correlation between catalytic activity and bonding and coordination

5
Proc. NatI. Acad. Sci. USA Vol. 82, pp. 2207-2211, April 1985 Chemistry Correlation between catalytic activity and bonding and coordination number of atoms and molecules on transition metal surfaces: Theory and experimental evidence (catalysis/transition metals/surface structure/electronic structure) L. M. FALICOV*t AND G. A. SOMORJAItt Departments of *Physics and fChemistry and tCenter for Advanced Materials, Lawrence Berkeley Laboratory, University of California, Berkeley, CA 94720 Contributed by L. M. Falicov, Nov'ember 13, 1984 ABSTRACT Correlation between catalytic activity and low-energy local electronic fluctuations in transition metals is proposed. A theory and calculations are presented which indicate that maximum electronic fluctuations take place at high-coordination metal sites. Either (i) atomically rough surfaces that expose to the reactant molecules atoms with large numbers of nonmagnetic or weakly magnetic neighbors in the first or second layer at the surface or (ii) stepped and kinked surfaces are the most active in carrying out structure-sensitive catalytic reactions. The synthesis of ammonia from N2 and H2 over iron and rhenium surfaces, 'H2/2H2 exchange over stepped platinum crystal surfaces at low pressures, and the hydrogenolysis (C-C bond breaking) of isobutane at kinked platinum crystal surfaces are presented as experimental evi- dence in support of the theory. Metal catalysts are used to carry out a multitude of selective chemical reactions that range from the synthesis of ammonia from nitrogen and hydrogen gases, to hydrogen-deuterium exchange, to selective hydrocarbon conversion reactions (dehydrocyclization, isomerization, dehydrogenation, and hydrogenolysis to mention a few). These catalysts perform selective bond breaking and rearrangements for millions of turnovers and produce product molecules continuously without deactivation (poisoning). Yet modern surface sci- ence studies have revealed by isotope labeling and other appropriate experiments that most of the surface is covered with strongly adsorbed species whose turnover is much longer than the times required to produce the desired prod- ucts for the catalytic reaction (1). It is therefore obvious that those reaction steps that need metal sites must occur at those few locations on the metal surface that remain uncovered. Thus, the questions arise: How does catalysis occur in the presence of these strongly chemisorbed stagnant layers of almost monolayer coverage? What is the nature of the active metal sites that remain uncovered? Surface science experiments have also shown that the chemical bonding of adsorbed atoms and molecules is struc- ture sensitive. The heats of adsprption of atoms and mol- ecules could vary as much as 10-30% from adsorption site to adsorption site on the same crystal surface (2). Top, bridge, threefold, and fourfold sites are available on flat, low Miller index surfaces. Surfaces that are rough on the atomic scale are often found, by chemisorption or by catalytic reaction studies, to bind atoms and molecules more strongly and to dissociate strong chemical bonds (NN, C-H, C-C, C=O) more readily (2). Rough surfaces are those with higher Miller index, in which the outermost layer is more sparsely populated, thereby exposing atoms in the second and other inner layers [for example the (111) face of a body-centered cubic (bcc) crystal as opposed to the (110) face, or the (1120) surface of a hexagonal-close-packed (hcp) crystal as opposed to the closed packed (0001) face], or have a high concentration of step edges and kink vertices. The structures of some of these surfaces are shown schematically in Figs. 1-5. One of the key questions in heterogeneous catalysis that emerges from modern surface science studies concerns the determination of those metal sites where catalytic processes occur (3-6); these must be distinctly different from the sites where strong bonding of atoms and molecules takes place accompanied by a low reactivity and slow turnover. In this paper a theory is presented that argues that catalysis is predominant at high-coordination sites: the more bulk-like the site, the greater the probability of electron-state fluctua- tions, which participate catalytically in the breaking and formation of chemical bonds and yield high turnover. This model points out the catalytic importance of metal atoms located in the second layer of open surface structures-the iron (111) surface for example-which are highly coordi- nated but are nonetheless exposed to the incoming reactant molecules. Similarly, the increased activity at stepped sur- faces is explained by the exposure of high-coordination atomic sites at the bottom of the steps. Adsorbates tend to form strong stable bonds at lower- coordination sites, where the atoms have few electron-state fluctuations, where s- and p-bonding predominates, and where charge-transfer processes are more common; all these effects lead to the formation of more stable surface com- pounds. These low-coordination sites have only secondary d FIG. 1. Geometry for a stepped (111) surface in the face-centered- cubic (fcc) structure. Open circles represent atoms in the second layer, which is complete. Solid circles represent atoms in the first layer, in which every third row is removed to create steps. This arrangement can also describe a stepped (0001) surface of a hcp structure (the difference between the two appears in the third, unseen, layer). 2207 The publication costs of this article were defrayed in part by page charge payment. This article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. §1734 solely to indicate this fact.

