You are on page 1of 322
NRO Re Rae Aantal eterogeneous ~ World Scientific | CONTENTS | CONCEPTS INCATALYSIS. 5 sd | L_References 12 CHAPTER 2: THEORIES OF CHEMISORPTION,» ss 2.1_Introduction u 2.2 Theoretical Intermezzo 1. Background to the Extended-Hiickel Method and other Semi-Empirical Methods. . . B 2.3 Elementary Theory of Chemisorption. One-Dimensional Models. __44 2.3.1 Elementary free-electron gas models. 46 2.3.2 Tight-binding calculations, all atoms equal; no hybridization 51 2.33 One-dimensional models, hybridization... _ 63 | 2.3.3.1 Physical background. 63 2.3.3.2 Elementary quantum chemistry of the Shockley surface state 68 | 2.4 Theoretical Intermezzo 2. Introduction to the Green’s Function Method, Application to One-Dimensional Surfaces Green’s function calculation of local density of states o | open chain . 78 242 ‘The resolvent method. Application to Shodiley surface state 80 2.5 Elementary Quantum Chemistry of Chemisorption to Metal Surfaces. 76 Tamm Surface States, Surface Molecule Limit : . 85 2.5.1 __A model for three-dimensional chemisorption. Bethe lattice approximation; the concept of group orbitals... _ 8 | 2.6 Theoretical Intermezzo 3. Formal Chemisorption Theory... 102 | 2.6.1 The expressions for the bond energy. 102 | 2.6.2 Relationship with HOMO-LUMO theory ‘and frobtier aia | theory i - 113 | 2.6.3 Relationship with bond-order caleul ‘ions i 4s gS | 2.1 The Effect of E -E I ; | 211 ; icted HL Ri approximation... ss ss 2.7.2 ‘The unrestricted Hartree-Fock solution of the Newns-Anderson chemisorption model. 12 2.73 ‘The effective one-center electron-clectron repulsion integral Time-dependent charge fluctuations of the chemisorbed state _124 2.7.3.1 The physical picture. sw 108 % CONTENTS 2.7.3.2 Theoretical Intermezzo 4; tunnel time and electron m acoming effects. Quantum chemistry of Uers - - 126 2.8 Application of Bethe Lattice Method to CO and H Chemisorption . _140 281 TI 7 f a 1 surh 140 2.8.2 Atop and bridge coordination of COandH . ws. 148 283 Coordination of CO to different £. on i 56 2.9 Free-Electron Approaches to Chemisorption. Some First-Principle Calculations 0. CHAPTER 3: MOLECULAR BASIS OF METAL CATALYSIS Ci‘ 3.1__Introduction 178 3.2 fees in Chemisorption _. 179 3.2. Adsorption energy as a Tunction of coverage; surface reconstruction... 3.3 Quantum Chemistry of Surface Dissociation and Association Reactions * 2 F 90 3.4 Shustorovich’s Empirical Approach to Surface Association and Dissociation Reactions . . » + 204 3.4.1 | The Shustorovich Bond Order ‘Conservation ‘method . 204 3.4.2 Derivation of the formula for the heat of molecular sbceyhion: 206 3.4.3 Derivation of the formula representing the activation energy for dissociation of a chemisorbed molecule se ee 208 3.5 Catalysis of Synthesis Gas Conversion . . 3 . . 3...) 21 3.5.1 General features. » 21 3.5.2 Quantum chemistry of CO chemisorption and dissociation 213, 3.5.3 CH, fragment chemisorption and reactivity... s 284 3.6 Hydrocarbon Conversion Catalysis... ess 3.6.1 General background sw sO 3.6.2 Theory of surface ligand effects. . 246 3.6.3 Quantum chemistry of cation-induced changes in chemisorption to metal clusters . . . =. 0. 3.7 Selective Oxidation Catalysis hg) 87 3.7.1 General background . 267 3.72 Quantum chemistry of chemisorption of the O, molecule tosilver . . 270 3.7.3 Epoxidation of ethylene; the electronic basis of the role of subsurface oxygen...» . ss ss 7h conrents xi 3.8 References 0. . . ss CHAPTER 4: | | | | | IDE SURFACE: 281 4.1 General Physicochemical Properties of the Tonogente Surface. Pauling’s Valency Concept... ee 4.1.1 Surface relaxation effects __. -__281 4.1.2 _ The Madelung potential of the surface layer; Pauling’s valencies 285 4.1.2.1 Ton charge excess of the dense TiO, surface. _. 291 4.1.2.2 Elementary theory of surface Bronsted and Lewis acidity and basicity. Implications for catalyst preparation and acidcatalysis . ww we we es 04 2 ThE ic 8 f Transition-Metal Oxide Surt ans 4.2.1 Ligand-field theoretical considerations, the surface dangling bond analog. . 303 j 7 | 4.22 Valence electron stmctore of the rutile surface a1 | 4.3 Quantum-Chemical Basis of Bronsted Acidity in Zeolites... __316 | 4.4 The Quantum-Chemical Basis of Heterogeneous Oxide Catalysis. 326 4.5 References 2. 2. . ee eee eee BBL | 5.1 Introduction a | 5.2__Gibbs’ Rule 336 | 5.2.1 Surface composition ofalloys . . =. =. =. =. —- ‘34 5.2.2 Statistical- mechanical theory ‘of surface enrichment in alloys 343 | 5.2.2.1 Regular solution model for a two-component ally .. . 343 5.2.3 Application of the cluster method “_ * 2 5 = wee | 5.3 Lattice Stability of Zeolites =... . . . ss 852 5.3.1 Background. te 5.32 The lattice energy of zeolites... ~—~—=s=~SC~—ws~C“CS~Csi‘“‘<=‘“ 1-10 mbar). Hydrocarbon conversion catalysts may become covered with hydrogen or a carbonaceous layer, ot may become converted to a carbide as is the case for Fe. In view of this it is remarkable that for selected reactions and conditions, kinetic data determined at vacuum conditions can be used to predict the kinetic behaviour of the catalyst over a pressure gap of ten orders of magnitude. This has been found for the ammonia synthesis reaction!) as well as for the reaction of CO and NOM). Apparently the latter reactions can be classified as condition invariant, whereas the former example of ethylene epoxidation is a condition variant reaction. Bonding between adsorbate and reactant and especially the rearrangement of adsorbate molecules to chemisorbed reaction intermediates is usually accompanied by a change in the formal valency of the surface atom to which the reaction inter- mediate is coordinated. This will be discussed for the metal catalysts in chapter 3 and for the oxidic catalysts in chapter 4. Solid acids such as zeolites and clays can be synthesized as Brgnsted acids, stable at temperatures up to 700 or 800 °C. They catalyze reactions of the Bronsted acidity type that are known from classical carbo- nium or carbenium ion organic chemistry. The presence of micropores in the zeolite may change the selectivity of reaction due to spatial restraints. Also Lewis acid cat- alysts are known. A familiar example is the epoxidation reaction of an alkene with hydrogen peroxide or a hydroperoxide, catalyzed by silica or silica-zeolites contain- ing small amounts of Ti. The large Ti ions substitute