You are on page 1of 14

feature

Catalysis of gold nanoparticles deposited on metal oxides


Gold in bulk is chemically inert and has often been regarded to be poorly active as a catalyst. However, when gold is small enoughwith particle diameters below 10 nmit turns out to be surprisingly active for many reactions, such as CO oxidation and propylene epoxidation. This is especially so at low temperatures. Here, a summary of the catalysis of Au nanoparticles deposited on base metal oxides is presented. The catalytic performance of Au is defined by three major factors: contact structure, support selection, and particle size, the first of which being the most important because the perimeter interfaces around Au particles act as the site for reaction.

Masatake Haruta
Research Institute for Green Technology National Institute of Advanced Industrial Science and Technology (AIST) 16-1 Onogawa, Tsukuba 305-8569, Japan tel: +81-298-61-8240 fax: +81-298-61-8240 e-mail : m.haruta@aist.go.jp

Heterogeneous catalysts currently used are classified into three types of compounds: metal oxides, metal sulfides, and metals. Metal oxides are used mainly for selective oxidation of hydrocarbons [1] and selective reduction of NOx with NH3 [2] . Metal sulfides are used for hydrodesulfurization of petroleum [3] . Metals are most widely used for a variety of reactions [4] , including hydrogenation, complete and partial oxidation, and reduction of NOx with hydrocarbons [5] . Actually, catalytic metals are limited to 12 elements of group VIII and Ib of the Periodic Table. Most widely used are the 3d metals of Fe, Co, Ni and Cu, the 4d metals of Rh, Pd and Ag, and the 5d metal of Pt. Ruthenium(4d) and Ir(5d) have only limited applications. Osmium is excluded as a catalyst component because its oxide is toxic. Gold (5d) is the only exception and has usually been regarded to be poorly active as a catalyst. The catalytic excellence of group VIII metals can be ascribed to the optimum degree of d-band vacancy. The elements of group Ib, the so called coinage metals, Cu, Ag and Au, have fully occupied d-bands. Owing to relatively low ionization potential, Cu and Ag readily lose electrons to yield d-band vacancies. In fact, Cu is used for methanol synthesis and Ag is used for ethylene oxide
102
Volume 6, no. 3, 2002

synthesis in the chemical industry. On the other hand, Au has a high ionization potential and accordingly has poor affinity towards molecules. Surface science investigations [6~8] and theoretical calculations [9] have proved that no dissociative adsorption of H 2 and O2 takes place over the smooth surfaces of Au at temperatures below 473K, indicating that Au is catalytically inactive for hydrogenation and oxidation. Indeed, the conventional supported Au catalysts were much less active than supported Pt group metal catalysts. But, it should be noted that these supported Au catalysts were not as highly dispersed as other supported noble metals. When they were prepared by the impregnation method, Au particles were usually larger than 30 nm, while Pt particles were distributed at around 5 nm in diameter[10] . This is because the melting point of Au is much lower than those of Pd and Pt (Au:1336K, Pd:1823K, Pt:2042K). Due to the quantum-size effect, the melting point of Au particles with a diameter of 2 nm is lowered to 573K [11] , and these small Au nanoparticles tend to coagulate much more readily than Pd and Pt nanoparticles during calcination of catalyst precursors at temperatures above 573K. Accordingly, we did not know whether the rough surfaces or nanoparticles of

Unraveling the bite angle effect

Figure 1 TEM micrograph of a Au/TiO2 catalyst prepared by the impregnation method.

Au, having a substantial number of steps, edges and corners, were really catalytically inactive or not. The first hint that Au might not always be poorly active when dispersed as small nanoparticles was presented by Bond and Sermon, for hydrogenation over Au/SiO2 prepared by calcination at a temperature as low as 383~401 K [12,13] , and by Paravano and his coworkers, for oxygen and hydrogen transfer reactions over Au/MgO and Au/Al 2O3 catalysts [14,15] . These landmarks, achieved prior to 1981, are reviewed by Schwank[16] . It should be kept in mind that most supported Au catalysts in the past may have been severely contaminated by Cl- and/or Na+, the amount of which depended on Au loadings. We found in 1983, and patented in 1984, that once Au is deposited as hemispherical nanoparticles on selected metal oxides, it exhibits surprisingly high catalytic activity for CO oxidation, even at 200K [17,18] . This finding has gradually evoked renewed interest in Au catalysts [1923] . The present article deals with up-to-date progress in research on the unique catalytic properties and mechanistic understandings of Au nanoparticles deposited on metal oxides and activated carbon. The ongoing and future applications of Au catalysts are also presented. Preparation of highly dispersed gold catalysts Traditional Au catalysts were prepared by the impregnation method (IMP). A metal oxide support is immersed in an aqueous solution of HAuCl4 and then water is evaporated to disperse HAuCl4 crystallites over the support surfaces. The dried precursor is calcined in air, usually at temperatures above 473K, and is often reduced in a diluted H 2 stream. In this case, as shown in Figure 1 [10] , the size of Au particles is larger than 30 nm, because the interaction of HAuCl4 crystallites with the metal oxide support is weak, and chloride remaining on the support surfaces markedly promotes the coagulation of Au particles. We have developed four techniques which can deposit Au nanoparticles on various types of metal oxides [24] : coprecipitation [17] , co-sputtering [25] , deposition-precipitation

(DP) [26] , and gas-phase grafting (GG) [27,28] . In addition to these, the amorphous alloy method [29] and the liquid phase grafting (LG) method [30,31] are also applicable. In the case of Pt group metals, calcination of catalyst precursors in air produces metal oxides which strongly interact with the support metal oxides. Also, the reduction in the H 2 stream brings about the strong interaction of metallic particles with the supports and removes Cl- as HCl. Because Au does not form stable metal oxides during calcination, in order to disperse Au as small nanoparticles preventing them from coagulationAu precursors should strongly interact with the support. The above six techniques can be classified into two categories. The first is based on the preparation of wellmixed precursorsfor example, hydroxide, oxide and metal mixtures of Au and the metal component of the supportby coprecipitation, co-sputtering and alloying, respectively. These precursor mixtures are then transformed during calcination in air into metallic Au particles strongly attached to the crystalline metal oxides, -Fe2O3, Co3O 4, and ZrO2, respectively. The second technique is to utilize the deposition or adsorption of Au compounds, for example, Au hydroxide by DP, and organogold complex by GG and LG. Scheme 1 shows a detailed procedure for DP. Due to the amphoteric properties of Au(OH)3, the pH of aqueous HAuCl4 solution is adjusted at a fixed point in the range from 6 to 10. Careful control of the concentration (around 10-3M), pH(6~10), and temperatures(323K~363K) of the aqueous HAuCl4 solution can lead to the selective deposition of
HAuCI4 aq. pH = 2-3 NaOH aq.

HAuCI4 aq. pH = 6-10

Au(OH)4-

support

Au(OH)3 /support

wahsing drying calcination (573-673K)

Au /support

Scheme 1 Flow chart of the procedure in the deposition-precipitation method. 103

Unraveling the bite angle effect Ordered hexagonal arrays Mesoporous SiO2 MCM-41 Surface Area: 1100 m2/g
Figure 2 phase grafting method.

