Вы находитесь на странице: 1из 15

Electrochimica Acta 47 (2002) 3595 /3609

www.elsevier.com/locate/electacta

Catalytic oxidation of CO by platinum group metals: from ultrahigh


vacuum to elevated pressures
A.K. Santra, D.W. Goodman *
Department of Chemistry, Texas A&M University, P.O. Box 30012, College Station, TX 77842-3012, USA

Received 10 October 2001; received in revised form 15 February 2002

Abstract

CO oxidation over platinum group metals has been investigated for some eight decades by many researchers and is considered to
be the best understood catalytic reaction. Nevertheless, there has been a renewed interest in CO oxidation recently because of its
technological importance in pollution control and fuel cells. Removal of COx from automobile exhaust is accomplished by catalytic
converters using supported Pt, Pd and Rh catalysts. Catalysts are used in fuel cells to remove traces of COx from the H2 feed gas to
the few ppm level necessary for their efficient operation. Efforts have been made in our laboratory to understand the adsorption of
CO and the kinetics of CO-oxidation on both single crystals and supported metal catalysts over a wide temperature (100 /1000 K)
and pressure (1/10 7 /10 Torr) range. By comparing the results of single crystals, model supported catalysts, and supported
technical catalysts the relationship between particle size and catalytic activity can be better understood. Also discussed is CO
oxidation on model supported Au catalysts, a promising new candidate for low temperature CO oxidation. # 2002 Published by
Elsevier Science Ltd.

Keywords: CO oxidation; Catalysts; Automobile exhaust

1. Introduction conventional hydrogen production technologies such as


steam reforming, partial oxidation and autothermal
Catalytic oxidation of CO over platinum group reforming of hydrocarbons produce large amounts of
metals (Pt, Ir, Rh and Pd) has been the subject of CO as a by-product [52,53]. Therefore, it is extremely
many experimental and theoretical investigations [1 /46] important to have a CO oxidation catalyst with very
since the classic work of Langmuir [47] in 1922 and has high efficiency and one that can preferentially oxidize
been extensively reviewed [10,48/50]. Recently, CO CO for the production of CO-free hydrogen stream.
oxidation has attracted renewed attention due to its Numerous adsorption and kinetic studies on single
technological importance in the area of pollution con- crystals and supported metal catalysts have been re-
trol [51] and fuel cells [52,53]. Currently the removal of ported in the last two decades from our laboratory as a
CO from automobile exhaust is accomplished by the function of O2 and CO partial pressure from ultrahigh
oxidation of CO in catalytic converters using supported vacuum (UHV) to elevated pressures (10 Torr) over a
Pt, Pd and Rh catalysts. It is well established that for wide temperature range of 100 /1000 K [1 /6,10 /12,54 /
optimum operation of low temperature fuel cells it is 59]. Although the CO oxidation reaction is the best
essential to have a continuous supply of CO-free understood among the industrially important catalytic
hydrogen. Although, proton exchange membrane fuel reactions, there are many complex aspects to be resolved
cells can tolerate a few ppm level of CO in the hydrogen with respect to a unified reaction mechanism.
stream, alkaline fuel cells require CO-free hydrogen. The Historically, gold is chemically inert compared with
the other Pt group metals, however, recently it has been
shown that Au, deposited as finely dispersed, small
* Corresponding author. Tel.: /1-979-845-6822; fax: /1-979-845-
particles (B/5 nm diameter) on reducible metal oxides
0214 like TiO2, is an excellent catalyst for CO oxidation at
E-mail address: goodman@mail.chem.tamu.edu (D.W. Goodman). relatively low temperatures [54,55,60 /63]. Furthermore,
0013-4686/02/$ - see front matter # 2002 Published by Elsevier Science Ltd.
PII: S 0 0 1 3 - 4 6 8 6 ( 0 2 ) 0 0 3 3 0 - 4
3596 A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609