Si‘* in tetrahedral positions. Exposed to the peroxide, the Ti‘* polarizes the electrons in the -O-OH group, en- hancing the electrophilic nature of the OH group that is inserted into the alkene bord. CONCEPTS IN CATALYSIS Table 1.1. Examples of important catalytic materials and reactions. ~ SiO2/TiO2 (Lewis acids) Sulfides - NiS/MoS2, CoS/MoS» Reduci i ~ Bi03/MoOs3 V203/TiO2 Chlorides ~ TiCls Catalyst Reactions Group-VIII/IB metals ~ Pt, Ni dehydrogenation, isomerization Pt, Co, Ag oxidation - Fe, Rh, Co synthesis gas conversion NO reduction Solid acids ~ zeolites, clays hydrocarbon cracking (Bronsted acids) methanol to aromatics alkylation isomerization shape-selective reactions epoxidation hydrodesulfurization hydrodenitrogenation selective oxidation propylene to acrolein NO reduction polymerization 5 In the presence of a vapor phase with sulfur-containing molecules, metal surfaces become sulfidized. A familiar class of sulfidic catalysts are MoS; and WS. They are usually promoted with NiS or CoS. Such catalysts are very active in desulfurization and hydrodenitrogenation catalysis. 6 CHAPTER However, other applications are also known. MoS catalysts can be used to pro- duce higher alcohols from synthesis gas and for the water-gas shift reaction. The elementary reaction steps on the sulfide metal center proceed in a way analogous to those in organometallic complexes. On the edges of the sulfide the metal ion ligand positions are partially occupied by reacting molecules. Catalysis occurs by oxidative addition, insertion or reductive elimination steps. The electronic features controlling these events are very similar to those known from coordination chemistry]. The situation is more complex for the reducible oxides, represented by the BiyO3/MoOs compounds. They are used for the selective oxidation of propylene to acrolein or as ammoxidation catalyst. According to the Mars-van Krevelenl® mechanism the Mo ion is the cation interacting with the organic molecule and Bi2O3 mainly serves to conduct oxygen atoms to the catalytically active site. A recent development is the use of mixed oxides of V203/TiO2, notably for the conversion of NO with NH to No in stack-gas. Another important catalyst is based on the metal chorides. TiCls dispersed on MgCl, is used as an olefin polymerization catalyst. In similar fashion to the sulfide case, catalysis occurs by a coordination-type mechanism. Ethylene or propylene coordinates to an unoccupied ligand position of Ti and inserts into the growing hydrocarbon chain positioned on another ligand position of TiCls. MgClz acts as support to TiCl3. There exist many other examples of coordination-type mechanisms occurring at the catalytically active sites of heterogeneous catalysts. We will just mention the Cr203/SiOz catalyst used for ethylene polymerization, and the MoO03/SiO2 or Re2Os/SiO> catalyst used for the olefin metathesis reaction. Most of the catalytic systems discussed result from accidental discovery or pur- poseful development. Initially, both synthesis chemistry and reaction mechanism used to be little understood and optimization studies were often highly empirical. At present a considerable change is taking place in the toolkit of the practical cat- alytic chemist or chemical engineer. In large industrial laboratories the use of X-ray diffraction, Solid State NMR, Infrared Spectroscopy and X-ray spectroscopy is be- coming state of the art. This adds to the more ”classical” analytical tools such as Electron Microscopy and structural characterization techniques based on adsorp- tion and desorption of gases. Such changes illustrate the impact of spectroscopic techniques and more indirectly of theoretical chemistry on practical catalysis. Catalysts are usually multicomponent systems in view of the need to stabilize the active phase, to suppress nonselective sites and to optimize surface area. For instance, metal particles are usually dispersed on inert high-surface oxidic supports. Promotors or additives may be added to stabilize the particles or to poison nonselec- tive sites (e.g. Sin catalytic reforming). There is not a single technique that enables a complete characterization of such a system, but a combination of several spectro- scopic techniques") may be very succesful to identify the different compound phases present on a catalyst. Apart from developments with a direct impact on practical catalysis, advances in spectroscopy as well as theory have significantly benefited catalysis science, mostly through the study of well-defined model systems. The sur- face science approach to catalysis as pioneered by Somorjail') and Ertl!) highlights these efforts. The use of metal single crystal faces in combination with Ultra High CONCEPTS IN CATALYSIS. T Vacuum techniques makes the study of well-defined surfaces possible. In addition, surface spectroscopic techniques can be used to identify chemisorbed species and their changes as a function of temperature or upon contact with another reactant gas. This approach has been very fruitful and has generated a significant amount of detailed information on the behaviour of chemisorbed molecules indispensable for the theoretical development of catalysis. The study of chemisorbed molecules is of the essence for catalysis. Catalytic events consist of elementary reactions on the catalyst surface. Chemical bonds are formed between the surface atoms and the adsorbing molecule. These interactions cause rupture of chemical bonds within the adsorbing molecule and formation of new chemical bonds between the fragments. Related to the Surface Science approach to catalysis is the use of molecular beam techniques'®l, With Mass Spectroscopy and other detection techniques the reactivity of small metal particles effusing into a vacuum is studied. Again reaction conditions are well defined, but the temperature is close to zero Kelvin. This approach enables the study of chemical reactivity as a function of particle size. Knowledge on the reactivity of isolated particles is relevant especially for theoretical studies since these data refer to well-defined particles without interaction with a support. Theoretical studies can be conveniently done with isolated small clusters. As with surface science studies, molecular beam experiments are fundamental to the study of elementary reactions as a function of surface or particle structure. We will frequently refer to data collected by these approaches. Zeolites form another class of materials useful for fundamental studies(!