Au(OH)3 only on the surfaces of support metal oxides without precipitation in the liquid phase. Because the precursor can be washed before drying, sodium and chloride ions are removed down to a level of about 50 ppm. One of the constraints of DP is that it is applicable only to metal oxides, the isoelectric points of which are above 5. Gold hydroxide cannot be deposited on SiO2(IEP=2), SiO2-Al2O3(IEP=1) and WO3(IEP=1). In contrast, GG using dimethyl-gold acetylacetonate is unique because it can deposit Au nanoparticles even on SiO2 and SiO2-Al2O3. Figure 2[28] shows a TEM photograph for Au nanoparticles deposited on MCM-41. With an increase in Au loading, wire-like metal rods appear, suggesting that most Au particles might be incorporated into the meso-tubes. Fine structure of gold nanoparticles Carefully prepared Au catalysts have a relatively narrow size distribution of Au particles, giving mean diameters in the range of 2 to 10 nm with standard deviation of about 30%. A major reason why Au particles remain as nanoparticles even after calcination at temperatures above 573K is the epitaxial contact of Au nanoparticles with the metal oxide supports. Gold particles always exposed its most densely packed plane, (111) plane, in contact with -Fe2O3 (110), Co3O 4 (111), and anatase TiO2 (112 ), and rutile TiO2 (110). Figure 3 shows a typical TEM image of Au particles epitaxially attached on anatase TiO2 [32] . The surface atomic configuration is better matched for a Au(111) plane sitting on the oxygen layer of anatase TiO2 than on the Ti layer. 3-dimensional nanostructure analyses by electron holography together with high-resolution TEM revealed that smaller hemispherical Au particles with a thickness under 2 nm had contact angles with the support that were below 90 (wet interface), whereas larger Au particles with a 5 nm thickness had that angle above 90 (dry interface) [33] . This difference in the wettability of Au particles may arise from the change in the electronic state of the contact interfaces with the particle size. When Au/TiO2 was calcined at temperatures above 573K, Au particles coagulated with each other forming larger particles, mostly gathered in the valleys at the junctions between the TiO2 particles [34] . Factors controlling catalytic activity and selectivity of gold Except for H2 oxidation and hydrocarbon hydrogenations, most reactions are remarkably structure-sensitive over supported Au catalysts. Two typical reactions are CO oxidation and propylene epoxidation. The oxidation of CO is the simplest reaction and has been the most intensively studied since Langmuir first presented a theory of adsorption and catalysis for this reaction[35]. It is also practically important in connection with the purification of engine exhaust gases and of hydrogen produced by steam reforming of methanol and hydrocarbons for polymer electrolyte fuel cells. The direct epoxidation of propylene is regarded as a Holy Grail of sorts because current industrial processes need two-stage reactions.

{1

Figure 3 (a) TEM micrograph of a Au/TiO2 catalyst prepared by the depositionprecipitation, and (b) its schematic model at the interface.
104

3
12 }

nm
TEM micrograph of a Au/MCM-41 catalyst prepared by the gas-

{1 11 }2
2. 33

Au

(b)
<110>

.3

TiO2 <201>

Unraveling the bite angle effect

intermezzo 1

A brief history of catalysis by gold

Until 1975, catalysis by Au was often studied by using wire, foil, film, gauze and sponge and, accordingly, at high temperatures. A landmark was the work of Wood and Wise, who demonstrated that a Au film was active for the hydrogenation of cyclohexene and 1-butene if dissociated hydrogen atoms were supplied through a Pd-Ag alloy timble. In the 1970s, G. C. Bond and P. A. Sermon, and also G. Paravano, paid attention to the size effect of Au particles on the catalytic activity. They tried to disperse small Au particles on MgO, Al2O3 and SiO2 by low-temperature calcination or by liquid-phase reduction. Some of them exhibited appreciable catalytic activities for the hydrogenation of linear alkenes, oxygen and hydrogen transfer reactions, and NO reduction with H2 at moderate temperatures. Although supported gold possessed unique selectivity, a constraint was that the catalytic activity was always inferior to those of other noble metals.

Another approach toward the size effect was bottom-up, starting from a single atom or clusters. In the late 1970s, Ozin reported that atomic Au could react with a CO+O2 matrix at a very low temperature (10K) and liberate CO2 when the temperature was raised to 30-40K. Cox reported specific numbers of atoms in small Au clusters (a few to twenty atoms) for reactions involving H2, O2 and CH4. These observations clearly showed that small gold particles could exhibit noticeable activity for hydrogenation and oxidation reactions. However, few people recognized at that time the implications of these valuable contributions. The interest in catalysis by Au was revitalized when Hutchings predicted in 1985, and subsequently confirmed, that Au would be the most active catalyst for the hydrochlorination of acetylene to produce vinyl chloride.

What characterizes Au catalysts in these two reactions are the three factors which rule the activity and selectivity; strong contact of Au particles with the support, suitable selection of the support, and size control of Au particles.
(1) Strong contact of gold particles with the support

Figure 4 shows turnover frequencies (TOFs), the reaction rate over one single surface metal atom per second, of CO oxidation at 300K over Au/TiO2 and Pt/TiO2 catalysts prepared by DP, photocatalytic deposition and IMP methods [36] . The DP method yields hemispherical metal particles with their flat planes strongly attached to the TiO2 support (see Figure 3), while photocatalytic deposition and IMP methods yield spherical particles simply loaded on the TiO2 support (see Figure 1) and, therefore, much larger particlesparticularly in the case of Au. Over Pt/TiO2, the reaction of CO with O2 can take place on the Pt surfaces and the metal oxide support is not directly involved in the reaction. This can account for why different preparation methods do not make any appreciable difference in the TOF of Pt catalysts. On the other hand, the TOF of Au/TiO2 markedly depends on the contact structure changing by four orders of magnitude. The TOF of strongly attached hemispherical Au particles exceeds that of Pt by one order of magnitude. The strong contact of Au particles is also indispensable for the epoxidation of propylene in the gas phase containing O2 and H 2 [37] . Figure 5 shows that spherical Au particles simply loaded on TiO2 (prepared by IMP) needs higher temperatures for reaction to occur and causes complete oxidation to produce CO2. The yield of H 2O is much larger than that of CO2, showing that H 2 oxidation takes place much faster than propylene oxidation. On the other hand, hemispherical Au particles strongly attached to the TiO2 support (prepared by DP) produce propylene oxide with

almost 100% selectivity at a lower temperature, 323K. The H 2 consumption is only about three times that of propylene conversion and appreciably smaller than that of spherical Au particle catalysts. The sharp contrast between the above two catalysts in CO oxidation and propylene epoxidation suggests that the reactions may take place at the perimeter interfaces around Au particles. To confirm this hypothesis, Vannice prepared an inversely supported catalyst, namely, TiO2 layers deposited on a Au substrate, and observed appreciable catalytic activity[38]. We prepared the Au/TiO2 catalyst by mechanically mixing a colloidal solution of Au particles 5 nm in diameter with TiO2 powder, then enacting calcination in air at different temperatures[111]. Calcination at 873K promotes the coagulation of Au particles to form larger particles with
1 10-1 turnover frequencies (s-1) 10-2 10-3 10-4 10-5 10-6 Au TiO2
Figure 4

gold

platinum

Au TiO2

Pt TiO2

Pt TiO2

Turnover frequencies for CO oxidation over spherical and


105

hemispherical particles of Au and Pt supported on TiO2.

10 9 8 7 6

product yield (%)

Unraveling the bite angle effect


106

CH3CH=CH2 + O2 + H2

CH3CH CH2 + H2O O impregnation Au TiO2 H2O

deposition - precipitation 5 4 H2O 3 2 1 0 310 320 CH3CH CH2 O CO2 330 340 350 reaction temperature (K) 360 Au

TiO2

yielding mainly N2 and CO2, while N2O is mainly produced over Pd and Pt catalysts even at higher temperatures. For the selective oxidation of hydrocarbons in the co-presence of O2 and H 2, as listed in Scheme 2, only TiO2 and Ti-silicates act as effective supports [4549] . The support requirements are very strict. Only anatase, neither rutile nor amorphous, makes Au selective, although the reason is still unclear. TEM observations showed that Au particles were more often epitaxially contacted on anatase than on rutile, indicating that the location of Ti cations around Au particles is more regular on the anatase surfaces [32] . When Ti cations are isolated from each other on the surface or in the bulk network of SiO2, Au is also selective to epoxidation. The distance between Ti cations may be important; on the surfaces of anatase and Ti silicate, Ti cations are separated from each other at a distance of the diameter of oxygen anion or farther, while on the rutile surface, they are located closer. For methanol synthesis, ZnO, which is also the best support for commercial Cu catalysts, works as the best support [50,51] . For NO reduction with hydrocarbons in the presence of O2 and H 2O, Al 2O3 gives the most selective reduction catalysts [52] .
(3) Size control of gold particles