Upload: buithuan

Post on 30-Dec-2016

217 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Correlation between catalytic activity and bonding and coordination

Proc. NatI. Acad. Sci. USAVol. 82, pp. 2207-2211, April 1985Chemistry

Correlation between catalytic activity and bonding and coordinationnumber of atoms and molecules on transition metal surfaces:Theory and experimental evidence

(catalysis/transition metals/surface structure/electronic structure)

L. M. FALICOV*t AND G. A. SOMORJAIttDepartments of *Physics and fChemistry and tCenter for Advanced Materials, Lawrence Berkeley Laboratory, University of California, Berkeley, CA 94720

Contributed by L. M. Falicov, Nov'ember 13, 1984

ABSTRACT Correlation between catalytic activity andlow-energy local electronic fluctuations in transition metals isproposed. A theory and calculations are presented whichindicate that maximum electronic fluctuations take place athigh-coordination metal sites. Either (i) atomically roughsurfaces that expose to the reactant molecules atoms with largenumbers of nonmagnetic or weakly magnetic neighbors in thefirst or second layer at the surface or (ii) stepped and kinkedsurfaces are the most active in carrying out structure-sensitivecatalytic reactions. The synthesis of ammonia from N2 and H2over iron and rhenium surfaces, 'H2/2H2 exchange overstepped platinum crystal surfaces at low pressures, and thehydrogenolysis (C-C bond breaking) of isobutane at kinkedplatinum crystal surfaces are presented as experimental evi-dence in support of the theory.

Metal catalysts are used to carry out a multitude of selectivechemical reactions that range from the synthesis of ammoniafrom nitrogen and hydrogen gases, to hydrogen-deuteriumexchange, to selective hydrocarbon conversion reactions(dehydrocyclization, isomerization, dehydrogenation, andhydrogenolysis to mention a few). These catalysts performselective bond breaking and rearrangements for millions ofturnovers and produce product molecules continuouslywithout deactivation (poisoning). Yet modern surface sci-ence studies have revealed by isotope labeling and otherappropriate experiments that most of the surface is coveredwith strongly adsorbed species whose turnover is muchlonger than the times required to produce the desired prod-ucts for the catalytic reaction (1). It is therefore obvious thatthose reaction steps that need metal sites must occur at thosefew locations on the metal surface that remain uncovered.Thus, the questions arise: How does catalysis occur in thepresence of these strongly chemisorbed stagnant layers ofalmost monolayer coverage? What is the nature of the activemetal sites that remain uncovered?

Surface science experiments have also shown that thechemical bonding of adsorbed atoms and molecules is struc-ture sensitive. The heats of adsprption of atoms and mol-ecules could vary as much as 10-30% from adsorption site toadsorption site on the same crystal surface (2). Top, bridge,threefold, and fourfold sites are available on flat, low Millerindex surfaces. Surfaces that are rough on the atomic scaleare often found, by chemisorption or by catalytic reactionstudies, to bind atoms and molecules more strongly and todissociate strong chemical bonds (NN, C-H, C-C,C=O) more readily (2). Rough surfaces are those withhigher Miller index, in which the outermost layer is moresparsely populated, thereby exposing atoms in the secondand other inner layers [for example the (111) face of a