“. As mentioned earlier, zeolites are microporous silica-aluminates with micropores of di- mensions comparable to organic molecules. The materials are unique, because these micropores are determined by the three-dimensional crystallographic structure of the material and catalytic events occur at the interphase of zeolite micropore and zeolite lattice. As a result the catalytically active sites are well defined. Zeolites are used in practice in the acidic form or promoted with metal or sulfide particles. High Resolution Electron Microscopy, Neutron Diffraction and Solid State NMR are tech- niques that are applied for structural characterization and to study the behaviour of chemisorbed molecules. Theoretical methods are required to derive structural information from spectro- scopic data, which usually concern measurements of electronic features. Because of the availability of large and efficient computer power and the current state of the art of theoretical chemistry, electronic structure calculations on model systems of relevance to experimental studies can be made. In addition, the catalytic chemist needs insight into the factors that determine the transition-state potential energy surface of reacting molecules. Also methods are needed to predict the geometry of the adsorption site as a function of metal surface composition or charge distribution in the zeolite. These methods will be extensively discussed in the next chapters. For mechanistic and structural studies of practical catalysts, the possibility to 8 CHAPTER do in-situ spectroscopy of the working catalyst is important. Catalytic reaction intermediates as well as active catalyst phases are often unstable with respect to the noncatalytic condition. Techniques such as Infrared Spectroscopy using electro- magnetic radiation are especially valuable so that high pressure-gas conditions can be studied. For mechanistic studies pulse-type experiments with isotope-labelled compounds are often very succesfully employed. A recent interesting technique is the TAPl!1] (Temporal Analysis of Products) reactor system that enables massspectrometric identification of reaction interme- diates at a milliseconds timescale. The mechanistic knowledge generated by such techniques is used not only to unravel details of elementary reaction steps, but also to study catalyst synthesis ”in-situ” and as a function of catalyst activation. Usu- ally an active catalyst is made by reaction of a support with an active catalytic site precursor and activation procedures are applied to produce a catalyst with the required performance. Two developments are of importance that enable the ”molec- ular catalyst engineering approach” to be realized. Firstly there are the state of the art and the rapid extension of organometallic chemistry. The arsenal of synthetic methods can be used to produce a rich variety of complexes involving similar metal centers. Clusters can be synthesized containing also oxygen or sulfur. Secondly we have the molecularization of colloid chemistry. As a result, the properties of OH groups adsorbed on a support surface, its basic or Bronsted acid properties, aad also Lewis acid or basic sites become understood on a molecular level. By proper variation of the conditions for catalyst support preparation, its reactivity with re- spect to active catalytic site precursors may be controlled. This enables production of more uniform and more selective catalysts. Four major concepts appear to be important to molecular heterogeneous catal- -lysis, They are: 1. Chemisorption and catalytically active site 2. Structure and reactivity relationships 3. Long-range versus short-range electronic and electrostatic effects 4. Surface and phase stability ad 1, CHEMISORPTION Earlier we described the catalytic reaction as a series of consecutive steps at the surface, in which adsorbate and adsorbate-surface bonds are formed and/or broken on the reaction path towards the product molecule, The forces between surface atoms and adsorbate atoms responsible for rearrangement of the chemical bond are similar to those responsible for strong adsorption (Egg > 10 kcal/mol). The adsorption process dominated by such interaction is called chemisorption. Even on a single crystal metal surface, several adsorption modes are conceivable and for dissociation of a diatomic molecule many different reaction paths can be envisioned. However, usually only one particular surface atom configuration is preferred to lead to the idea of catalytic active site. If catalysis of a molecule is studied that has several reaction possibilities, some desirable and others not, a selective reaction usually requires a particular surface atom composition and rearrangement. CONCEPTS IN CATALYSIS. 9 ad 2. REACTIVITY RELATIONSHIP The concept of chemisorption as a function of surface atom rearrangement implies a relation between substrate reactivity and catalytic site rearrangement. Especially surface-science-type experiments have demonstrated the dependence of catalytic ac- tivity on single crystal surface exposed. Usually surfaces containing atoms of increas- ing coordinative unsaturation appear more reactive. Molecular beam experiments have shown dramatic changes in chemical reactivity as a function of particle size-An optimum in reactivity is generally found for particles containing 6-10 metal atoms. Increasing particle size at first results in a decrease in reactivity, which again slowly increases for larger particle sizes. These differences relate to changes in electronic structure, especially for the smaller metal particles, as well as changes in exposed surface geometry. The latter becomes very apparent from studies involving alloys of catalytically active and nonactive metal atoms or coadsorption with catalytically inert adsorbate atoms. It appears that some reactions (e.g. dissociation) require coordination of the reactive molecules to a surface site with a minimum number of neighbouring reactive surface metal atoms in order for the reaction to proceed. Such reactions become suppressed if the concentration of these reactive surface atom en- sembles decreases by dilution with nonreactive atoms. The earlier discussed concept