear function of specific surface area calls for small particles. Monoliths with thin channel walls are also suitable. A major reason for choosing monoliths in this process is the low pressure drop. Throughputs are very high (ug,s > 3 m/s, even if a very large cross sectional area is chosen), and the cost of a compressor is large. Another issue is that the feed contains many dust particles from the combustion stage. This dust would certainly plug a packed bed. Figure 9.3 shows a typical design of a monolith reactor for the DeNOx process.

distribution layer

catalyst layers

module

honeycomb monolith

Figure 9.3 Typical design of a monolith reactor for DeNOx. The reactor shown in Figure 9.3 is filled with monolithic modules that are stacked to fill the reactor. At the entrance of the reactor ammonia is fed. In this particular configuration, the position of the inlet at the side will lead to maldistribution. The flow rate and amount of dust will be highest at the side opposite of the entrance. This would cause ammonia to locally slip through and - even using monoliths - severe local erosion of the catalytic material. To counteract maldistribution, a special noncatalytic layer is mounted at the top to redistribute the gas. Pore-mouth poisoning, e.g. by arsenic oxide, and erosion by dust particles tends to favor fully active walls over washcoats.

9.5 Example: Car Converters


In 1970, the US Clean Air Act called for a reduction of polluting gases from car exhaust by 90% in 5 years. Today, the reduction of exhaust pollution still goes on. In the seventies, a catalytic system was introduced to convert gases like NOx and CO to less polluting gases in a packed bed-like 9-5

Special Topics

Monolith reactors

configuration. There are several reasons for preferring monoliths to packed beds in car converters. The most important ones are: Pressure drop. At the very high temperature of the exhaust gas, the reaction is very fast, and high surface areas are needed to meet the pollution requirements. In a packed bed, the pressure drop is so high that the engine performance drops considerably. Car manufacturers are very sensitive to reduction of engine power. Catalyst mass. A larger eggshell catalyst could be used to obtain a reasonable pressure drop, but the reactor would have to become to long, and more importantly, too heavy. Weight reduction is a constant consideration in car design, but there is a more important reason for weight reduction. Cars do not - in contrast to petrochemical reactors - operate at steady state. During start-up, the catalyst is still cold and the system is inactive. Even today, using monoliths with very thin walls, most of the emissions occur during the inevitably cold start-up of the car. Therefore, it is vital that the catalyst heats up fast. The fast heat up of monolith car converters has caused thermal stress problems. While the entrance of the monolith is already hot, the exit side is still rather cold. Due to thermal expansion the structure is likely to break. For this reason the ceramic material cordierite is used, which has a very small thermal expansion coefficient.

9.6 Three-Phase Flow


Due to the successful application of monolith reactors for gas-phase reactions, the use of the monolith reactor has been extended to three-phase processes. Three-phase processes require high surface-tovolume ratios, because diffusion through the liquid layer is often slower than reaction. Hydrodynamics play an important role in contacting of the phases. Conventional three-phase reactors are trickle-bed reactors - packed beds through which the gas and liquid move downwards - and slurry reactors - stirred liquid tanks, with gas and fine catalyst particles in suspension. The hydrodynamics of a gas and liquid, moving through a monolith channel closely resembles the hydrodynamics in two-phase pipe flow. The main difference is that due to the small channel sizes, surface tension forces play a large role, whereas they are usually ignored in pipe flow. The flow patterns found in monoliths are summarized in Figure 9.4 for increasing gas flow rates.

Figure 9.4 Flow patterns in channels. 9-6

Special Topics

Monolith reactors

At very low flow rates, the bubbles are smaller than the channel diameter, and bubble flow is observed. For higher gas flow rates, the gas bubbles are larger than the channel diameter, and the twophase system moves through the reactor as a train of separate gas and liquid slugs. This flow pattern is called Taylor flow, and is of most interest to monoliths. For even higher flow rates, the pattern becomes less well defined and will eventually go to annular flow. In Taylor flow, the catalyst wall is separated from the gas bubble by a very thin film. The thin film gives rise to high concentration gradients in the case of a fast reaction at the wall, and thus high fluxes. The three-phase monolith can be operated even at zero pressure drop. In this case, downflow in the channel is driven by gravity alone. In this case, the reactor design depicted in Figure 9.5 can be used. Liquid is fed at the top of the reactor and enters the monolith blocks through a distributor. The gas is sucked in by the falling liquid plugs and recirculated inside the column, eliminating the need for a compressor. The liquid is collected at the bottom and can be recirculated.