body-centered cubic (bcc) crystal as opposed to the (110)face, or the (1120) surface of a hexagonal-close-packed (hcp)crystal as opposed to the closed packed (0001) face], or havea high concentration of step edges and kink vertices. Thestructures of some of these surfaces are shown schematicallyin Figs. 1-5.One of the key questions in heterogeneous catalysis that

emerges from modern surface science studies concerns thedetermination of those metal sites where catalytic processesoccur (3-6); these must be distinctly different from the siteswhere strong bonding of atoms and molecules takes placeaccompanied by a low reactivity and slow turnover. In thispaper a theory is presented that argues that catalysis ispredominant at high-coordination sites: the more bulk-likethe site, the greater the probability of electron-state fluctua-tions, which participate catalytically in the breaking andformation of chemical bonds and yield high turnover. Thismodel points out the catalytic importance of metal atomslocated in the second layer of open surface structures-theiron (111) surface for example-which are highly coordi-nated but are nonetheless exposed to the incoming reactantmolecules. Similarly, the increased activity at stepped sur-faces is explained by the exposure of high-coordinationatomic sites at the bottom of the steps.

Adsorbates tend to form strong stable bonds at lower-coordination sites, where the atoms have few electron-statefluctuations, where s- and p-bonding predominates, andwhere charge-transfer processes are more common; all theseeffects lead to the formation of more stable surface com-pounds. These low-coordination sites have only secondary

d

FIG. 1. Geometry for a stepped (111) surface in the face-centered-cubic (fcc) structure. Open circles represent atoms in the secondlayer, which is complete. Solid circles represent atoms in the firstlayer, in which every third row is removed to create steps. Thisarrangement can also describe a stepped (0001) surface of a hcpstructure (the difference between the two appears in the third,unseen, layer).

2207

The publication costs of this article were defrayed in part by page chargepayment. This article must therefore be hereby marked "advertisement"in accordance with 18 U.S.C. §1734 solely to indicate this fact.

Page 2: Correlation between catalytic activity and bonding and coordination

2208 Chemistry: Falicov and Somorjai

fcc (977) fcc (755) fcc (533)fcc (14,11,10) fcc (10,8,7) fcc (13,11,9)

fcc (443) fcc (332) fcc (331)

FIG. 2. Various stepped surfaces in the close-packed fcc lattice.

importance in catalytic reactions that require specific metalsites, as they may sterically block catalytically active high-coordination atomic sites, participate in diffusion or atomtransfer processes, or change the local work function.

THEORYThe chemisorption of an atom or molecule B at a givenlocation in an A transition-metal surface is essentially aconsequence of the local electronic properties both of theatom or molecule and of the metal site or sites in theimmediate neighborhood.A given electronic configuration of, say, an isolated

transition-metal atom A and atom B either leads to theformation of a stable AB compound or yields an unreactiveno-compound-formation situation. This typical either-or re-action situation is the opposite to what is needed in an

effective catalyst, where in principle A should attract B toform a weakly bound AB intermediate compound but bereadily available to release B in the presence of C so that thefinal product BC be formed and A be ready to start theprocess all over again.

This traditional view of the intermediate-compound theoryof catalysis can be viewed from a different point of view: theparticular transition-metal atom A under considerationshould be susceptible to stable low-energy electronic fluc-tuations in its original state. In that case the intermediateAB compound-intrinsically related to the fluctuations-canbe easily formed. Since low-energy fluctuations are alsopresent after a weak AB is formed, the compound can alsobe easily broken and the catalytic reaction may thus takeplace. The electronic fluctuations of importance may be ofmany kinds. In particular:

(i) Charge fluctuations, in which the electronic charge inthe vicinity of A fluctuates about its mean value. Atoms of ametal are the traditional examples of charge fluctuators. Theclassical noninteracting band model of alkali metals consistsof a bulk atom neutral on the average, but with a 1/4probability each of being positively or negatively singlycharged. Electron-electron coulomb repulsion tends to de-crease or suppress large charge fluctuations.