this catalytic activity has been shown to be a critical 107 Torr O2 at 1300 K) followed by annealing at 1500
function of the Au cluster size. K. The Ru(0001) crystal was cleaned by oxidation (2 /
In this article, we will compare CO adsorption and 107 Torr O2 at 1450 K for 3 min) followed by
oxidation on single crystals with model supported annealing at 1500 K.
catalysts of Pt, Ir, Rh, Pd and Au. Primary emphasis The IR cell with CaF2 window is connected to the
will be given to the results obtained in our laboratory UHV chamber through a double differentially pumped
over wide temperature (100 /1000 K) and pressure sliding seal. The configuration of the elevated pressure
ranges (1 /107 /10 Torr). The goal of this work is cell is similar to that described by Campbell et al. [65].
an understanding of the effects of temperature, pressure This arrangement allows IR experiments in the pressure
and particle size on CO oxidation that spans the range of 10 8 /103 Torr and also provides convenient
material and pressure ‘gaps’ between ‘real world’ access to the sample without opening the UHV cham-
catalysis and ‘surface science’. ber. The pressure in the IR cell was determined using an
ionization gauge and a capacitance manometer at their
respective working pressure ranges.
2. Experimental TiO2(110) single crystals (Commercial Crystal La-
boratories) were used in these studies mainly due to their
The UHV systems used for this work were equipped suitability for atomically-resolved STM and STS experi-
[64] with Auger electron spectroscopy (AES), a quadru- ments. The n-type semiconductor form, sufficiently
pole mass spectrometer for temperature programmed conductive for STM and electron spectroscopic mea-
desorption (TPD), low energy electron diffraction surements, was prepared by cycles of Ar  sputtering
(LEED), scanning tunneling microscopy/spectroscopy and annealing to 700/1000 K. Deposition of the metal
(STM/STS) and an Ar ion sputter gun, contiguous to a was typically accomplished by resistive evaporation of
high-pressure reaction chamber. The single crystal high-purity metal wire wrapped around a W or Ta
samples were mounted on a retractable manipulator,
filament in vacuum. Such dosers provide an excellent
allowing the sample to be moved between the two
means of obtaining a clean and stable metal flux after
chambers in situ. Pt(100), Ir(110) and Pd(110) crystals of
thorough outgassing. By controlling the filament cur-
0.92 cm in diameter and 0.11 cm in thickness were used
rent, doser to substrate distance and the substrate
for the experiments. The Ir(111) sample was elliptical in
temperature, fine control can be exercised over cluster-
shape and 0.75 /0.55 /0.03 cm3 in size. The samples
size and density.
were heated resistively by two high purity (0.051 cm)
Gas chromatography with flame ionization detection
tungsten leads spot-welded to the back of the crystal;
(FID) was used to analyze the reaction products in
sample temperature was measured by a 0.08 cm
which CO and CO2 were catalytically converted to
chromel-alumel thermocouple spot-welded to the sam-
methane before analysis. Rates of reaction are expressed
ple edge.
AES was used to check the cleanliness of the samples. as turnover frequencies (TOF), defined as the number of
In addition, the Pt(100) sample was cleaned of carbon CO2 molecules produced per active metal site per
by O2 adsorption/desorption and of Si and Ca impu- second. For Pd, Rh or Ir, the entire crystal (front,
rities by high temperature oxidation (1123 K, 1 /10 7 back and edge) were included in determining the total
Torr O2) and Ar sputtering. The Pd crystal was cleaned number of sites; for Pt, only the front face was included,
in the reactor with 8 Torr of CO and 8 Torr of O2 and as the back and edge of the crystal were not subjected to
heating to 600 K for 1 /2 min. One to three cycles of this the sputter cleaning procedure. Research grade CO
treatment produced a clean surface. Large carbon (99.99%) and O2 (99.995%) were supplied by Matheson.
impurities were cleaned from the Ir samples by oxida- The CO was further purified by slowly passing it
tion in 1 /105 Torr O2 at 1000 K for 5 /10 min, through a molecular sieve trap at 77 K. No metal
followed by a 3 min anneal to 1600 K. Small traces of carbonyls (e.g. Ni(CO)4 were detected in any experiment
carbon were removed by reaction in 8 Torr O2 and 4 in post-reaction AES analysis.
Torr CO for 2 min at 600/625 K, followed by a brief The experimental procedure has been described in
flash to 1600 K. The Pt sample was cleaned by detail elsewhere [64]. Briefly, after cleaning, the sample
sputtering at 1100 K in 5/105 Torr of Ar for 30 was moved to the reactor and charged with reactants.
min (1 kV beam energy) mainly to remove Si and Ca Most experiments were performed with 16 Torr of CO
impurities. This treatment was followed by heating in and 8 Torr O2. The sample is heated to the desired
0.1 Torr of O2 at 1100 K for 30 min to remove traces of temperature for a specified time, the products are then
carbon and then annealing at 1300 K. Repeated cycles allowed to mix for 15 min and then an aliquot of the
of this procedure produced a clean Pt surface, which product mixture was analyzed by GC. The reactor was
could not be oxidized under UHV conditions at high then evacuated, and the sample returned to the UHV
temperature. Rh crystals were cleaned by oxidation (2 / chamber for post-reaction analysis.
A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609 3597