of shape selectivity observed for zeolites also implies a structure-activity relation- ship. In this particular case, the selectivity of a reaction is governed by the size of product molecules and by the transition state. Geometric considerations related to the size of the crystalline micropore channels and cavities in the zeolite here prevail. ad 3. LONG RANGE INTERACTION Long-range effects can be defined as changes in the environment of the catalytically reactive site due to alterations in composition or geometry of atoms in the second coordination shell with respect to the chemisorbed molecule. The particle size effects discussed above for metal particles, if not due to changes in geometry of the catalyt- ically active site, belong to the long-range effect category. Changes in the surface dipole layer for different metal surfaces or on surface kinks and steps also belong to this category. An example of a work function change that affects the topology of the chemisorbed molecule is the effect of positively charged potassium ions on the coordination of CO. Under the influence of the positive charge of potassium on the Pt surface, CO molecules shift from atop to bridging coordination sites. In zeolites also changes in acidity occur due to changes in composition at positions beyond the first coordination shell tetrahedron with respect to the Bransted acid site. An interesting unresolved question is the range of the long-range interactions, a subject to which we will also refer in this book. ad 4, SURFACE COMPOSITION In alloy catalysis it has been discovered that the surface composition and bulk composition of a material may significantly differ. As a result, catalysis has to be studied in relation to the surface composition rather than the bulk composition. In addition, if the reactivity of the surface atoms is different the surface composition may change, favouring enrichment of surface atoms of highest reactivity. The driving 10° CHAPTERI force to this phenomenon is minimization of surface-energy. As a consequence of the surface energy minimization principle, small metal particles will have particular shapes which may depend on gas-phase composition. Ionogenic materials in vacuum will only expose their electrostatically neutral surfaces. As discussed in chapter 4 the stability of microporous solids in contact with a liquid phase is also ruled by this surface-energy minimization principle. One may wonder whether there is a relation between kinetic data collected on uniform model catalyst surfaces and practical catalysts with a nonuniform distribu- tion of catalytically active sites. A thorough analysis of this question is found in the book by Boudart and Djéga-Mariadassow!4l, based upon pioneering studies of Temkin and other Russian kineticists on ammonia synthesis. Assuming an expo- nential distribution of adsorption energies and a linear variation of the activation energy, one finds that the form of the overall rate constant is very similar for uni- form and nonuniform distribution of active sites, so that the same type of kinetic equation applies. But the interpretation of the kinetic constants is different for the two cases, Fundamental to a kinetic analysis is the concept of the most abundant reaction intermediate (MARI) and the rate-limiting step in a reaction sequence. On a working catalyst on which many surface species may be present, some not active in the reaction but just as a spectator, one has to identify the species that partici- pates in the catalytic reaction chain. The dominant species is the one that is most difficult to convert. It is one of the challenges of in-situ spectroscopy to identify the MARI or MARIs proposed in the derivation of an overall kinetic equation. With the help of this information the theoretical basis for a reaction mechanism can be developed. We will consider the overall kinetic equation for a heterogeneous reaction and compare this with the expressions used in homogeneous or enzyme catalysis!!), Let us consider the sequence of two reversible elementary steps: A Ait 5S; = Bit Sz (1.1) ka ka Sz+ Az = Bot+Si (1.2) kaa A, + Az = Bi + Bo S; and Sz denote empty and occupied surface site positions respectively, with L=54+5. As is shown by Boudart, the turnover rate becomes: kika[Ai}[Aa] — k-1k—2 [Bi)[B2) = BIA + kalBil + falda] + bBA] 8) < and CONCEPTS IN CATALYSIS. 11 (Si) _ kali) + bolAo] [Se] Ai[Ai] + &-2[Ba] Let us assume that step k; is rate determining. Define @ as the fraction of occupied sites: (1.4) IS) _ 1 Sa] 7 a) The expression for the turnover rate becomes: al Aa]@ = ky[Ao] K fy = 1.6 = 1+ x fy a8) with k AL =e (1.7) The turnover rate becomes proportional to the surface coverage 9. Equation (1.6) is the Langmuir- Hinshelwood expression and is very similar to the Michaelis - Menten expression used in enzyme catalysis. Consider the following mechanism describing enzyme catalysis: p+sixEeyp (1.8) ki ke where E is the enzymatic site, S the substrate, X the enzyme substrate complex and P the product of the reaction. Using the steady-state condition and assuming step ky to be rate limiting, one derives for the turnover rate for product formation: &: 1 dy _ 9, = Ease! __bKCs ie x = 1.9) Cay’ a T+ pbges 14K (1.9) with 0, the steady state fraction of enzyme sites complexed with S4. Cpo is enzyme site concentration, Cs is substrate concentration. One notes the great similarity of Eq.(1.9), the Michaelis-Menten rate equation of enzyme catalysis and Eq.(1.6), the equation for the surface reaction rate. The similarity between the expressions stems from the conservation of surface sites or enzyme molecules available to the overall reaction. Essential factors here are the complexation constant f- (Eq.(1.9)) or adsorption equilibrium constant K (Eq.(1.6)) and the rate of decomposition of the adsorption complex. Since at low reactant concentration the activation energy is determined by the relation: din(ky.K) dh) (1.10) (Hits Bact = ~Ro 12° cuarreni but at high reactant concentration the activation energy becomes: , __p dint) [Hing Pact = Pa “acy one derives that relevance of kz or K with respect to the overall reaction rate depends on the occupation of the catalytically active sites 0. In Eq.(1.10) and Eq.