Figure 5

Product yields of the reaction among propylene, oxygen and

hydrogen over Au/TiO2 catalysts prepared by the deposition-precipitation and impregnation methods.

diameters above 10 nm, but at the same time with stronger contact (observed by TEM), leading to much higher catalytic activity than calcination at 573K.
(2) Selection of suitable supports

For CO oxidation, except for strongly acidic materials such as SiO2-Al 2O3 and activated carbon, many oxides can be used as a support. Semiconductive metal oxides such as TiO2, Fe2O3, Co3O 4 and NiO provide more stable Au catalysts than insulating metal oxides like Al 2O3 and SiO2. Among Au supported on Al 2O3, SiO2 and TiO2, TOFs at room temperature are nearly equal, indicating that the contribution of metal oxide supports is more or less similar in intensities [39] . The difference appears in the moisture effect: Al 2O3 and SiO2 need a higher H 2O concentration above 10 ppm for CO oxidation at room temperature than TiO2 [40] . Alkaline earth metal hydroxides such as Ba(OH) 2 and Mg(OH) 2 are the best to exhibit high activity at a temperature as low as 196K. However, the Au catalysts are stable for only 3 or 4 months [41] . For the combustion of hydrocarbons, Co3O4, which is the most active base metal oxide for complete oxidation, gives rise to the highest activity[42,43]. For nitrogen-containing compounds, ferric oxide and nickel ferrites lead to the highest activity owing to their good affinities to nitrogen[44]. The oxidative decomposition of trimethylamine, which is a typical odor compound, proceeds at temperatures below 373K,

Figure 6 shows that at a critical diameter of 2 nm, the main product in the reaction of propylene with O2 and H2 switches from propylene oxide to propane[37]. When Au particles are smaller than 2 nm (clusters), they behave differently in the presence of O2 from the bulk Au and become similar to Pd and Pt. Over Pd and Pt catalysts the product is only propane, irrespective of the presence and absence of O2, while propane is formed over Au catalysts only when O2 is present. This implies that a change in the surface property of Au clusters is induced by O2 through electron abstraction from the Au clusters to form negatively charged oxygen species, O2-. This critical diameter corresponds to a layer of 3 or 4 atoms thick if the Au clusters are hemispherical in shape. The work by Goodman and his coworkers[53] shows that the electronic state of Au clusters changes with layers of 2 or 3 atoms thick.

313 - 473 K CH3CH=CH2 + O2 + H2 TiO2 (anatase) Ti-MCM41 Ti-, TS-1, TS-2 TiO2 (rutile) TiO2 (amorphous)

CH3CH CH2 O

CH3CCH3, CO2 O

Scheme 2 Products of the reaction among propylene, oxygen and hydrogen over gold catalysts supported on the various materials containing titanium.

Unraveling the bite angle effect

intermezzo 2

How metal support interactions affect the catalytic properties of metals

Metal support interactions can be classified into three types. The first is that metal particles work as the active phase, with the support modifying the electronic structure of the metal. This happens only when metal particles are small enough to be substantially altered by some electron transfer from or to the support. The critical size is quantum size, which is about 2.0 nm in diameter for spherical particles of noble metals. For such sizes, the fraction of atoms exposed to the surface exceeds 50%. The electronic structure of such metal particles deviates from that of bulk metal even when they are unsupported. However, the interaction with a metal oxide support, especially semiconductive metal oxides, can enhance this change through electron transfer. Goodman observed by scanning tunneling spectroscopy that the electronic state of Au islands deposited on TiO2 changes when their thickness is less than 3 atomic layers. The second type involves new reaction sites appearing at the metal-support perimeter interface. The surface of the metallic

particles is, however, still indispensable to provide sites for the adsorption of at least one of the reactants. For example, CO on the edge and corner sites of Au nanoparticles. The supports may also work as a reservoir for counter-reactants. For example, O2 on metal oxide supports. Even though molecular O2 may not dissociate by itself at the perimeter interface, it is speculated that it can react with CO at the periphery sites to form CO2. An open question is why the perimeter interface works as a reaction zone. One possibility is that the interface is composed of Au oxide or hydroxide, which is stabilized and recycled by interaction with metal oxide supports. The last type of effect is the reverse of the first case. Thin metal oxide layers covering the metal particles may work as the catalytically active phase, in particular when the layer thickness is smaller than a few atoms. This happens, for example, when Pd/CeO2 is prepared by co-precipitation or depositionprecipitation.

kCO (mmol min-1 m-2cmHg-1)

Figure 7 plots the rates of CO oxidation per exposed surface area of Au, which are equivalent to TOFs, over unsupported Au powder and over Au/TiO2 [54] . In the case of Au/TiO2, the rates were about one order of magnitude larger when measured by lowering the temperature from 353K than when measured by raising the temperature from 203K. This difference is assumed to arise from the accumulation of carbonate species on the surfaces of the support at low temperatures resulting in the loss of the activating power of the perimeter interfaces for O2. Therefore, the rate over Au/TiO2 which was deactivated during experiments at lower temperatures is regarded to be close to the rate of CO reaction with O2 over the surfaces of Au particles without the contribution of O2 activation at the perimeter interfaces. The linear relation among the rates over unsupported Au powder and the ones over deactivated
Au/TiO2 (p-25) 3 product yield (%) C3H6 353K C3H6 / O2 / H2 / Ar = 1 / 1 / 1 / 7 SV : 4,000 h-1 ml/g-cat

Au/TiO2 supports our assumption and shows that the normalized rate increases with a decrease in the mean diameter of Au particles by a factor of 2/3. This increase can be explained if the active sites are edge, corner or step sites, the fractions of which increase with a decrease in the size of Au particles [61] . The one order of magnitude difference in the rate between fresh (obtained by high temperature 101 Au/TiO2 100 273K

fresh

10-1

deactivated Au powder

10-2

CO2 CH3CH CH2 O 1.6 2.0 2.4 Au particle size (nm) 2.8

10-3 100 101 102 103 Au particle size (nm)


Figure 7 Reaction rates for CO oxidation over fresh and deactivated Au/TiO2 catalysts as well as unsupported Au powder.
107

Figure 6

Product yields of the reaction among propylene, oxygen and

hydrogen over Au/TiO2 as a function of Au particle size.

kCO (10-5 molCO/min . g-cat)

10

adsorbance

Unraveling the bite angle effect


108

Icosahedron Icos.: Cubo-oct. CO Conversion (-70C) 58 : 42

Cubo-octahedron 7 : 93

100 %

7 : 93

Figure 8 CO oxidation over Au13 clusters supported on Mg(OH) 2 with different composition of icosahedral and cubo-octahedral structures.

measurements) and deactivated Au/TiO2 can be ascribed to the contribution of the TiO2 support. Among supported noble metal catalysts, Au supported on Mg(OH) 2 is the most active for CO oxidation at 196K, giving 100% conversion at an hourly space velocity of 20,000h-1mg/g-cat. However, it suddenly dies after 3 to 4 months, losing activity even at 473K [55] . We have estimated the most probable size distribution of Au particles by using Debye Functional Analysis of X-ray scattered by Au particles for fresh and aged catalysts. It is suggested that the active Au species are 13-atom Au clusters and when they grow to 55-atom clusters with truncated decahedral structure, the catalytic activity is completely lost. Figure 8 shows that among the two structures of 13-atom clusters, the icosahedron is active, whereas the cubo-octahedron is