0.86 ML. The species has been identified as the a-top


type and three different ordered structures at uCO /
0.33, 0.5 and 0.80 ML have been identified by LEED.
A regular increase in the C /O frequency in the IRAS
data along with an increase in the CO desorption
temperature with respect to the increase in coverage
(uCO) have been observed and understood to be due to
lateral CO /CO repulsion. On the reconstructed (5 /1)
as well as on the (1 /1) Ir(100) surface, formation of a
c(2 /2) overlayer structure has been reported [76]. In
this article, we would like to concentrate on the
adsorptive behavior of CO over a wide pressure range
from UHV to 10 Torr and a temperature range of 100/
1000 K using Pd as an example. We will also discuss the
effect of particle size on the catalytic activity.
CO adsorption on Pd(100) and Pd(111), studied under
UHV conditions using an array of surface science
techniques, exhibits structure sensitivity depending on
the face of the Pd crystal. For example, on Pd(111), at a
CO coverage (uCO)B/1/3 ML, CO adsorbs at 3-fold
hollow sites, with a structure corresponding to (3/
3)R 308 as revealed by LEED and a carbon /oxygen
stretching frequency of 1850 cm 1 [77,78]. However at a
coverage (uCO) of 0.5 ML, a c (2 /2) LEED pattern with
a C /O stretching frequency of 1918 cm 1 has been
observed and is assigned to a CO adsorbed onto bridge-
bound sites. At very high coverages, a-top and bridge-
Fig. 1. IR spectra of CO on Pd(111) at Pco /10.0 Torr as a function of
temperature. The spectra were collected from high temperature to low bound CO co-exist yielding a (2 /2) LEED pattern with
temperature [85a,85b]. C /O frequencies near 1951 and 2097 cm 1, respec-
tively. In contrast, only bridge-bonded CO has been
3. CO adsorption

Carbon monoxide adsorption under UHV conditions


especially on transition metal surfaces has been studied
extensively [10,48/50,66 /68]. Three different types of
adsorption sites have been observed, in general, namely
a-top, bridge-bound and 3-fold hollow. The interaction
of a CO molecule with a transition metal can be
understood by a simple donor /acceptor model first
described by Blyholder [69].
Both linear and bridge bonded CO has been observed
on Rh(100) for the entire coverage and temperature
range of 90 /300 K [70]. A transition from a well ordered
c (2 /2) structure at uCO /0.5 ML to a p (42 /
42)R 458 coincidence structure at uCO /0.7 ML has
been observed by LEED and infrared reflection absorp-
tion spectroscopy (IRAS). On Pt single crystal surfaces,
however, all three types of CO have been observed
depending on the CO coverage (uCO), adsorption
temperature, pressure and crystal face [20,35,45,67,71/
74]. Three different desorption features of CO have been
observed on the Ir(110) surface [75], whereas the
corresponding IR spectra showed a continuous increase Fig. 2. IR spectra of CO on Pd(111) at Pco /1.0 /10 6 Torr as a
in the C /O stretching frequency from 2001 cm 1 at the function of temperature. The spectra were collected from high
lowest coverage to 2086 cm 1 at a coverage (uCO) near temperature to low temperature [85a,85b].
3598 A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609