(ii) Configuration fluctuations, in which the occupationnumbers of single-electron orbitals are no longer good quan-tum numbers. Transition-metal and rare-earth compoundsand structures are typical examples of configuration fluctua-tors. The nickel configuration alternates, for instance, be-tween (3d)10, (3d)9(4s)1, and (3d)8(4s)2. The cerium atom, a

commonly known valence or configuration fluctuator, has its

FIG. 3. Various kinked surfaces in the close-packed fcc lattice.

four outer electrons fluctuating among the 4f, 5d, and 6sshells.

(iii) Spin fluctuations, in which the magnitude and theorientation of the local magnetic moment fluctuates about itsmean vectorial value. Spin fluctuations are common in therare-early and in the 3d-transition-metal compounds.

(iv) Term and multiplet fluctuations, present in incom-plete shells of complex structure, and common in transition-metal, rare-earth, and actinide elemental and compoundsolids.

All these fluctuations require the presence of low-energyelectronic excitations which must have a large spectralweight in the catalytically active site.The most important criterion in determining the presence

or absence of fluctuations is the location of a given elementin the periodic table. There are several tabulations of theheat of adsorption of small molecules (CO, H2, 02, N2) ontransition metals that exhibit decreasing values with increas-ing d-shell occupancy (7). There are also correlations re-ported for the variations of catalytic activities for hydro-genolysis, hydrogenation, and other chemical reactions thatshow trends dependent on d-shell occupancy (8, 9). Fromthis evidence it is clear that the catalytic activity of thetransition-metal atoms and ions is directly related to theirunfilled d shells.Both s and p shells tend to form deep-lying filled (bonding)

orbitals, highly energetic unoccupied antibonding orbitals,or wide metallic conduction bands with a relatively lowdensity of states at the Fermi level. None of these situationsleads to the sought-for low-energy fluctuations.The highly degenerate f shells are dominated by the very

strong electron-electron intrasite coulomb repulsion. There-fore f levels are essentially atomic in character: most localcharge and configuration fluctuations are suppressed. Thelevels are discrete and only occasional configuration or termfluctuations, as is the case of cerium, survive.The d shells provide the ideal happy medium: the 10-fold

degeneracy and relative narrowness of the bands providemetallic character with a high density of states.- But thebands are not dominated by coulomb repulsion effects withthe consequent "'discretization" of the spectrum, appear-ance of energy gaps, formation of magnetic moments, andloss of "metallic" character of the d orbitals-all conditionsthat lead to the suppression of local electronic fluctuations.The effect changes across the periodic table. The elec-

tronic fluctuations increase in going from the 3d to the 4d tothe 5d series: the d bandwidth increases in that direction, thecoulomb repulsion effects are less dominant, magnetismdisappears, and ordinary configuration fluctuations increase.The effect increases from left to right in the periodic table:the group VIII elements are the most favorable.

Because of computational considerations we have ex-plored theoretically in detail the local electronic structure ofvarious surfaces of the late 3d transition metals: Ni, Co, Fe,and Cr. We have also explored heterogenous structuresinvolving either two of these elements or one of the elementsand their noble-metal neighbor, Cu.

Proc. Natl. Acad. Sci. USA 82 (1985)

Page 3: Correlation between catalytic activity and bonding and coordination

Proc. Natl. Acad. Sci. USA 82 (1985) 2209

Table 1. Number n of d holes for nickel atoms invarious configurations

Configuration Z(Ni) Z(Cu) n

Bulk nickel 12 0.56(111) Surface nickel atom 9 0.38Outstep nickel atom a (Fig. 1) 7 0.25Outstep nickel atom b (Fig. 1) 7 0.24Instep nickel atom c (Fig. 1) 11 0.52Instep nickel atom d (Fig. 1) 10 0.45Ni under a (111) Cu monolayer 9 3 0.52Ni monolayer on Cu (111) 3 6 0.46