inactive [41] . It is hard to explain why such a drastic difference was observed. The coordination number of the Au atom is 5 for the icosahedron and 4 for the cubo-octahedron. The active Au/Mg(OH) 2 catalyst, which is mainly composed of icosahedral Au clusters of 13 atoms, showed negative apparent activation energy in the temperature range from 196K to 273K [56] . This can be explained by the enhanced transformation of the icosahedron into the cubo-octahedron with a rise in reaction temperature. Heiz and coworkers prepared model catalysts by depositing size-selected Au anion clusters onto a single crystal of MgO. Appreciable size dependency of CO adsorption was observed and the highest reactivity to CO was observed for the anion clusters of eleven atoms [57] . It was also reported that 8 and 11 are the smallest and the second smallest number of atoms to exhibit the catalytic activity for CO oxidation over the MgO support. Higher activity of Au clusters on defect-rich MgO than on defect-poor MgO was observed. Ab initio simulations indicated that partial electron transfer from the surface of the Au clusters, and oxygen-vacancy defects in the support, play an essential role for the genesis of catalytic activity [58,59] . Kinetic behaviour and reaction mechanism As typically shown in Figure 9 for Au/TiO2, the rate of CO oxidation over Au/TiO2, Au/Fe2O3 and Au/Co3O 4 is independent of the concentration of CO down to 0.1 vol% and only slightly dependent on the concentration of O2 (0~0.25 order) [18] . Actually, this fact was a surprise to us because the conversion of CO increased with a decrease in the concentrations of CO and O2. Over unsupported Au powder, with the mean diameters of primary Au particles at 17 nm, the rate is almost independent of the concentrations of CO and O2 [54] . These independencies suggest that CO and O2 are adsorbed on the catalyst surfaces nearly to saturation and the reaction of CO with O2 is the rate determining step.

100 273 K CO

2176

4 mbar CO at 90 K

O2

.5A

Au/TiO2 3.3 wt%

2098

dAu 573K (2.5 nm)

2154

473K (2.4 nm) 873K (10.6 nm)

rate [CO]0.05[02]0.24 0.1 0.1 1 10


Figure 10 473, 573 and 873 K. 2200 2100 wavenumbers (cm-1)

2000

Concentration of CO or O2 (vol%)
Figure 9 Dependence of the reaction rates for CO oxidation over Au/TiO2 on the concentration of reactant gasses.

FT-IR spectra for CO adsorbed on Au/TiO2 calcined at

Unraveling the bite angle effect

101 region I 100 Ea 2 kJ/mol

CO + Au

O C-Au Au/TiO2...O2 O2 (g) slow

O2 + Au/TiO2 k CO (mmol min-1 m-2cmHg-1) Au/TiO2...O2 + 2 [O C-Au] Au/TiO2 10-1 region II Ea 30 kJ/mol O C-Au O

O C-Au + CO2 O CO2

O C
perimeter

surface

10-2 Ea 10-3 3 4 1000/T (K-1) Figure 11 Dependence of the reaction rates for CO oxidation over Au/TiO2 as well as unsupported Au powder on the reaction temperature. 5 5 kJ/mol Au powder

spectator intermediate

O C

O
region III

C O O
Au TiO2

O C O

Scheme 3 Reaction mechanisms proposed for CO oxidation over Au/TiO2.

Figure 10 shows FT-IR spectra for CO adsorption at 90K over Au/TiO2 calcined in air at different temperatures [60] . The most active sample (calcined at 573K, mean diameter of Au particles 2.5 nm) has the largest intensities of the peak at 2110~2120cm-1, which can be assigned to CO linearly adsorbed on the metallic Au sites. When Au particles become larger than 10 nm in diameter (sample calcined at 873K), the intensity of this peak is markedly reduced, indicating that CO adsorption may take place only on steps, edges and corners of Au particles, not on the smooth surfaces. This agrees well with what is obtained by self-consistent density functional theory calculations [61] . No direct experimental evidence has yet been presented where oxygen is activated for reacting with CO adsorbed on the Au surfaces, and whether the oxygen molecule is dissociatively or non-dissociatively adsorbed. A TAP (temporal analysis of products) study of O2 adsorption and the reaction of O2 with CO [62,63] , 18O2 isotope experiments [62-64] and ESR measurements [64,65] , indicate that molecularly adsorbed O2, most likely O2- at the perimeter interface, is involved in the oxidation of CO. Figure 11 shows the Arrhenius plots in a wide range of temperatures from 90K to 400K [66] . There are three temperature regions where different kinetics are operating with markedly different rates and apparent activation energies. The probable pathways for CO oxidation are schematically shown in Scheme 3. At temperatures below 200K, the TiO2 surfaces and its perimeter interfaces around Au particles are covered with carbonate species formed by surface reactions of CO. Reaction of CO with O2 takes place only on the surfaces of Au, more specifically, on the step, edge and corner sites, with apparent

activation energies of almost zero kJ/mol. This means that when Au particles are small enough, the catalytic activity can be detected at any temperature. Actually, unsupported Au powder (primary particles with mean diameters of 17 nm) exhibits measurable activity for CO oxidation at 200K with an apparent activation energy of nearly zero[54]. At temperatures above 300K, reaction takes place at the perimeter interfaces between CO adsorbed on the surfaces of Au particles and oxygen (most likely molecular) adsorbed at the support surfaces. This reaction also gives nearly zero apparent activation energy, but proceeds much faster by more than one order of magnitude than the reaction over the Au surfaces. At intermediate temperatures from 200K to 300K, the reaction proceeds at the perimeter interfaces, which are partly covered with carbonate species. The coverage of the species may change depending on temperature, thus giving rise to an apparent activation energy of around 30 kJ/mol. There are still other arguments about the active species of Au, especially in the case of Au/Fe2O3: oxidized Au species, Au(I)[67], Au+[68] or metal oxide support surfaces with modified reducibility by the interaction with Au nanoparticles[69,70]. It is unlikely that oxidic Au species are major catalytically active phases because the most active supported Au catalysts are prepared by calcination in air at 573 K, where Au precursors (hydroxides or organo complexes) can be transformed mostly into metallic particles. A certain fraction of Au species remains as atomically dispersed species in the matrix of the support, which was proved by EXAFS[22,68,71,72], XPS[69], Mssbauer[70,73] and IR for CO adsorbed[60,67]. However, no correlation between the amount of oxidic Au species and catalytic activity has yet been presented. It is speculated that
109

conversion (mol%)

Unraveling the bite angle effect


110

the samples mainly consisting of oxidic Au could exhibit high catalytic activity because oxidic Au species are transformed into metallic particles during reaction. Even though metallic Au particles are indispensable, a question arises as to why the periphery of Au particles can activate O2 molecules at low temperatures. As proposed by Bond and Thompson[74], it is probable that the perimeter interfaces contain oxidic Au(III) or Au(I) species, most probably Au(OH)3 or Au(OH) under usual conditions, where H2O is present at a concentration of 10 ppm. These hydroxides may be stabilized and reversibly formed and decomposed with the aid of the metal oxide supports. Another argument proposed by Goodman and his coworkers is that the non-metallic nature of Au clusters leads to the highest catalytic activity[53]. This conclusion is questionable. The transition of electronic state was measured on one specific Au cluster of a defined diameter by scanning tunneling spectroscopy, whereas the catalytic activity was measured by using a specimen of Au/TiO2 model catalyst with a mean diameter of all Au clusters. A maximum in catalytic activity, with respect to the mean diameter of Au clusters, was observed where the transition of electronic state from metallic to non-metallic started. However, this fact can be more reasonably explained by assuming that a maximum in the total surface area or the number of step sites[61] of metallic Au clusters is obtained at a certain diameter where the transition to the non-metallic state starts. Reactions catalyzed by gold Supported Au catalysts can also catalyze many reactions other than CO oxidation and propylene epoxidation. The following reactions usually take place at much lower temperatures or with much higher selectivities than over the other metal catalysts.
(1) Water Gas Shift Reaction and CO Removal from H2