sentative case, 1/106 Torr data are shown in Fig. 2


[85a,85b]. Transition from 3-fold hollow to bridging
sites can clearly be seen as the C /O stretching frequency
changes from 1855 to 1900 cm 1. As the population of
CO increases the frequency shifts toward higher values
and the peak becomes sharper. This blue-shift has been
attributed to repulsive lateral interactions between CO
molecules. The frequencies of the a-top and 3-fold
hollow sites are 2110 and 1895 cm 1. The highest
frequency observed for the bridge-bound CO was 1962
cm 1.
An equilibrium phase diagram for CO on Pd(111) is
shown in Fig. 3, based on the adsorption data collected
at the temperature range 90/1000 K and a CO pressure
range of 107 /10 Torr [85a,85b]. It is important to
mention that this phase diagram is valid only at
equilibrium conditions. This is the case for the
bridging 0/a-top/3-fold hollow phase transition. Non-
equilibrium CO adsorption can also occur at low
Fig. 3. CO /Pd(111) equilibrium pressure /temperature phase diagram
showing the different CO adsorption phases: a-top/3-fold hollow, adsorption temperatures (crosshatched region of Fig.
bridging, 3-fold hollow and non-adsorbed [85a,85b]. 3). In this region appropriate adsorption conditions
have to be used to ensure a fully equilibrated adsorbate
observed on Pd(100) even up to uCO /0.5 ML with a layer. This equilibrium phase diagram illustrates that
c (22 /2)R 458 LEED pattern [79 /84]. In this study, low pressure/low temperature adsorption information
for uCO /0.5 ML, a uniaxial compression takes place can be extrapolated into the high pressure/high tem-
resulting in the formation of incommensurate over- perature regime provided that certain initial adsorption
layers. As a consequence of this compression of the conditions are used.
CO overlayer, a sharp decrease in the heat of adsorption In Fig. 4 two series of IR spectra of CO adsorption at
has been observed at uCO /0.5 ML due mainly to lateral Pd(100) surface are shown at CO pressures of 1/106
repulsive interactions. and 1 Torr [85a,85b], respectively. In agreement with
previous studies [83,84] only bridge-bound CO was
detected near the C /O stretching frequency of 1895
3.1. Effects of temperature and pressure cm 1 at the lowest coverage. The vibration frequency
that corresponds to a coverage near 0.5 ML was blue-
A series of IR spectra of adsorbed CO on Pd(111) shifted to 1950 cm 1, which is further blue-shifted to
obtained at a CO pressure of 10 Torr is shown in Fig. 1 1995 /1998 cm1 as the coverage is increased to 0.8 ML.
[85a,85b]. At high temperatures (i.e. low uCO) only 3-
fold hollow sites are occupied. With decreasing tem- 3.2. Particle size effects
perature (i.e. increasing uCO) the population at the
bridging sites increases. At temperatures near 200 K, a As we have seen in the previous section CO adsorp-
sudden change in the spectrum is apparent, correspond- tion behavior changes dramatically with respect to the
ing to a transition from bridging sites to a mixture of a- structure of the crystal surface indicating particle size
top and 3-fold hollow sites. The bridging0/a-top/3-fold effects. In Fig. 5, IRAS data for CO adsorbed on Pd/
hollow transition takes place within a very narrow Al2O3/Ta(110) model catalysts for uPd /5.0 and 1.0 ML
temperature range while the transition from the 3-fold as a function of temperature is presented [86]. These
hollow to bridging sites is much less defined. The narrow results were independent of the direction of temperature
spectral line-width, due to the 3-fold hollow sites, variation provided the samples were annealed in CO to
indicates the presence of a very well ordered CO 600 K. On the uPd /5.0 ML catalyst (Fig. 5a), a peak at
adsorbate layer, and has been confirmed by a very 1894 cm1 was observed at 500 K. With a decrease in
sharp (2 /2) LEED pattern. temperature, this peak shifted towards higher frequency,
In addition to the 10.0 Torr data shown in Fig. 1, became broader, and at 300 K split into two peaks at
isobaric adsorption data were collected at every decade 1984 and 1942 cm.1 In addition, a peak at 2076 cm 1
from 107 to 10 Torr. In each of these isobaric series, appeared at 400 K. These three features continued to
adsorption site progression was identical; the only grow, gradually shifting to higher frequency until, at 150
apparent difference was the temperature at which the K, a new peak appeared at 1890 cm1. These peaks
adsorption site transformation took place. As a repre- have striking similarities to those observed on single
A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609 3599

Fig. 4. IR spectra of CO on Pd(100) as a function of sample temperature at CO pressures of (a) 1.0 Torr and (b) 1/10 6 Torr [85a,85b].

crystal Pd(100) and Pd(111) surfaces as discussed in the cm 1 at 500 K can be attributed to a combination of 3-
previous section. The broad peak corresponding to 1894 fold hollow and bridge-bound CO species, shifting

Fig. 5. Temperature dependent IR spectra of CO adsorbed on uPd /5.0 and 1.0 ML Pd/Al2O3/Ta(110) catalysts at 1/10 5 Torr [86].
3600 A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609