Nickel is particularly interesting, since it is the last of the3d group VIII transition metals. When nickel is in its metallicform, the ground state is a mixture of three configurations(10) that, on the average, result in the d shell being occupiedby 9.44 electrons. The presence of the n = 0.56 d holes,which are a measure of the configuration fluctuations, isresponsible for the catalytic activity of Ni. The larger n, thelarger the configuration fluctuations of a particular Ni atom,and the greater the catalytic activity of that particular site.Our calculations for nickel in its paramagnetic state-whichshould also be applicable to palladium and to platinumyield a close correlation between n and the coordinationnumber Z of a particular site (10). Our results are shown inTable 1, where Z(A) denotes the number of nearest neigh-bors of chemical species A.An analysis of these results yields the following simple

conclusion: configuration fluctuations are larger for thoseatoms that have a large number of metallic neighbors.Therefore transition-metal atoms in sites that are bulk-like,have a high metallic coordination, and are accessible to the

reactants should be, according to this argument, catalyticallyactive.Another measure of the fluctuations in transition metals is

the local spin imbalance at each site. When the coulombrepulsion between electrons becomes strong enough, thesystem tends to lessen or minimize its influence by suppress-ing metallic fluctuations. This suppression results in a ten-dency either towards an insulating state or to the formationof magnetic states (11, 12) with local spin imbalance (localmagnetic moment) and, at low temperatures, long-rangemagnetic order (ferromagnetic, antiferromagnetic, fer-rimagnetic, spiral arrangements, etc.). Those conditions thatlead to the formation or enhancement of spin imbalances donot favor-i.e., they suppress or decrease-catalytic activ-ity.We have calculated the spin imbalance S (twice the local

magnetic moment divided by the g-factor) for a variety ofstructures, sites, and circumstances (13-19). Our results areshown in Table 2.

Since spin imbalances are one of the manifestations ofreduced fluctuations, we expect, for a given element, cata-lytic activity to be anticorrelated to its spin imbalance. Thisanticorrelation and our calculations (Table 2) point once

again to catalytically active sites in locations of high co-

ordination, and preferentially to atoms surrounded by a

large number of nonmagnetic or weakly magnetic metallicneighbors.

EXPERIMENTAL EVIDENCEWe present here experimental evidence for the high catalyticactivity of atomically rough surfaces, where high-coor-dination sites in the first or second layer at the surface areexposed. We also show how stepped surfaces that expose

Table 2. Spin imbalance S of the magnetic transition metals in various configurationsConfiguration Z(Ni) Z(Cu) Z(Fe) Z(Co) Z(Cr) S

Ferromagnetic bulk fcc nickel 12 0.56Nickel in Ni3Cu quenched alloy 8 4 0.42Nickel in NiCu quenched alloy 4 8 0.08Nickel in NiCu3 quenched alloy 0 12 0Nickel at the (100) free surface 8 0.74Nickel at the (111) free surface 9 0.65Nickel under a (100) Cu monolayer 8 4 0.37Nickel under a (111) Cu monolayer 9 3 0.38Nickel monolayer on top of Cu (100) 4 4 0.48Nickel monolayer on top of Cu (111) 6 3 0.08

Ferromagnetic bulk hcp cobalt 12 1.56Cobalt in the FeCo ordered alloy 8 0 1.78Cobalt at the (100) surface of the FeCo alloy 4 0 2.03Cobalt at the (110) surface of the FeCo alloy 6 0 1.86Cobalt monolayer on top of Cu (111) 3 6 1.63

Ferromagnetic bulk bcc iron 8 2.12Iron in the FeCo ordered alloy 0 8 2.66Iron at the (100) free surface 4 2.90Iron at the (110) free surface 6 2.55Iron at the (100) free surface of the FeCo alloy 0 4 2.95Iron at the (110) free surface of the FeCo alloy 0 6 2.75Iron under a (100) Cr monolayer 4 4 1.95

Antiferromagnetic bulk chromium (incommensurable phase) 8 0.59Antiferromagnetic bulk chromium (commensurable phase) 8 0.67Chromium at the (100) free surface 4 3.00Chromium at the (110) free surface 6 2.31Chromium monolayer on top of Fe (100) 4 0 3.63Chromium monolayer on top of Fe (110) 2 4 2.25