(2) Hydrogenation of Unsaturated Hydrocarbon

A characteristic feature of Au catalysts for the hydrogenation of unsaturated hydrocarbons is that partial hydrogenation takes place very selectively: butadiene to butene and acetylene to ethylene[12,13,82,83]. Hydrocarbon hydrogenations are known to be structure-insensitive, proceeding at approximately the same TOF on metal particles of various sizes[84]. It was also the case in the hydrogenation of butadiene over Au catalysts as far as Al2O3, SiO2 and TiO2 were concerned as supports[82]. Turnover frequency (TOF) differed only by a factor of three among Au particles with mean diameters ranging from 3.5 to 7 nm which were deposited on Al2O3, SiO2 and TiO2. These Au catalysts were perfectly selective to partial hydrogenation to form butene(1-butene 70%, 2-butene 30% at 323K), whereas other noble metal catalysts by-produced butane. In the cases of acetylene hydrogenation over Au/Al2O3 and acrolein hydrogenation over Au/TiO2 and Au/ZrO2, the catalytic activity per unit weight of Au increased with a decrease in the size of Au particles down to 2 nm, indicating that the metallic nature of Au is also important for hydrogenation. In the hydrogenation of , unsaturated aldehyde, selectivity to the hydrogenation of C=O versus that of C=C was reported to reach 40~50% when Au particles deposited on TiO2 or ZrO2 were larger than 2 nm in diameter[85,86]. It has been reported recently that Au/ZnO exhibits a selectivity of about 80% to C=O hydrogenation to produce the unsaturated alcohol, but-2-en-1-ol, from but-2-enal[87].
(3) Liquid Phase Oxidation and Hydrolysis

Owing to ongoing applications of polymer electrolyte fuel cells to automobiles and to residential electricity-heat delivery systems, low-temperature water gas shift reaction is attracting renewed interest. In comparison with commercial catalysts based on Ni or Cu, which operate at 900K or at 600K respectively, supported Au catalysts appear to be advantageous in operation at a temperature as low as 473K. As a support, TiO2[75] and ZrO2[76] (IV A group metal oxides), and Fe2O3 [76] are especially effective, and the crystallinity of these metal oxides appreciably affect the catalytic activity[77]. Although the stability may not be good, Au/NaY is reported to be active for a water gas shift reaction at 373K[78]. The redox cycle of Au+ and Au0 is assumed to operate for CO activation, while the NaY support activates H2O. For selective CO removal in an H2 stream, supported Au catalysts are potentially advantageous over other noble metal catalysts because only supported Au catalysts are much more active for CO oxidation than H2 oxidation. Au/Mn2O3[79] and Au/Fe2O3[80,81] in particular exhibit good stability as well as high catalytic activity and selectivity.

Prati and Rossi have found that Au supported on activated carbon is more active and selective than other noble metal catalysts for the oxidation in a Methanol-H2O(6:4) solvent of glycols to -hydroxy acids, which are used in the cosmetics and food industries[88]. Not only activated carbon, but also -Al2O3 and TiO2 make Au active and selective[88,89]. An interesting feature in Figure 12 is that over activated carbon

100 80 60 40 20 0 3 4 5 6 7 Au particle size (nm) 8 9 Au/C

Au/Al2O3

Figure 12 Dependence of the conversion for glycol oxidation over Au/Al2O3 and Au/C catalysts on the Au particle size [88] .

Unraveling the bite angle effect

support, a maximum activity is observed when the mean diameter of Au particles is 7-8 nm, whereas the smaller Au particles give higher activity over -Al2O3 and TiO2 support. This can be explained as follows: smaller Au particles can be easily fixed on the internal surfaces of carbon, and are consequently less accessible to reagents in the liquid phase than larger Au particles located on the external surfaces. It should also be noted that Au sols stabilized with polyvinylpirrolidone or tetrakis(hydroxymethyl)phosphonium chloride can give uniformly dispersed Au particles even on activated carbon by simply dipping the support materials. Unsupported Au nanoparticles also exhibit unique catalytic properties in liquid phase reaction. N-methylimidazolefunctionalized Au particles are active for the hydrolysis of a carboxylic acid ester to about 30 times as high as a reference monomeric catalyst without Au particles[90].
(4) NO Reduction with Hydrocarbons

dioxin decomposition below 423 K


Cl O O Cl CO HCl HCHO etc. CO2

carrier Au Fe2O3 Pd SnO2 Ir La2O3

Figure 13

Schematic structure model of the three-component integrated

catalyst for dioxin decomposition.

The reduction of NO with hydrocarbons to N2 in the co-presence of excess O2 takes place over some supported Au catalysts[52]. Alkenes (C2H4, C3H6) are more effective as a reductant than alkanes (CH4, C2H6, C3H8) because the former can adsorb on the Au surfaces more strongly. The temperature and efficiency for NO reduction depend on the kind of metal oxide supports and increase in the order of ZnO(523K, 49%), -Fe2O3(523K, 12%), MgO(623K, 42%), TiO2(623K, 30%), Al2O3(673-723K, 80%). The NO conversion to N2 obtained over Au/Al2O3 in the presence of 5 vol% O2 and 10 vol% H2O is comparatively higher than over the other catalysts reported so far[5]. The reaction passes through NO oxidation with O2 to form NO2, which then reacts with propylene. Therefore, enhanced activity is obtained for a mechanical mixture of Au/Al2O3 with MnOx which is active for NO oxidation to NO2[91]. Gold wire[92] and film[93] were known to be active for N2O decomposition to N2 and O2. Recently, Au/Co3O4 has been reported to be active for this reaction even at 523K in the co-presence of 10 vol%O2 and 5 vol%H2O[94].
(5) Reactions involving Halogens

Because of the stability of Au against halogens, some supported Au catalysts have been reported to be more active and stable for reactions involving halogens than other noble metal catalysts. Gold supported on activated carbon is the most active for the hydrochlorination of acetylene to produce vinylchloride[95]. Gold on Co3O4 or Al2O3 is active for the oxidative decomposition of CCl2F2 and CH3Cl[96]. Gold supported on LaF3 is active for the synthesis of HCN by the reaction of fluorinated hydrocarbons with NH3[97]. Ongoing and future applications The characteristic features of supported gold catalysts are: active at low temperatures, activated by moisture and often very selective. Gold deposited on Fe2O3, which is supported on a zeolite wash-coated honeycomb, has been commercially

used as an odor eater in modern Japanese toilets since 1992. Gold catalysts are more active at temperatures below 400K for the oxidative decomposition of trimethlamine, a typical odor compound, than Pd and Pt catalysts, and more selective to N2 with the least formation of N2O[44]. It is probable that odor eaters using gold catalysts will expand their applications to air cleaners, home garbage treatment equipments, volatile organic vapor treatments and so on. Gold supported on Fe2O3 or La 2O3 is the most active among noble metal catalysts for the oxidative decomposition of dioxin at temperatures below 473K. Recently, as shown in Figure 13, we have integrated Ir, Pt and Au catalysts supported on La 2O3, SnO2 and TiO2, respectively, and 95% decomposition of dioxin was obtained, even at 423K, at an hourly space velocity of 12,000h-1ml/g-cat. [98] . Owing to the booster effect of Ir catalysts, the integrated noble metal catalysts are applicable to the removal of dioxin from the outlet gases of incinerators. The third promising field of application is in CO2 lasers [99] . It has long been known that coating the glass inside-walls of a laser discharge tube with Au results in a substantial improvement in the performance of sealed-off direct current excited CO2 lasers [100] . The performance of radio frequency excited CO2 lasers can also be enhanced by a joint action of the Au coating of electrodes and the addition of CO at a high concentration(4~8 vol%) [101,102] . The role of CO is explained by the adsorption on the Au surfaces to prevent them from being deactivated by oxygen species. This phenomenon is consistent with the results obtained at 90K for the reactivities of CO and oxygen preadsorbed; the latter is more stably adsorbed and less reactive to form CO2 [60] . It will take longer to utilize Au catalysts in the chemical industry. There are possibilities of earlier commercialization in selective partial oxidation in liquid phase for fine chemicals. As for bulk chemical production, the direct epoxidation of propylene can be commercialized if the yield of propylene oxide (PO) is improved to 10% from the present level of 3%, and the efficiency of H2 is improved to 50%, at least, from 20~30%. Gold supported on anatase TiO2 is selective only at temperatures below 353K. Because this temperature is not sufficiently higher than the boiling point of PO (307K),
111