Fig. 6. Arrhenius plot of the CO/O2 specific rates of reaction (TOF) for (a) single crystal (Pd, Ir and Pt) and their supported catalysts and (b) for Rh
single crystals [10,11].

towards exclusively bridge-bound species with decreas- 4. CO oxidation


ing temperature. At 300 K, the features at 1984 and 1942
cm 1 are assigned to contributions from bridging CO 4.1. Pt, Ir, Rh and Pd single crystals and model supported
from Pd(100) to Pd(111) facets, respectively. The feature catalysts
at 2076 cm 1 and 400 K is associated with an a-top
species on 111 facets. Near saturation coverage, the 4.1.1. Steady-state reaction kinetics
peak at 1998 cm 1 corresponds to a bridging species on The CO2 formation rate as a function of inverse
100 facets while the features at 2108, 1960 and 1890 temperature (1/T ) for Pd, Pt and Ir single crystals is
cm 1 correspond to a-top, bridging and 3-fold hollow shown [10] in Fig. 6a and compared with the data
sites, respectively, on 111 facets. obtained on several supported metal catalysts [38]. Data
The results obtained for a uPd /1.0 ML catalyst (Fig. obtained on Rh(100) and Rh(111) single crystals are
5b) differ significantly from the uPd /5.0 ML sample. shown [11] in Fig. 6b. The Pd, Pt and Ir single crystal
The broad peak near the region of bridge-bound species data are for a (1:2) O2: CO mixture at a total pressure of
never splits cleanly into individual components as is the 24 Torr, whereas, the Rh single crystal data are for a
case for larger particles. Moreover, no 3-fold hollow (1:1) O2: CO mixture at a total pressure of 16 Torr.
feature is evident at saturation coverage and the ratio of Within this pressure range the reaction rate is zero-order
a-top/bridge intensity is higher compare to the larger with respect to total pressure. The TOF for the single-
particles indicating that the smaller particles contain a crystal catalysts traverse four orders of magnitude over
higher proportion of edge/defects sites. The broader a temperature range of 450/600 K. Kinetic measure-
features observed at saturation coverage for the smaller ments over such a wide temperature range with sup-
particles are due to higher surface curvature on the small ported catalysts are not possible due to heat and mass
transfer limitations encountered at high temperatures.
particles and a correspondingly less compressed CO
Thus a direct comparison between the two types of
overlayer assuming roughly hemispherical shape of the
catalysts is limited to a relatively small temperature
particles. That the types of transitions (bridging 0/a-top/
range. Nevertheless, it is very clear from Fig. 6a and b
3-fold hollow) occur on the Pd(111) single crystal and,
that there is excellent agreement between the single
to a more limited extent, on the uPd /5.0 ML catalyst crystal and model supported systems with respect to the
do not occur on the uPd /1.0 ML particles is consistent specific reaction rates and apparent activation energies
with the reduced CO density on the smaller particles. [28,38].
The effect of particle size with respect to CO oxidation Fig. 7a /c [10,11] show reaction rate dependence on
will be discussed subsequently. CO partial pressure for Pd(110), Ir(111) and Rh(111) /
A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609 3601