Chemistry: Falicov and Somorjai

Page 4: Correlation between catalytic activity and bonding and coordination

2210 Chemistry: Falicov and Somorjai

(111)

C4 (solid)C7 (dashed)

(111)

(100) (110)

C4 (solid) C6 (solid)

C8 (dashed) C8 (dashed)

lO,000

(1120) (1010)

C7 (solid) C6 (solid)

C11 (dashed) C10 (dashed)

(0001)

C9 (solid)

5,000

(100)

*0 1,000

t 5005)

;10050

(110)

FIG. 4. Surface structure sensitivity of iron-catalyzed ammoniasynthesis. The symbols C,, indicate n-coordinated sites. Yield isexpressed in arbitrary units.

(1120)

(1010)

(0001)Nonactive

FIG. 5. Strong surface structure sensitivity of ammonia synthe-sis on rhenium single crystals. Symbols as in Fig. 4.

high-coordination sites at the step inner edges and at kinksexhibit high turnover rates for various reactions.

Greatly Enhanced Rates of Ammonia Synthesis on OpenCrystal Faces of Iron and Rhenium That Expose the High-Coordination Atoms in the First and Second Layers. Therate-determining step in the synthesis of NH3 from N2 andH2 is the dissociation of N2 (20). Chemisorption studiesindicate that on the open (111) crystal face, where thehigh-coordination iron atoms in the second layer are exposedto the incoming nitrogen molecules, the dissociation prob-ability of N2 is much higher than on the close-packed (110)face. Chemical reaction studies on iron single crystal sur-

faces, shown in Fig. 4, revealed that the rate of ammoniaformation is 500 times greater on the Fe (111) than the Fe(110) face, while the Fe (100) face exhibits catalytic activitythat is intermediate between the other two crystal faces (3).The closed-packed hexagonal (0001) face of rhenium is

virtually inactive for the synthesis of ammonia, while the open

(1010) and (1120) crystal faces have high reactivities, many

orders of magnitude higher than the (0001) face (4). Theseactive crystal surfaces exhibit an open structure with manyhigh-coordination sites exposed to the incoming moleculesboth in the first and in the second layers at the surface. In Fig.5 we show the relative reaction rates on the different singlecrystal surfaces along with the surface structure, displayingatoms of different nearest-neighbor configurations.

Reactive Molecular Beam Surface Scattering Studies of'H2/2H2 Exchange on Stepped Platinum Single Crystal Sur-faces. The reaction of hydrogen and deuterium molecules toform 1H2H is one of the simplest reactions that take placereadily on transition-metal surfaces, even at temperatures as

low as 100 K. This reaction was studied by using mixedmolecular beams of 1H2 and 2H2 on the (111) and steppedsurfaces of platinum (5). The reaction probability was 0.35on the stepped surfaces, much higher than the 0.07 value on

the (111) surface under identical experimental conditions.The reaction probability also depended markedly on the direc-tion of approach of the reactant molecules to the steppedsurface. This is shown in Fig. 6. The production of 1H2H was

the highest when the reactants struck the open side of the stepstructure, and it decreased by about a factor of 2 when theinner edge of the step was shadowed. These results indicate

that the inner corner atom, which is also the one with the

highest coordination, is the most active for breaking the largebinding energy (103 kcal/mol) of the H-H bond.Enhanced Hydrogenolysis Activity at Kink Sites on Plati-

num Crystal Surfaces. Platinum is widely used as a catalystfor the rearrangement of hydrocarbons. It is particularlyuseful for the conversion of straight chain alkanes to aro-

matic molecules and branched isomers, to produce gasolinewith high octane number. The conversion of n-hexane or

n-heptane is often used as a test reaction that probes theselectivity of various platinum crystal surfaces to yield the

1.0 I

0.9-H (a)_9W)Y EW0t= 900

Pt (332)0.8 -

0.7

0

+ 05aPt (332)

0.4-

0.3-

0.2

0.1 - Pt (111)

-80 -60 40 -20 0 20 40 60 80

Angle of incidence (6), degrees

FIG. 6. Reaction probability for 'H2/2H2 exchange over a plati-num single crystal surface as a function of angle of incidence of themolecular beam, normalized to the incident 2H2 intensity. Curve a,the platinum (332) surface with the step edges perpendicular to theincident beam. Curve b, the platinum (332) surface where theprojection of the beam on the surface is parallel to the step edges.Curve c, the smooth, close-packed platinum (111) surface. 4, Angleof azimuth; I, beam intensity. Frequency of chopping was 10 Hz andsurface temperature was 800°C.