(1.11) Rg is the gas constant. If the surface coverage or site occupation @ increases from 0 to 1, the overall reaction order changes accordingly from 1 to 0. The examples considered concern reactions with only one product. Analogous expressions can be derived for reactions with several products. The selectivity is found to depend on the ratios of decomposition rates and complexation constants. (1.11) Clearly the kinetics of homogeneous and enzyme catalysis as well as heteroge- neous catalysis are closely related. They are governed by the geometric and elec- tronic details of the catalytically active site. The molecularization of heterogeneous catalysis implies an increasing aptitude to manipulate catalyst preparation, so as to direct catalytic reactivity towards required selectivity. At the root of this molecu- lar knowledge is the increasing sophistication of spectroscopy that can be usefully applied to complex catalytic systems. Spectroscopic as well as mechanistic interpre- tation is founded on advances in theoretical chemical techniques currently able to simulate systems of catalytic relevance. It is to the description of those endeavours that this book is committed. 1.1 References la. M. Boudart, G. Djéga - Mariadassou, "Kinetics of Heterogeneous Reactions”, Princeton University Press, Princeton, N.J. (1984). 1b. J.M. Thomas, W.J. Thomas, "Introduction to Heterogeneous Catalysis”, Academic Press (1967). 2. P.Stoltze, J.K. Norskov, Phys. Rev. Lett, 55 (1985) 2502; P. Stoltze, Physica Scripta 36 (1987) 824. 3. S.H. Oh, G.B. Fischer, J.E. Carpenter, D.W. Goodman, J. Catal. 100 (1986) 360. R. Hoffmann, Science 211 (1981) 995. 5. G.A.Somorjai,” Chemistry in Two Dimensions: Surfaces”, Cornell University Press (1981). 6. P. Mars, D.W. van Krevelen, Chem. Eng. Sci. Suppl. 3 (1954) 41. 7. J.M. Thomas, R.M. Lambert, "Characterization of Catalysts”, J. Wiley (1980). 8. G. Ertl in Catalysis 4, chapter 3, Ed. J.R. Anderson, M. Boudart, Springer (1983). 9. Kaldor, R.L. Whetten, D.M. Cox, D.J. Trevor, A Kaldor, Phys. Rev. Lett. 54 (1985) 1494. 10. J.M. Thomas, Ang. Chem. Int. Ed. Engl. 27 (1988) 1673. * CONCEPTS IN CATALYSIS 13 11. J.T. Gleaves, J.R. Ebner, T.C. Kuechler, Catal. Rev. - Sci. Eng. 30 (1) (1988) 49. 12. R.A. van Santen, W.M.H. Sachtler, Adv. in Catal. 26 (1977) 69. CHAPTER 2 ‘THEORIES OF CHEMISORPTION 2.1 Introduction. In order to appreciate the different theoretical approaches of use in chemisorption theory, it is necessary to know some of the fundamental observations on chemisorp- tion derived from surface science studies. Each theoretical method has its limita- tions, so it is also useful to have an idea of the current status of theoretical chemistry. In this section these two topics will be highlighted and an outline of the material to be presented in this chapter will be given. Before the era of modern surface science in metal catalysis alloys were inves- tigated to solve the question of the so-called ’electronic factor’ in heterogeneous catalysis. Experiments were aimed at gaining an insight into the electronic basis for the unique position of the group VIII metals as hydrogenation catalysts. The group VIII metals, i.e. Fe, Co, Ni; Ru, Rh, Pd; Os, Ir, Pt, have d-valence electrons distributed over a narrow nearly filled d-valence electron band and a broader s,p- valence electron band. The d-valence electron-band is completely filled for the IB metals, Cu, Ag and Au, that have a high activation energy as hydrogenation cata- lysts and are little active as hydrocarbon conversion catalysts. For this reason one of the propositions of that time was that the catalytic activity relates to the number of d-valence electron holes. This seemed reasonable since the number of s, p-valence electrons varies little. It is of the order of one electron per atom. The distribu- tion of valence electrons over d- and s,p-electron bands is sketched in Fig.(2.1). Information on the distribution of valence electrons is derived from photoemission measurements, in which the kinetic energy of emitted electrons is measured as a function of the frequency of irradiating light. The kinetic energy of the emitted electrons (Exjn) is related to the valence electron energy of the electrons (Fyq)) via the Einstein relation: Ein = hy — Evat (2.1) In Eq.(2.1) v is the frequency of light and 4 Planck’s constant. Before the develop- ment of photoemission techniques one could measure the electronic structure only indirectly, e.g. by measuring the thermal or electric conductance. These properties are mainly governed by the energy density of the electrons in the highest occupied orbitals of the metal p(Er). The energy of the highest occupied molecular orbital and the lowest unoccupied orbital is the same in a conductor and is defined as the Fermi level Ep. In the fifties, catalytic chemists tried to establish relations between catalytic activity and p(Ep). Alloying may change (Ep) considerably. Sometimes these changes correlate with catalytic activity, sometimes they do not. THEORIES OF CHEMISORPTION 15 fe “(itt EF Figure 2.1. Distribution of valence electrons over d- and s, pelectron bands for a transition metal (schematic). a: group VIII metal; b: group IB metal. E kcal 9 O 20 Au Pd 80 4 6 Figure 2.2. Activation energy of H2/D2 by Pd/Au alloys as a function of conversion. ‘The broken line x denotes the paramagnetic susceptibility in arbitrary units!"), ‘An example is provided by the work of Couper and Eley!!!, who demonstrated that the rate of the reaction: Hy + Dz —+ 2HD decreases dramatically if palladium is alloyed with gold. As shown in Fig.