Rh(100), respectively. In these experiments the CO


pressure was allowed to vary keeping the oxygen
pressure fixed. In the case of Pd and Ir, up to a CO:
O2 ratio of approximately 1:12 the reaction was first-
order with respect to CO partial pressure; whereas below
this ratio the reaction became negative-first-order with
respect to CO. For Rh(111) and Rh(100) such a role-
over with respect to the partial pressure of CO has been
observed with a maximum activity occurring approxi-
mately at a O2: CO ratio of 30:1. However, at a lower
partial pressure of O2 (8 Torr), for Pd(110), Rh(100) and
Rh(111), the reaction order has been observed to be
negative-first-order with respect to CO partial pressure
(Fig. 3a and c, respectively). It is noteworthy (Fig. 7c)
that the rate of CO2 formation on Rh(100) is higher
compared with that on Rh(111) at any given CO partial
pressure indicating that the reaction may be structure
sensitive.
For Pd, Ir and Rh the reaction order with respect to
the partial pressure of O2 (Fig. 8a/c) is generally
positive-first-order [10,11]. Above an O2:CO ratio of
12:1, the same ratio at which the CO order changes from
negative to positive, the reaction rate begins to decrease,
becoming negative-order in O2 partial pressures. For
example, at extremely high O2:CO ratios, the order of
reaction for O2 is /0.79/0.2 on Ir(111). For Pd,
changing the CO partial pressure at a constant tem-
perature only shifts the curve, i.e. the maximum rate and
the ratio at which the rate turns over, remain un-
changed. Whereas, for Ir the O2:CO ratio at which the
rate varies from first-order oxygen dependency changes
somewhat (from 12:1 to 16:1) as does the order of the
reaction. The reaction rate on both Rh surfaces (Fig. 8c)
increases linearly at low oxygen partial pressures. This
first-order dependence is altered at high oxygen pres-
sures where the rates roll over and become negative-
order with respect to the partial pressure of oxygen.
Note that this rollover occurs at different oxygen partial
pressures on the two single crystal surfaces indicating
the possibility of structure sensitivity.
On Pt(100), the order of the reaction with respect to
CO partial pressure changes with temperature (Fig. 9)
[10]. In the range where the activation energy is
changing (425 /490 K), the reaction order varies from
0.0 to 0.6. Above 500 K the order becomes more
negative and rapidly approaches negative-first-order.
The reaction never becomes positive-first-order with
respect to CO partial pressure even at an O2: CO ratio of
200:1, as observed for Pd, Ir and Rh.
The Pt(100) surface shows (Fig. 10) [10] only positive-
first-order behavior with respect to the partial pressure
of oxygen. The range of O2:CO ratios studied at a given
temperature was limited to TOF’s where there was
differential conversion (B/5%) of CO. No decrease in
Fig. 7. CO partial pressure dependence at constant oxygen pressure
the order of reaction is observed for O2:CO ratios of 1:5
and temperature: (a) on Pd(110), (b) on Ir(111) and (c) on Rh(111) and to 150:1, and temperatures from 475 to 650 K, indicat-
Rh(100) [10,11].
3602 A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609

Fig. 8. O2 partial pressure dependence at constant CO pressure and temperature: (a) on Pd(110), (b) on Ir(111) and (c) on Rh(111) and Rh(100)
[10,11].

ing that under these experimental conditions no strongly on Rh single crystal surfaces wherein the CO2 deso-
bound, deactivating oxygen species is present. In order rption temperature differs by /50K on Rh(100) and
to form a similar species on Pt, much higher O2 pressure Rh(111) surfaces. The Rh(100) surface shows signifi-
and/or temperatures would be necessary, conditions not cantly higher reaction rates compared with that on the
accessible in our experiments. Rh(111) surface. The other difference between the two
systems is that on Rh(100) all CO is oxidized to CO2,
4.1.2. Structure sensitivity and particle size effects whereas on Rh(111) a fraction of the CO desorbs. The
Although it appears from the data in Fig. 1 that the reason for such high activity and selectivity on Rh(100)
CO-oxidation reaction is structure insensitive, it should towards CO2 formation is assumed to be due to the
be noted that the single crystal rates are compared with surface reaction step, COads/Oads 0/CO2(g), being in-
the least dispersed supported catalyst. Very recently, trinsically faster on Rh(100) than on Rh(111). In
under UHV conditions structure sensitivity has also contrast to the high-pressure data (Fig. 7c and Fig.
been observed by Niemantsverdriet and co-workers [90] 8c), the reaction exhibits first order kinetics with respect
A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609 3603

Fig. 9. CO partial pressure dependence as a function of temperature at constant oxygen pressure on Pt(100). (a) The Arrhenius plot for CO oxidation
on Pt(100) is shown to illustrate the temperature regimes in which (b) the CO pressure dependence data were obtained [10].

Fig. 10. O2 partial pressure dependence at constant CO pressure and


temperature on Pt(100). In all cases the reaction order is 1.09/0.1 [10].