Unraveling the bite angle effect


112

the conversion of propylene remains small due to the slow desorption of PO. Therefore, a variety of titanium silicates as well as Ti deposited on SiO2 have been tested as a support for Au particles[45-49,103]. Among titanosilicates such as TS-1, Ti- zeolite, Ti-MCM41 and Ti-MCM48, having different sizes and structures of pores, Ti-MCM48 is proven to give the best catalytic performance[104]. The atomic ratio of Ti/Si should be within 1/100~3/100 to keep Ti cations isolated from each other in the matrix. When these Ti/SiO2 support materials are used, Au is selective to epoxidation up to 473K, yielding improved conversion almost to 5%[47,103,104]. A reaction pathway has been proposed in which H2O2 is formed on the Au surfaces and then transformed into OOH on Ti4+ tetrahedrally coordinated in SiO2 matrices to react there with propylene[47,49,104]. Outlook Gold catalysts have been attracting growing interest from the point of view of both applications and fundamental study [23] . Owing to their excellent performance in complete oxidation at ambient temperature with enhancement by moisture,

Au catalysts are expected to penetrate into our daily life and make it more comfortable and healthy. Through appropriate selection of support materials, gold can exhibit high selectivity in hydrogenation and partial oxidation, and may contribute to the development of highly selective and green chemical processes. The genesis of catalysis by gold is, in most cases, ascribed to the perimeter interfaces around Au particles. This presents us with a new guiding principle to create a wide range of new catalytic systems, because the combination of catalytic metals with a variety of support materials gives rise to a great deal of novel catalysts. Good examples are Pd/CeO2 and Pt/ZrO2, prepared by coprecipitation for the lowtemperature decomposition and synthesis of methanol[105]. Recently, fundamental work has been emerging to understand the unique catalysis of Au and its sizedependency. By means of surface science techniques using well-defined catalytic materials, most often size-selected gold clusters deposited on single crystalline metal oxides such as MgO[57] and TiO2[53,106,107]. An atomic scale understanding is being accumulated in combination with

questions and answers

When and how did you get involved in catalysis by gold? In 1976 when the Japanese chemical industry was suffering heavily from the second oil crisis and pollution problems, fortunately, I was offered a permanent position at the Osaka National Research Institute where I was involved in the development of hydrogen catalytic combustors under a national project for new energy technologies. I planned to develop inexpensive metal oxide catalysts that could replace Pd and Pt. Firstly, I prepared transition metal oxides by a variety of methods and then mixed metal oxides. A few years later, stimulated by a publication of Professor Moro-oka, I plotted the catalytic activities of metal oxides for H2 oxidation against the metal-oxygen (M-O) bonding energies and obtained a volcano-like relationship. My idea was then to combine Ag or Au, having weaker M-O bonding energies than those of Pd and Pt, with other metals having stronger M-O bonding energies. I observed that some Ag composite oxides, Ag-Mn and Ag-Mn-Co oxides in particular, were more active than Ag oxide alone and showed an activity comparable to that of Pd and Pt for H2 oxidation. For CO oxidation, the composite oxides were active at room temperature and superior to Pd and Pt. I presented this work at the 3rd International Symposium on Catalyst Preparation held at Louvain-laNeuve in 1982. Dr. C. S. Brooks, who had retired from

United Technology Co. Ltd., asked me whether I had prepared the composite oxides of Au. I answered that Au composite oxides might not be interesting because Au was (at that time) more expensive than Pt. He changed my mind by saying, The cost does not matter at this stage of research. It is important to find evidence and confirm your hypothesis that active catalysts may emerge from Ag or Au. In the USA, even researchers working for industry recognize a significance in checking a hypothesis. Why dont you demonstrate that Au becomes catalytically active when combined with the oxides of base metals? The active sites responsible for the catalytic activity of gold are located at either the accessible periphery of the metal particles where the supporting oxide may also contribute directly, or on the surface of the gold particles whose properties are altered by the strong interaction with the support. How does one or the other situation prevail depending on the reaction type? When gold is deposited as nanoparticles on a variety of metal oxides and carbonaceous materials, it exhibits noticeably high catalytic activity for many reactions. Among them, CO oxidation and propylene epoxidation are structure sensitive reactions, the rate and reaction products of which are markedly dependent on the size of the Au particles, the type

Unraveling the bite angle effect

theoretical calculations[9,58,61]. In this approach, gold may be the best target of research because gold itself is poorly active, and the detectable activity is obtained only when the specific structure and size of Au particles are given. What is also advantageous in Au model catalysts is that the activity can be often observed at lower temperatures where designed structure can be maintained. In order to make recent attempts more effective to combine experimental work using real catalysts and model catalysts with theoretical calculations [108] , the effect of moisture should be taken into account. Most surface science work is carried out in an ultra-high vacuum, while catalytic activity measurements for real catalysts are carried out in a fixed bed flow reactor using reactant gas containing moisture at least 1 ppm, usually 10 ppm. The catalyst surfaces are covered with OH groups and water molecules at room temperature. For CO oxidation which does not produce H 2O and proceeds at temperatures below 373K, moisture markedly changes the catalytic activity of metal oxides and gold catalysts. Under dry conditions, with a H 2O concentration of 80 ppb, CO oxidation can take place even

at 210K over Co3O 4 without Au [109,110] , while supported Au catalysts prefer moisture as shown in Figure 14 [40] .

10-5 k CO (mol/l s g-cat)

Au/TiO2

10-6

10-7 270 K 0.1 1 10 100 1000 10000

H2O concentration (ppm) Figure 14 Dependence of the reaction rates for CO oxidation over Au/TiO2 on the H2O concentration in the reactant gas.

of support and the structure of the interface. In these cases, the periphery is working as reaction sites but metallic gold particles are also necessary as sites for the adsorption of one of the reactants. On the other hand, the hydrogen oxidation and the hydrogenation of alkenes and alkynes are structure insensitive reactions whose rates in terms of turnover frequency and product selectivity are almost independent of the size of the Au particles (in the range of 2 nm to 10 nm) and of the type of support. This is because these reactions take place on the surface of the Au particles, the electronic state and defect density may change markedly at a critical size around 2 nm. The activity of gold catalysts has been reported mainly for reactions occurring below 500K. Do you know of any example where gold is active at a higher temperature? One of the characteristic features of catalysis by Au is that the apparent activation energy is very low, usually below 40 kJ/mol. and sometimes even below 10 kJ/mol. This feature may be a reflection of the weak interaction of the reactants with stepped or defect sites of Au. Such low activation energies mean that Au catalysts are more competitive at lower temperatures and especially useful for environmental applications at ambient temperatures. Consequently, Au catalysts are, in principle, inferior to other noble metal

catalysts at temperatures above 500K. The only known exception, presently, is the selective reduction of NOx with hydrocarbons. Gold supported on Al 2O3 is active and selective for NO reduction with propylene in the presence of excess O2 and H 2O at temperatures from 400K to 700K. An advantageous feature is that the conversion of NO does not decrease with increasing temperature (in this range) because the combustion of propylene is less promoted over Au/Al 2O3. Because the melting point of Au is much lower than those of Pd and Pt, and can be as low as 600K when the particle diameter is 2 nm or less, it appears that the upper reaction temperature limit for Au catalysts (which usually contain Au particles with diameters of 3-4 nm) should range from 600K to 700K.