1000

500

;,% 100

*t 50la

5)

m 10z

1

Proc. Natl. Acad. Sci. USA 82 (1985)

Page 5: Correlation between catalytic activity and bonding and coordination

Proc. Natl. Acad. Sci. USA 82 (1985) 2211

0.03

t.)

ct0

._

0.0

0.0

0

Pt (100) Pt (111) Pt (13,1,1) Pt (10,8,7)

FIG. 7. Reaction rates for the hydrogenolysis of isobutane onvarious single crystal surfaces of platinum. Rates are given in unitsof molecules per site per second. The kinked (10,8,7) crystal face isthe most active for this reaction.

desired products (21). The hydrogenolysis of alkanes to pro-duce low molecular weight gases-methane through propane-by C-C bond breaking is an undesirable but important re-action path. It was found that platinum crystal surfaces thathave high concentrations of kink sites are the most active onesin catalyzing this unwanted reaction (22, 23). This is shown inFig. 7. Hydrogenolysis can be inhibited by the introduction ofa small concentration of sulfur, which strongly binds to thesesites and blocks the C-C bond-breaking process.

STRUCTURE SENSITIVITY OF CATALYTICREACTIONS

The examples of catalytic reactions described above (am-monia synthesis, 1H2/2H2 exchange at low pressures, andhydrogenolysis) all indicate pronounced structure sensitiv-ity. This structure sensitivity of catalytic reactions can bedetected either through systematic variation of metal particlesize (24, 25) or by reaction studies using metal single crystalsurfaces of different surface structure (26). For structure-sensitive catalysis processes the concentration of the mostactive metal sites determines the turnover rates. In thesecircumstances the availability of atoms with high coordina-tion and correspondingly large electronic fluctuations is ofkey importance. There are also, however, structure-insen-sitive reactions, such as many hydrogenation processes (27,28), in which the structure of the metal sites is not asimportant. Reactions of this type are likely to occur on thestrongly adsorbed overlayer (29) that essentially insulatesthe metal from most of the reactants and reaction productsand may reduce the metal role to that of atomizing some ofthe smaller reactant molecules (H2 or 02, for instance).As an example, during ethylene hydrogenation a stable

ethylidyne monolayer forms on top of the metal surface, andthe hydrogenation occurs via hydrogen transfer through thislayer to the ethylene molecules that are weakly adsorbed on

top of it. The rates of structure-sensitive reactions (e.g., thebreaking of the H2 molecule) are controlled by the surfaceroughness, whereas the structure-insensitive processes(e.g., the hydrogenation of ethylene by the atomic hydrogen)are mediated by the tenaciously held overlayer. It is impor-tant to realize that most catalytic reactions may require bothingredients: the proper atomic surface structure and a

strongly adsorbed overlayer.

CONCLUSIONThe catalytic activity in structure-sensitive processes isdirectly related to the existence of sites with low-energyelectronic fluctuations. These fluctuations are most pre-

dominant in transition metal atoms of high coordinationnumber that are surrounded by a large number of othermetallic nonmagnetic or weakly magnetic atoms.

This proposition immediately points out the suitability, forthis purpose, of rough, stepped, or kinked surfaces of metalsand properly chosen alloys, which make accessible to thereactants atomic sites of high coordination.

It is also important to remark that not all intermediatesteps of a catalytic reaction are structure sensitive, and thatstructure-insensitive processes (which in general are not therate-determining ones) may be in fact mediated by tena-ciously held overlayers, which insulate the metal from directcontact with some of the reactants.Our proposition has obviously profound implications in

the manufacturing of catalysts, their chemical compositionand morphology, as well as in the understanding of theprocesses that lead both to their effectiveness and to theirdemise (poisoning).