to both the partial pressure of CO and O2 under UHV Fig. 11. Effect of particle size on the CO2 formation rate on Ir/SiO2
conditions, however, the structure sensitivity is consis- and Ir single crystal catalysts [38].
tent with the high-pressure data (Fig. 7c and Fig. 8).
Dramatic structure sensitivity has been observed on the dispersion values obtained by Cant and co-workers
Ir/SiO2 catalyst [38] (Fig. 11). It is clear from Fig. 11 [38]. It is noteworthy that the single crystal data in Fig.
that larger particles are more active for CO oxidation 11 corresponds to the supported data extrapolated to a
and the data more comparable to the single crystal data. particle size of 40 nm. This is close to the ‘effective’
The data for the larger particles were calculated from particle size of the single crystals used in this study,
3604 A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609

shown that on Pt(335) surface dissociation of CO and


O2 occurs preferentially at step sites. It also has been
observed that CO adsorbed at 111 terraces is more
active compared with CO at 100 step sites whereas
chemisorbed oxygen atoms at step sites is more active
compared with CO adsorbed at terraces. Using isotopi-
cally labeled 13C18O molecules, it has been shown that
the oxidation rate at step sites is twice the oxidation rate
at terrace sites. However, recent STM studies of CO
oxidation on Pt(111) have shown that the reaction
occurs exclusively at the boundaries between (2 /
2)Oads and c(4 /2)COads domains when co-adsorbed
Fig. 12. CO oxidation with O2 over model Pd/SiO2/Mo(100) and a
conventional 5% Pd/SiO2 catalyst. Reaction conditions were PTorr / under UHV condition [17,87 /89]. Therefore, to increase
0.5 Torr and CO /O2 /0.2 [56,86]. the reactivity of the surface it is necessary to increase the
interface between the Oads and COads domains, i.e. to
increase the coverage of Oads.

4.1.3. CO oxidation versus metal /oxygen bond energy


The oxidation rate of CO under steady-state condi-
tions on various Pt group metals are compared in Fig.
13 [94]. The temperature of the reaction was chosen
specifically as 793 K in order to avoid surface site
blocking due to CO adsorption; the pressure of CO and
O2 was 1 /107 Torr. From the data of Fig. 13 it is
clear that the most active metals (Pd, Pt, Ir and Rh) for
CO oxidation have M /O bond energies within the range
Fig. 13. Dependence of the rate of CO oxidation on metals on the 320 /390 kJ mol 1. In the case of metals having M /O
oxygen bond energy EMO (793 K, PO2 /PCO /10 7 Torr) [94]. bond energies less than 320 kJ mol 1, e.g. Ag or Au, the
rate-determining step for CO oxidation is the adsorption
taking into account step densities and edge effects. The and dissociation of oxygen. On the other hand, for
decreasing activity of the reaction with Ir particle size metals having M/O bond energies larger than 390 kJ
can be understood as due to a preferential poisoning of mol 1, the rate-determining step is the reaction between
the active sites by carbon formed by CO dissociation, a COads and Oads. In other words those metals on the right
competing process to CO oxidation under such high- hand side of Fig. 13 have a higher tendency to form
pressure conditions. This reaction would be expected to stable oxides. In fact, on Rh(111) and (100) single
occur more rapidly on defect sites and step edges, crystals, high O2:CO ratios result in a decrease in the
present at much higher concentration on the smaller overall rate and a change from positive-order in oxygen
particles. to negative-order. This change was directly correlated
CO oxidation has also been investigated over Pd/ with the formation of an oxide-like species, as deter-
SiO2/Mo(100) model catalyst [56,86]. The reaction mined by post-reaction AES and TPD [10,11]. Pd(110),
conditions for the catalysts were 10 Torr CO, 5.0 Torr Ir(111) and Ir(110) exhibit partial pressure dependence
O2, and reaction temperatures in the range 540 /625 K. and high oxygen pressure behavior similar to Rh. For
The conversions were maintained at less than 50% and Rh, the formation of a near surface oxide (probably
were measured by monitoring the pressure decrease in a Rh2O3 [10,11] which is much less active) is responsible
static reactor of known volume (750 cm3). The average for the deactivation. The similar behavior of Pd and Ir
cluster sizes shown in Fig. 12 were determined by CO suggests a similar deactivation mechanism. In contrast,
TPD, O2 TPD and ex situ STM/AFM and all were in on Ru the oxide was found to be substantially more
good agreement. The specific reaction rates were some- active than the clean surface, and the reaction order in
what higher for the model catalysts than the high- oxygen pressure increases by approximately 3-fold
surface-area catalysts, but the activation energies are [9,13,80,95].
remarkably similar. There was no noticeable depen-
dence of the CO2 formation rate on the Pd cluster size, 4.1.4. Reaction mechanism
indicating that CO oxidation over Pd/SiO2 is structure The Langmuir /Hinshelwood reaction between COads
insensitive. and Oads is well established as the dominant reaction
Structure sensitivity has also been observed for Pt. mechanism for conditions where CO is the primary
Yates and co-worker [91 /93] using IRAS and TPD have surface species [48,49]. This mechanism has been con-
A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609 3605