113

Unraveling the bite angle effect


114

curriculum vitae

references
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] G. I. Panov, CATTECH, 4 (2000) 18. P. Forzatti, Appl. Catal. A: General, 222 (2001) 221. H. Topse, B. S. Clausen, and F. E. Massoth, Hydrotreating Catalysis, Science and Technology, Springler, Berlin, 1996. V. Ponec and G. C. Bond, Catalysis by Metals and Alloys, Elsevier, Amsterdam, 1996. M. Iwamoto, Stud. Surf. Sci. Catal., 130 (2000) 23. A. G. Sault, R. J. Madix, and C. T. Campbell, Surf. Sci., 169 (1986) 347. N. Saliba, D. H. Parker, and B. E. Koel, Surf. Sci., 410 (1998) 270. J. Wang and B. E. Koel, J. Phys. Chem. A102 (1998) 8573. B. Hammer and J. K. Nrskov, Nature, 376 (1995) 238. K. Tanaka, T. Hayashi, M. Haruta, Interf. Sci. Material Interconnection, Proc. JIMIS-8, Jpn. Inst. Metals., 1996, pp.547550. Ph. Buffet and J-P. Borel, Phys. Rev. A, 13 (1976) 2287. G. C. Bond and P. A. Sermon, Gold Bull., 6 (1973) 102. G. C. Bond and P. A. Sermon, J. C. S. Chem. Comm., (1973) 444. D. Y. Cha and G. Parravano, J. Catal. 18 (1970) 200. S. Galvano and G. Parravano, J. Catal. 55 (1978) 178. J. Schwank, Gold Bull., 16 (4) (1983) 103. M. Haruta, N. Yamada, T. Kobayashi, and S. Iijima, J. Catal., 115 (1989) 301. M. Haruta, S. Tsubota, T. Kobayashi, H. Kageyama, M. J. Genet, and B. Delmon, J. Catal., 144 (1993) 175. M. Haruta, Catal. Today, 36 (1997) 153. M. Haruta, Catal. Surveys of Japan, 1 (1997) 61. G. C. Bond and D. T. Thompson, Catal. Rev. Sci. Eng., 41 (1999) 319. Osaka National Research Institute, Activity Report No. 393 (1999). M. Haruta and M. Dat, Appl. Catal. A: General, 222 (2001) 427. M. Haruta, S. Tsubota, and M. Okumura, in Advances in Catalyst Preparation (Japanese), eds. Y. Ono et al., Association for the Promotion of Catalyst Preparation Chemistry, Tokyo, 2000, pp39-50. [25] [26] [27] [28] [29] [30] [31] [32] [33] T. Kobayashi, M. Haruta, S. Tsubota, and H. Sano, Sensors and Actuators, B1 (1990) 222. S. Tsubota, M. Haruta, T. Kobayashi, A. Ueda, and Y. Nakahara, Stud. Surf. Sci.Catal., 63 (1991) 695. M. Okumura, K. Tanaka, A. Ueda, and M. Haruta, Solid State Ionics, 95 (1997) 143. M. Okumura, S. Tsubota, M. Iwamoto, and M. Haruta, Chem. Lett., (1998) 315. M. Shibata, N. Kuwata, T. Masumoto, and H. Kimura, Chem. Lett., (1985) 1605. Y. Yuan, A. P. Kozlova, K. Asakura, H. Wan, K. Tsai, and Y. Iwasawa, J. Catal., 170 (1997) 191. M. Okumura and M. Haruta, Chem. Lett., (2000) 396. T. Akita, K. Tanaka, S. Tsubota, and M. Haruta, J. Electron Microscopy, 49 (2000) 657. M. Ichikawa, T. Akita, M. Okumura, K. Tanaka, and M. Haruta, Proc. 7th Intern. Symp. Advanced Physical Fields, Tsukuba, Nov. 2001, T. Noda ed., Nat. Inst. Materials Sci., pp 369-372. [34] [35] [36] [37] [38] [39] [40] [41] [42] T. Akita, P. Lu, S. Ichikawa, K. Tanaka, and M. Haruta, Surf. Interface Anal., 31 (2001) 73. I. Langmuir, J. Amer. Chem. Soc., 40 (1918) 1361. G. R. Bamwenda, S. Tsubota, T. Nakamura, and M. Haruta, Catal. Lett., 44 (1997) 83. T. Hayashi, K. Tanaka, and M. Haruta, J. Catal., 178 (1998) 566. Z. M. Liu and M. A. Vannice, Catal. Lett., 43 (1997) 51. M. Okumura, S. Nakamura, S. Tsubota, T. Nakamura, M. Azuma, and M. Haruta, Catal. Lett., 51 (1998) 53. M. Dat and M. Haruta, J. Catal. 201 (2001) 221. D. A. H. Cunningham, W. Vogel, H. Kageyama, S. Tsubota, and M. Haruta, J. Catal., 177 (1998) 1. R. D. Walters, J. J. Weimer, and J. E. Smith, Catal. Lett., 30 (1995) 181. M. Haruta, Now and Future, 7 (1992) 13. [43]

Masatake Haruta is Director of the Research Institute for Green Technology at the National Institute of Advanced Industrial Science and Technology. In April 2001, he moved to Tsukuba from Ikeda, Osaka prefecture, due to a reorganization of Japanese national laboratories into semi-autonomous bodies. He graduated from Nagoya Institute of Technology, Department of Industrial Chemistry, in 1970, and received his PhD degree from Kyoto University in 1976 on the electrochemical fluorination of organic compounds. He then was granted a tenure position at Osaka National Research Institute (ONRI), at the Agency of Industrial Science and Technology (AIST), where he was responsible for the development of hydrogen-fueled catalytic space heaters. From 1981 until 1982 he was a visiting scientist at the Universit Catholique de Louvain, Belgium, and then joined, in 1984, the headquarters of the Sunshine Project at AIST, Ministry of International Trade and Industry. In 1990 he was promoted to Head of the catalysis section and in 1994 he became a chief senior researcher organizing a new interdisciplinary basic research laboratory for research exploring the potential of gold catalysts. In 1994 he was appointed as a guest professor at the Technical University of Vienna and as an adjunct professor of the graduate school of Osaka University, Faculty of Science (until 2001). He became Director of the Energy and Environment department of ONRI in 1999. From 2000 to this date, he has also been an adjunct professor at the Collaborative Research Centre for Advanced Science and Technology at Osaka University. Harutas research interests have moved from base metal oxides for catalytic combustion to the preparation of homodispersed particles, gas sensors using noble metals deposited on semiconductive metal oxides, and eventually to the application of gold catalysts. Currently, his research group at Ikeda investigates and develops further the fundamentals and applied aspects of catalysis by gold nanoparticles.

acknowledgement

The author is grateful to Dr. M. Dat of AIST Kansai for his critical reading and comments together with the drawing of figures and schemes.

Unraveling the bite angle effect

[44] [45] [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] [80]