(2.2) this decrease in catalytic activity correlates with a dramatic decrease in magnetic susceptibility, i.e. p(Ep). We will return later to the question which features of the 16 CHAPTER? mA ° SNS AS" Cc \ | i i i i } i } i i | : : : METAL, Ho Ho Hc c c CH \ on NY \ oe / ' | i | s ‘ . ALLOY Figure 2.3. Primary ensemble effect. oO c i : i i ’ mw. s site blocking Figure 2.4. Secondary ensemble effect. ‘THEORIES OF CHEMISORPTION 17 electronic structure of this alloy are responsible for the decrease of p(Ep) with Au composition and whether one can theoretically justify a relation between chemical reactivity and p(E£p). Progress in the development of surface techniques at present not only makes it possible to distinguish between surface and bulk electronic structure, it also pro- vides the possibility to distinguish between surface and bulk atoms. Electron emis- sion techniques can be employed exploiting the difference in electron freepath as a function of electron energy. Other techniques provide direct information on the sur- face composition; e.g. Low Energy Ion Scattering (LEIS) relates the kinetic energy of scattered noble atom ions directly to the mass of the surface atom scatterers. A classical method to probe the surface composition of an alloy is the use of chemisorp- tive titration. This can be applied if alloys are studied that consist of a mixture of elements such that the atoms of one element strongly interact with the gas atoms used for chemisorptive titration (e.g. Pd adsorbs CO or H2), but the other elements only very weakly (e.g. Au does not adsorb CO or Hz). One finds that for alloys the surface composition is usually different from the bulk. In chapter 5 we will return to theories useful to predict surface enrichment of alloys. Catalytic properties correlate well with the surface composition of alloys. The existence of a surface phase explains why correlations with p(E{p) measured in the bulk sometimes are and sometimes are not found. A correlation has to be established with a property that characterizes the surface electronic structure and not the bulk electronic structure. Alloys can also change the selectivity of reactions. Alloying of Ni with Cu changes hydrocarbon conversion by Ni in the following way: a. The rates of hydrogenolysis, the reaction that decreases the size of the molecule by breaking C-C bonds, become suppressed. b. The rates of isomerization reactions, in which the number of carbon atoms in the molecule remains the same, decreases much less, whereas hydrogenation or dehydrogenation reactions change little. The rates are expressed as molecules reacted per unit time per Ni atom exposed. One finds in general that association and dissociation reactions are affected most if one studies the effect of an inactive component such as Cu or Au on the catalytic properties of an active component Ni. Changes in selectivity can be explained on the basis of two effects: a, geometric b. electronic Geometric effects can be subdivided into so-called primary and secondary ensemble effects. According to the primary ensemble effect a change in surface reactivity occurs, if a molecule adsorbs with several atoms to the metal surface (see Fig.(2.3)). Molecules adsorbed in this way, when partially stripped of hydrogen atoms, will hydrogenolyse. If the surface is diluted with metal atoms that are catalytically much less active (e.g. Cu), the probability decreases that active metal atoms (e.g. Ni) keep active metal neighbors. As a result the probability for multisite adsorption is reduced 18 CHAPTER? and hence the rate of the hydrogenolysis reaction. The secondary ensemble effect decreases the strength of the metal surface-adsorbate atom bonds, by diminishing the surface metal atom coordination number of the adsorbate atom in contact with the metal surface Fig.(2.4). An example is the adsorption of CO to the Pd surface. CO coordinated to Pd surfaces may chemisorb atop to a single Pd atom or bridge between two Pd atoms. The CO frequency in the atop configuration is 2050 cm™!, in the bridging configuration 1980 cm~1. In a subsequent section we will discuss the reasons for this frequency difference. While on nonalloyed Pd predominantly CO in bridging coordination occurs, one observes that upon alloying with silver the probability increases for CO to chemisorb atop”) Fig.(2.5). A simple geometric explanation for this phenomenon can be given. The interaction of CO with Ag is much weaker than that with Pd and the probability for Pd neighbor atoms decreases. Ina later section we will discuss the clectronic basis of this phenomenon. Modern surface science as well as molecular beam techniques indicate very large variations in surface reactivity if one compares different metal surfaces (semi-infi- nite systems) or metal particles of different size. As an example in Fig.(2.6) for different Fe and Re single crystal surfaces the turnover numbers (rates expressed as molecules converted per unit time per surface atom exposed) are given for the con- version of Nz and H2 to ammonia. Both the geometric arrangement of surface atoms and their average coordination number change. The larger the average coordinative unsaturation of the surface atoms, the higher the turn over number for ammonia synthesis. One of the primary aimes of the theoretical discussion presented in this chapter is to study the reactivity of surface atoms as a function of coordination as well as valence electron occupation. The large differences in reactivity as a function of particle size Fig.(2.7), found for very small particles in molecular beams, are espe- cially of interest since current quantum-chemical ab-initio methods are best suited to study the reactivity of small metal particles. Metal catalysts used in hydrocarbon reforming processes contain metal particles of very small size (<2 nm). This is also the case for catalysts with metal particles distributed in the micropores of zeolites. The experimental results indicate that the behavior of small metal particles may be very different from that of large surfaces. As a matter of interest this is also found theoretically. The differences in covalent bonding that are the result of changes in electron density relate to some extent with differences in ionization potential. This is nicely illustrated by the molecular beam results presented in Fig.