This work was supported at the Lawrence Berkeley Laboratoryby the Director, Office of Energy Research, Office of Basic EnergySciences, Materials Science Division of the Department of Energyunder Contract DE-AC03-76SF00098.

1. Davis, S. M., Zaera, F. & Somorai, G. A. (1982) J. Catal. 77,439-459.

2. Somojai, G. A. (1981) Chemistry in Two Dimensions: Sur-faces (Cornell Univ. Press, Ithaca, NY).

3. Spencer, N. D., Schoonmaker, R. C. & Somorjai, G. A.(1982) J. Catal. 74, 129-135.

4. Ascher, M. & Somorjai, G. A. (1984) Surf. Sci. 143, L389- L392.5. Salmeron, M., Gale, R. J. & Somorjai, G. A. (1977) J. Chem.

Phys. 67, 5324-5334.6. Davis, S. M., Zaera, F. & Somorjai, G. A. (1984) J. Catal. 85,

206-223.7. Toyoshima, I. & Somorjai, G. A. (1979) Catal. Rev. Sci. Eng.

19, 105-159.8. Sinfelt, J. H. (1973) Adv. Catal. 23, 91-119.9. Sinfelt, J. H. (1974) Catal. Rev. 9, 147-168.

10. Tersoff, J. & Falicov, L. M. (1981) Phys. Rev. B 24, 754-764.11. Thalmeier, P. & Falicov, L. M. (1979) Phys. Rev. B 20,

4637-4644.12. Thalmeier, P. & Falicov, L. M. (1980) Phys. Rev. B 22,

2456-2465.13. Tersoff, J. & Falicov, L. M. (1982) Phys. Rev. B 25,

2959-2961.14. Tersoff, J. & Falicov, L. M. (1982) Phys. Rev. B 26, 459-461.15. Tersoff, J. & Falicov, L. M. (1982) Phys. Rev. B 26,

6186-6200.16. Victora, R. H. & Falicov, L. M. (1983) Phys. Rev. B 28,

5232-5242.17. Victora, R. H. & Falicov, L. M. (1984) Phys. Rev. B 30,

259-262.18. Victora, R. H., Falicov, L. M. & Ishida, S. (1984) Phys. Rev.

B 30, 3896-3902.19. Victora, R. H. & Falicov, L. M. (1985) Phys. Rev., in press.20. Ertl, G. (1983) in Robert A. Welch Conference Proceedings on

Heterogeneous Catalysis, ed. Hackerman, N. (Robert A. WelchFoundation, Houston, TX), Vol. 25, pp. 179-207.

21. Zaera, F. & Somorjai, G. A. (1982) J. Phys. Chem. 86,3070-3078.

22. Gillespie, W. D., Herz, R. K., Petersen, E. E. & Somorjai,G. A. (1981) J. Catal. 70, 147-159.

23. Davis, S. M., Zaera, F. & Somorjai, G. A. (1982) J. Am.Chem. Soc. 104, 7453-7461.

24. Boudart, M. (1969) Adv. Catal. 20, 153-166.25. Dominguez, J. M. & Yacaman, M. J. (1980) in Growth and

Properties ofMetal Clusters, ed. Bourdon, J. (Elsevier, Amster-dam), pp. 493-503.

26. Somorai, G. A. (1983) in Robert A. Welch Conference Proceed-ings on Heterogeneous Catalysis, ed. Hackerman, N. (Robert A.Welch Foundation, Houston, TX), Vol. 25, pp. 166-178.

27. Blakely, D. K. & Somorjai, G. A. (1976) J. Catal. 42, 181-196.28. Manogue, W. H. & Katzer, J. R. (1974) J. Catal. 32, 166-169.29. Zaera, F. & Somorjai, G. A. (1984) J. Am. Chem. Soc. 106,

2288-2293.

_ ) - C, - C, gases

Isobutane Methane. Ethane, Propane

12-_

1I

Chemistry: Falicov and Somorjai

V.V.