A. Ueda and M. Haruta, Shigen Kankyou Taisaku (Resources and the Environmental Technology), 28 (1992) 1035. M. Haruta, B. S. Uphade, S. Tsubota, and A. Miyamoto, Res. Chem. Intermed., 24 (1998) 329. Y. A. Kalvachev, T. Hayashi, S. Tsubota, and M. Haruta, J. Catal., 186 (1999) 228. B. S. Uphade, Y. Yamada, T. Nakamura, and M. Haruta, Appl. Catal. A: General, 215 (2001) 137. E. E. Stangland, K. B. Stavens, R. P. Andres, and W. N. Delgass, J. Catal., 191 (2000) 332. G. Mul, A. Zwijnenburg, B. Linden, M. Makkee, and J. A. Moulijn, J. Catal., (2001) 3239. H. Sakurai and M. Haruta, Appl. Catal. A: General 127 (1995) 93. A. Baiker, M. Kilo, M. Maciejewski, S. Menzi, and A. Wokaun, Proc. 10th Intern. Congr. Catal. (L. Guzci et al. eds.), Elsevier, Amsterdam, (1993) 1257. A. Ueda and M. Haruta, Appl. Catal. B: Environmental, 285 (1996) 81 and Gold Bull., 32 (1999) 3. M. Valden, X. Lai, and D. W. Goodman, Science, 281 (1998) 1647. Y. Iizuka, T. Tode, T. Takao, K. Yatsu, T. Takeuchi, S. Tsubota, and M. Haruta, J. Catal., 187 (1999) 50. W. Vogel, D. A. H. Cunningham, K. Tanaka, and M. Haruta, Catal. Lett., 40 (1996) 175. D. A. H. Cunningham, W. Vogel, and M. Haruta, Catal. Lett., 63 (1999) 43. U. Heiz and W.-D. Schneider, J. Phys. D: Appl. Phys., 33 (2000) R85. S. Abbet, U. Heiz, H. Hkkinen, and U. Landman, Phys. Rev. Lett., 86 (2001) 5950. U. Heiz, A. Sanchez, S. Abbet, and W-D. Schneider, J. Am. Chem. Soc., 121 (1999) 3214. F. Boccuzzi, A. Chiorino, M. Manzoli, P. Lu, T. Akita, S. Ichikawa, and M. Haruta, J. Catal., 202 (2001) 256267. M. Mavrikakis, P. Stoltze, and J. K. Nfrskov, Catal. Lett., 64 (2000) 101. M. Olea, M. Kunitake, T. Shido, and Y. Iwasawa, Phys. Chem. Chem. Phys., 3 (2001) 627. M. M. Schubert, S. Hackenberg, A. C. van Veen, M. Muhler, V. Plzak, and R. J. Behm, J. Catal., 197 (2001) 113. H. Liu, A. I. Kozlov, A. P. Kozlova, T. Shido, KL. Asakura, and Y. Iwasawa, J. Catal., 185 (1999) 252. M. Okumura, J. M. Coronado, J. Soria, M. Haruta, and J. C. Conesa, J. Catal., 203 (2001) 168174. M. Haruta, M. Dat, Y. Iizuka, and F. Boccuzzi, Shokubai, 43 (2001) 125. S. Minic, S. Scir, C. Crisafulli, A. M. Visco, and S. Galvagno, Catal. Lett., 47 (1997) 273. Z. Hao, L. An, H. Wang, and T. Hu, React. Kinet. Catal. Lett., 70 (1) (2000) 153. D. Horvth, L. Toth, and L. Guczi, Catal. Lett., 67 (2000) 117. F. E. Wagner, S. Galvagno, C. Milone, A. M. Visco, L. Stievano, and S. Calogero, J. Chem. Soc., Faraday Trans., 93 (1997) 3403. H. Kageyama, N. Kamijo, T. Kobayashi, and M. Haruta, Physica B158 (1989) 183. S. Tsubota, D. A. H. Cunningham, Y. Bando, and M. Haruta, Stud. Surf. Sci. Catal., 91 (1995) 227. Y. Kobayashi, S. Nasu, S. Tsubota, and M. Haruta, Hyperfine Interactions, 126 (2000) 95. G. C. Bond and D. T. Thompson, Gold Bull., 33 (2000) 41. H. Sakurai, A. Ueda, T. Kobayashi, and M. Haruta, J. Chem. Soc. Chem. Commun., (1997) 271. D. Andreeva, V. Idakiev, T. Tabakov, and A. Andreev, J. Catal., 158 (1996) 354. T. Tabakova, V. Idakiev, D. Andreeva, I. Mitov, Appl. Catal. A: General, 202 (2000) 336. M. Mokhtar, T. M. Salama, and M. Ichikawa, J. Colloid Interface Sci., 224 (2000) 336. R. M. Torres Sanchez, A. Ueda, K. Tanaka, and M. Haruta, J. Catal., 168 (1997) 125. M. J. Kahlich, H. A. Gasteiger, and R. J. Behm, J. Catal., 182 (1999) 430.

[81] [82] [83] [84] [85] [86] [87]

M. M. Schubert, S. Hachenberg, A. C. van Veen, M. Muhler, V. Plzak, and R. J. Behm, J. Catal., 187 (2001) 113. M. Okumura, S. Nakamura, and M. Haruta, to be submitted. J. Jia, K. Haraki, J. N. Kondo, K. Domen, and K. Tamaru, J. Phys. Chem., B 104 (2000) 11153. Z. Xu, F.-S. Xiao, S. K. Purnell, O. Alexeev, S. Kawi, S. E. Deutsh, and B. C. Gates, Nature, 372 (1994) 346. P. Claus, A. Brckner, C. Mohr, and H. Hofmeister, J. Am. Chem. Soc., 122 (2000) 11430. C. Mohr, and H. Hofmeister, M. Lucas, and P. Clause, Chem. Eng. Technol., 23 (2000) 4. J. E. Bailie, H. A. Abdullah, J. A. Anderson, C. H. Roechester, N. V. Richardson, N. Hodge, J-G. Zhang, A. Burrows, C. J. Kiely, and G. J. Hutchings, Phys. Chem. Chem. Phys., 3 (2001) 4113.

[88] [89] [90] [91] [92] [93] [94] [95] [96] [97] [98] [99]

C. Bianchi, F. Porta, L. Prati, and M. Rossi, Topics in Catal., 13 (2000) 231. F. Porta, L. Prati, M. Rossi, S. Coluccia, and G. Martra, Catal. Today, 61 (2000) 165. L. Pasquato, F. Rancan, P. Scrimin, F. Mancin, and C. Frigeri, J. Chem. Soc. Chem. Commun., (2000) 2253. A. Ueda and M. Haruta, Appl. Catal. B: Environmental, 18 (1998) 115. C. N. Hinshelwood and C. R. Prichard, Proc. Roy. Soc. London, 108A (1925) 211. V. M. Stepanov, V. D. Yagodovskii, and H. Agilar, Russian J. Phys. Chem., 49 (1975) 1335. L. Yan, X. Zhang, T. Ren, H. Zhang, X. Wang, and J. Suo, Chem. Comm., (2002) 860. B. Nkosi, M. D. Adams, N. J. Coville, and G. J. Hutchings, J. Catal., 128 (1991) 333, 378. T. Aida, R. Higuchi, and H. Niiyama, Chem. Lett., (1990) 2247. Y. Takita, T. Imamura, Y. Mizuhara, Y. Abe, and T. Ishihara, Appl. Catal. B: Environmental, 1 (1992) 79. M. Okumura, M. Haruta, X. Wang, O. Kajikawa, and O. Okada, Abstract, 3rd Intern. Conf. Environmental Catal., Tokyo, Dec. 2001, pp.1516. NASA Conf. Publ. No. 3076, Low-Temperature CO-Oxidation Catalysts for Long-Life CO2 Lasers, D. R. Schryer and G. B. Hoflund eds., 1990.

[100] J. A. Macken, S. K. Yagnik, and M. A. Samis, IEEE J. Quantum Electron., 25 (1989) 1695. [101] S.A. Starostin, Y. B. Udalov, P. J. M. Peters, and W.J. Witteman, Appl. Phys. Lett., 77 (2000) 3337. [102] V. M. Cherezov, M. Z. Novgorodov, V. N. Ochkin, V. G. Samorodov, E. F. Shishkanov, V. A. Stepanov, and W. J. Witteman, Appl. Phys., B71 (2000) 503. [103] for example, Nippon Shokubai Co. Ltd. Jpn. Pat. Pub. No. H10-244156, Intern. Pub. No. WO97/34692; Dow Chemical Co., Intern. Pub. No. WO98/00413, WO98/00414, and WO98/00415; Bayer AG, DE 198 04 712 A1. [104] B. S. Uphade, T. Akita, T. Nakamura, and M. Haruta, J. Catal., in press. [105] Y. Usami, K. Kagawa, M. Kawazoe, Y. Matsumura, H. Sakurai, and M. Haruta, Appl. Catal. A: General, 171 (1998) 123. [106] F. Cosandey and T. E. Madey, Surf. Rev. Lett., 8 (2001) 73. [107] V. A. Bondzie, S. C. Parker, and C. T. Campbell, Catal. Lett., 63 (1999) 143. [108] R. Schlgl, CATTECH, 5 (2001) 146. [109] D. A. H. Cunningham, T. Kobayashi, N. Kamijo, and M. Haruta, Catal. Lett., 25 (1994) 257. [110] M. Haruta, M. Yoshizaki, D. A. H. Cunningham, and T. Iwasaki, Ultraclean Technology (Japanese), 8 (1996) 1. [111] S. Tsubota, T. Nakamura, K. Tanaka, and M. Haruta, Catal. Lett., 56 (1998) 131.

115

You might also like