(2.8). Changes in ionization potential do also play a role on metalsurfaces. Because of the electric double layer caused by spill-over of surface electrons to the vacuum, the ionization potential of a metal surface (the workfunction) decreases if one com- pares the workfunction of a close-packed surface containing surface atoms of high coordination, with a more open surface, containing surface atoms with a lower co- ordination. As we will discuss, not only changes of the surface-dipole layer occur if one compares different surfaces, but also metal-metal atom distances may change (they usually decrease if one compares dense surfaces with more open surfaces). In this chapter first the current status of quantum chemistry will be summarized with respect to the problem of surface reactivity and a start will be made with the ‘THEORIES OF CHEMISORPTION 19 2100 2000 1900 1800 1700 cm~1 Pd/Ag= 9.75/0.25 9.5/0.5 9/1 3.5/6.5 Figure 2.5. Infrared spectra of CO adsorbed to Pd-Ag alloys. —Pco = 0.01 Torr; -— Poo = 0.5 Torr!’ theoretical treatment of chemisorption. We begin with semi-empirical approaches such as the Extended-Hiickel method or the tight-binding method, as it is called by solid-state physicists, because they are convenient to analyze. The concepts devel- coped by applications of these methods to the surface reactivity problem will be used to analyze also the results of more extensive and accurate calculations. At present surface chemistry is in a comparable situation to that of organic chemistry in the | sixties. Using elegant approaches and a simple method such as the Extended-Hiickel method, Woodward and Hoffmann] developed important concepts extremely fruit- ful to organic chemists. An enormous body of theoretical work evolved to test the concepts with the aid of highly accurate quantitative calculations. Such calculations have been rendered possible by the advent of supercomputers, which are steadily improving in computation power. At present they are applied also to large-scale quantum-chemical calculations on small metal clusters and have even become avail- | able for use on metal surfaces. Again especially due to efforts of Hoffmann!4l, but also many others("), the Extended- 20 CHAPTER? Fe(111) Fe(100) Fe(110) Cg (solid) Cg (broken) Sh, SB C5 (solid) Cq(solid) C4o (broken) Figure 2.6. Top figure: Surface structure sensitivity of iron-catalyzed ammonia syn- thesis. Bottom figure: Strong surface structure sensetivity of ammonia synthesis on rhenium single crystals. ‘The symbols C, indicate n-coordinated sites. Yield is expressed in arbitrary unite), ‘THEORIES OF CHEMISORPTION 21 2 4 6 8 10 12 14 16 18 CLUSTER SIZE Figure 2.7. Reactivity of metal clusters with Dz as a function of clustersize. Triangles denote metal cluster responses in the absence of D2. Drawn peaks denote cluster responses in the presence of D2. Loss in intensity implies a large cluster reactivity'®), Hiickel method has been applied to small metal clusters and to the study of elemen- tary surface reactions], Important concepts have evolved, which will be discussed and can be verified with ab-initio calculations. Earlier, physicists had been oc- cupied with the question of the relation between chemisorption and surface states using related tight-binding techniques. Their adaptation of those techniques to the surface problem is of especial interest for the development of embedding methods, which study changes in chemisorption if a cluster becomes embedded in a semi- infinite surface. Currently this work is being extended to systems including explicit treatment of electron-electron interactions using extensions of local density approx- imation (LDA) solid-state methods. Interestingly the Hartree-Fock-Slater method, a quantum-chemical version of the LDA method developed by quantum chemists to study organometallic complexes, is being extended to the study of transition-metal surfaces(!®), In the next section we will start with a short theoretical intermezzo to show the approximations involved in the use of the Extended-Hiickel and related methods and introduce some basic theoretical analytical tools. Then the tight-binding method will be applied to study changes in electronic structure in a one-dimensional system 22 cuapren? lectron Binding Energy (eV) Relative Recctivity 5 10 15, 20 25 Cluster Size (No. of Atoms) Figure 2.8. Cluster ionization potential as a function of cluster size. For cluster sizes larger than 10 an apperent relationship with reactivity is found"®), simulating surface formation. In the next section this approach will be extended to the study of three-dimensional systems, where chemisorption will be related to surface states. Then the basic concepts of LDA theory will be introduced and their applications to chemisorption discussed. Since electron-electron interactions can now be accounted for, the physics of the image potential in relation with chemisorption can be introduced. The results of first-principle calculations on chemisorption to metal surfaces will be presented and interpreted. For chemical applications, analysis in terms of formal chemisorption theory is useful. This is illustrated by analyzing Hand CO chemisorption to transition-metal surfaces. The chapter will be finished with a discussion of the use of free-electron theories. Here again simplifications are possible and it will appear that such theories are complementary to theories based on molecular orbital approaches. ‘THEORIES OF CHEMISORPTION 23 2.2 Theoretical Intermezzo 1. Background to the Extended-Hiickel Method and other Semi-Empirical Methods. ‘The classical energy of a system can be written as: E = Exin + Epot (2.2) According to quantum mechanics the motion of the electrons is to be calculated from wave functions (r). The chance density to find an electron at position r is given by: AT) = b'(F).d(F) (2.3) Since the total chance to find an electron in space eqals 1, it follows that: [are @).0@) = [ao =1 (2.4) dr denotes an infinitesimally small volume element dr. Relation (2.4) defines the normalization of p(r). The wave functions (r) are the solutions of the Schrddinger equation: [P+ V7) vir) = Bd) (25a) AF ).vi(F) = Ev(F) (2.56) 7(*) and V(#) are kinetic and potential energy operators. The operator H{(7) is called the Hamiltonian operator. For a particle with only one degree of freedom along the z-axis, the kinetic energy operator 7'(z) equals: P “2m de® If the potential V(r) only depends on position, as presupposed in relation (2.5a), the potential energy operator equals the classical potential. (2) is the solution of: a yey + )o(z) = Ed(z) (2.7) 2m dz ave) = i . Because of the normalization condition Eq.(2.4) this equation has a solution for only a limited number of values Ej. These values are the eigenvalues of the Schrédinger equation. The corresponding solutions for (z) are called the eigenfunctions. We shall focus the discussion on the approximations that lead to the Extended-Hiickel method. In the course of its derivation we will also have an opportunity to refer to a few other methods such as the Local Density Approximation (LDA) and the Xa method, A major assumption involved is that the electrons move in the average potential of the other electrons. This implies that the for n-electrons the electron T(z) = (2.6) 24 CHAPTER? distribution p(71,......;7n) is equal to the product of the probability distributions of the single atoms: OF sonFn) = TL oC) (28) This is the equivalent of the mean field assumption in statistical mechanics. The n-electron wave function can be written as the antisymmetrized product of one-electron functions ¥;(T;,0;). 