Figure 9.5. Three phase monolith reactor.

9.7 Engineering Correlations Three-Phase Flow


Modeling of pressure drop (if any) in a monolith is very similar to modeling of single-phase pressure drop. Due to the capillary forces the gas bubbles are very rigid, and each liquid plug can be seen as single-phase entrance flow into a pipe. Ignoring the contribution of the gas plug to the pressure drop, which can be done safely, the Fanning equation with correction for entrance effects can be used (Eqn. 9).

4f =

64 d 1 + 0.0445 Re Re L

0.5

(9)

A plot of experimentally obtained friction factors versus Re, and assuming a constant plug length indeed gives a surprisingly good fit (see Figure 9.6). 9-7

Special Topics

Monolith reactors

Figure 9.6 three phase pressure drop. The success of the pressure drop model suggests applying the same principle to mass transfer: regard each liquid plug as an entrance region for developing mass transfer and use the single-phase model, replacing channel length with liquid plug length. For mass transfer from the liquid plug to the catalytic wall (or vice versa), this model has indeed been successfully applied. This mass transfer model would also implicate that a higher conversion is obtained in a liquid-solid reaction, if inert gas is used in a monolith. This has indeed been experimentally demonstrated. For mass transfer from the gas plug through the film to the wall we can make a first estimation by assuming a flat profile in the film, zero concentration at the wall (fast reaction), and saturation at the liquid interface. In this case we only need the film thickness as a function of flow and fluid properties.

Ca =


= 0.18 1 exp 3.1Ca 0.54

(10)

film
d channel

)]

(11)

Here, the surface tension is expressed as the dimensionless ratio Ca of viscous forces to capillary forces. The mass transfer coefficient is now estimated by

k GS =

film

(12)

A complete model for mass transfer from gas phase to solid wall would also have to describe mass transfer from the gas plug to the liquid plug, and is beyond the scope of this introduction.

9.8 Example: HDS


Desulfurization of vapor gas oil fractions is an important process in the petrochemical industry. One of the main problems is that this process is inhibited by its product H2S. The lumped (all sulfur compounds treated as one) reaction equation and kinetics are given by: 9-8

Special Topics
2 cS cH2

Monolith reactors

org=S +

H2 org + H2S

r = k

1 + Kc H 2 S

(13)

In this case, it is favorable to operate the reactor in counter-current mode. The bulk of the product is then formed in the first part of the reactor and removed immediately, reducing the inhibiting effect. (For an in depth discussion of HDS and counter-current flow, the reader is referred to Krishna and Sie [4]). A conventional packed bed cannot be operated in counter-current mode because the gas, flowing upward, entrains the down flowing liquid. This phenomenon is termed flooding. Several (structured) reactor concepts have been developed to allow counter-current operation, amongst which a special type of monolith, the internally finned monolith reactor (also see Figure 9.2). The (relatively) small amount of liquid runs along the walls of the channel. This falling film is stabilized by the fins. The gas moves upward through the center of the channel. So, in essence, this is a falling film reactor with a high area available for the falling film. It has been shown that (using a specially designed outlet geometry) flooding limits are good and mass transfer is acceptable.

Literature
1. J.A. Moulijn, A. and Cybulski (Eds.),Structured catalysts and reactors, Marcel Dekker (1998). 2. R.B. Bird, W.E. Stewart and E. Lightfoot, Transport Phenomena, John Wiley & sons (1960). 3. R. Perry and D. Green, Perrys chemical engineers handbook, McGraw-Hill, New York, 6th ed. (1984) 4. R. Krishna, and S.T. Sie, Strategies for multiphase reactor selection, Chem. Eng. Sci. 49, 40294065 (1994).

9-9

You might also like