0; denotes the spin of the electron: UF 1. Fnyonon) = AT] ay (F 1501)-ag(F2502).-Yan(F ny on) (2.9) Ais the antisymmetrization operator. One can improve on this one-electron approximation if one includes electron correlation by assuming the n-electron state function (71... 7 n,01...n) to be a linear combination of functions (7) ... Tn,o1---on) with the electrons distributed over different one-electron functions ¥a;(7,0). This leads to so-called configuration interaction methods and the n-electron density cannot be written any more in the form of Eq.(2.8) because now the electron motions become correlated. If one deals with weak interactions such as the van der Waals interaction, that depends on the polarization of the electrons in one fragment due to the fluctuating dipole of the electrons in the other fragment, the one-electron approximation does not satisfy and configuration interaction has to be considered explicitly. ‘The wave functions can be written as a product of a spatial function and spin function if electromagnetic interactions are ignored: vilF 0) = vir) (2.10) Different methods exist to compute wave functions (7) that are the solutions of equations of the type Eq.(2.5.) If the potential V(r) is small compared with the kinetic energy, it is useful to consider ¥(7) approximately as a sum of plane waves, an approximation often used in solid physics. If ¥(#) is approximated by a single plane wave the method is called the free-electron method. Such solutions are discussed later. We will first consider the wave function ¥;(7) to be a linear combination of atomic orbitals centered on the lattice atoms. vilF) = DO Ges) (2.11) d This method is called the LCAO (Linear Combination of Atomic Orbitals) method. The electron density p(7) becomes: AF) =D mpi) = Do mali)? (2.124) ‘THEORIES OF CHEMISORPTION 25 = x [zt IGE les(F)P + 2x Reef cf etres(7)} ou (2.128) nj is the orbital occupation. According to the Pauli principle the orbitals of a molecule are occupied such that only a maximum of two electrons per molecular orbital ¥;(r) are allowed. The contribution due to the first term in Eq.(2.12b) can be considered as the sum of the electron densities on the atoms. The term: Ee Rect & oi (F)9j(F me (2.13) igg due to the interference of atomic wave functions y;(7) and y;(F) is responsible for covalent bonding. It relates to the interatomic electron density. If the electrons are considered to move independently, Eq.(2.8) is satisfied and the total Hamiltonian can be written as: AF... Fn) = DAG +O Zrbye (2.14) yey | Ray | f(#;) is the one-electron operator that operates on a single electron and has a form similar to Eq.(2.5a). The second term in Eq.(2.14) is the nuclear repulsion term. Ryy represents the nuclear distance and Zy the nuclear charge. The general form of H(#) is: Weg A) = P)- Tes far Oe + Blezch,7) (2.15) Ry-F | [r-r] ‘The second term in Eq.(2.15) is the nuclear attraction between nuclei and electrons, the third the repulsion an electron experiences due to the presence of other electrons and the last operator /3(ezch, 7) is the contribution due to the so-called exchange energy. It arises from the Pauli requirement that the n-electron wave function has to be antisymmetric. H(exch, 7) partly compensates for the interaction of the electron with itself included in the electron repulsion term. Several approximations are used to compute H(ezch,7)¥,(7). In the Hartree- -Fock approximation it follows directly from substitution of Eq.(2.9) in the exact n-electron Hamiltonian and depends on the molecular orbital considered. For molec- ular orbital j the result becomes: EPP (exch, 7 )bj(r) = — me jf ar! ber) bj(r! oe(F Sopa; (2-16) | 7 a In the Local Density mene one uses the property that E(ezch. 7) is a function of the total density p(7). Often it is approximated by the functional dependence one derives for the exchange energy of a free-electron gas: B(exch,#) ~ p(r)$ (2.17) 26 CHAPTER? In the Xq method as originally introduced, in addition, assumptions are made con- cerning the form of Dy —“*—. It is computed within the muffin tin approxima- tion. Within a sphere radius ay the potential is assumed to be equal to a particular value assigned to atom 7, outside the sphere the potential is assumed to be equal to zero. The advantage of these assumptions is that scattering techniques can be used to find the eigenvalues of the electron distribution. These methods are efficient and can be used for very large systems. We will return to this approximation later. Usually its accuracy is insufficient to study the interaction energy of molecules. In the Hartree-Fock-Slater (HFS)-LCAO method the E(exch, 7) term is approximated bya term similar to Eq.(2-17) but further no assumptions on the potential are made. In the Extended-Hiickel method the interaction the electron experiences from elec- trons on neighboring atoms is supposed to be canceled by the attraction with nuclei. In addition the electron-electron interaction on one atom is averaged or can be sim- ulated by an effective one-center repulsion term. This implies that E(each,7) is independent of the molecular orbital j, The Extended-Hiickel method therefore is to be considered an approximation to the LDA. The approximation involved corre- sponds with the situation that the atoms are assumed to be neutral with localized electron distributions. Since the eigenvalues Ej are the solution of equations of type Eq.(2.5a), they are given by: By = [dr ve) Aw) (2.18) The values B; follow from the set of secular equations generated from Eq.(2.5) and Eq.(2.11): HF oF) = Bedi(7) (2.19) Multiplication at the left by the atomic orbital ¢$(7), substitution of Eq.(2.11) and integration yield: ry [rgd ol) = BIE Sy (2.20) k Defining: Sie = f ar e(rdex(r) (2.21a) Hyp = [erg mer) (2.218) one finds as secular equation: Dj — 215i) =0 (2.22) a ‘THEORIES OF CHEMISORPTION 27 Normalization gives as additional condition: 1= fdrvt@yiF) = Dial +20 Rech Se (2.28) k i

You might also like