4.1.5. Angle-resolved temperature programmed


desorption
In search of the possible reaction mechanism of CO
oxidation on both polycrystalline and Pt(111) surfaces,
angle resolved reactive temperature programmed deso-
rption (RTPD) has been performed using time of flight
(TOF) mass spectrometry [32,97 /99]. The results show
dramatic angular and velocity dependence of CO2
desorption from the reaction of CO and oxygen co-
adsorbed on a Pt(111) surface at 100 K. The velocity
integrated desorption spectra [99] (Fig. 14) show four
different peaks (a, b3, b2 and b1) for CO2 formation at
145, 210, 250 and 330 K, respectively. These feature
have been attributed to four different reaction mechan-
isms operating at various temperatures depending upon
the relative binding of oxygen and the geometric
arrangement and coverage of the adsorbed species.
Although the precise origin of b3 and b2 processes are
Fig. 14. Reactive temperature programmed desorption (RTPD) of
not clear at present, the a-CO2 formation temperature
CO2 as a function of emission angle after predosing 1.5 /1015 cm 2
molecules of O2 and 2.7/1015 cm2 molecules of CO successively at coincides with that of the molecular O2 desorption. The
Ts /100 K [99]. b1-CO2 formation is most likely the mechanism pro-
posed in the previous section due to the reaction at the
firmed by numerous UHV studies of the co-adsorption overlapping regions of COads and Oads island bound-
of reactants, transient kinetic studies, and steady-state aries. Very interesting oscillatory behavior of the CO
kinetics [1 /46]. The reaction steps can be written as oxidation reaction has been observed on Pt single crystal
under UHV conditions [15,16,100 /102], however, this
CO(g) 0 COads (1) behavior is beyond the scope of the present article.
O2 (g) 0 2Oads (2)
COads Oads 0 CO2 (g) (3) 4.2. CO oxidation on Au
where the recombinative desorption of adsorbed Oads
atoms (reverse of reaction 2) and the dissociative In the bulk form, Au is known to be chemically inert
chemisorption of CO2 (reverse of reaction 3) are compared with the other Pt group metals. However
neglected. It is assumed that CO is the dominant surface recently it has been shown that Au clusters, deposited as
species [96]. Considering the reaction steps (1/3) and finely dispersed, small particles (B/5 nm diameter) on
using the above assumptions, an approximate rate reducible metal oxides like TiO2, Fe2O3 and Co3O4,
expression, originally proposed by Langmuir [47] for dramatically enhance the rate of a number of indust-
Pt can be derived as: rially important reactions. Reactions catalyzed by Au
particles on TiO2 supports are CO oxidation, hydro-
d[CO2 ]=dtk exp (Edes;CO =F )PO2 =PCO (4) genation and partial oxidation of hydrocarbons and the
selective oxidation of higher alkenes [54,55,60 /63]. It
where the reaction rate is independent of total pressure,
has also been observed that catalytic activity of these
first-order in O2 pressure and negative-first-order with
catalysts is a function of cluster size.
respect to the CO pressure. The rate of the reaction is
then governed by the desorption of CO or the lifetime of
CO (tCO) on the surface [48], depending on the reaction 4.2.1. Characterization of the Au clusters by STM and
temperature, whereas the pressure dependence simply STS
reflects the competition for adsorption sites between The constant current STM micrographs in Fig. 15
oxygen and CO. It has been found [10] that the kinetics show [103] changes in the clusters with respect to the
on Pd, Ir and on Pt at high temperatures, are consistent amount of Au deposited. At a relatively low coverage of
with this model in that the pressure dependencies Au (uAu /0.1 ML) hemispherical 3D clusters are
predicted by equation 4 are observed. In addition, a observed with diameters of 2 /3 nm and heights of 1/
correlation of the activation energies between supported 1.5 nm. Interestingly the clusters mainly grow along the
and single crystal data, and among different single- step edges. Well-dispersed quasi-2D clusters, having a
crystal planes [28,38] reflect the fact that the binding height of 0.3 /0.6 nm and a diameter of 0.5 /1.5 nm, can
energy of CO does not vary greatly among these metal be seen on the terraces. With increasing Au coverage
catalyst surfaces. (uAu), the clusters steadily grow larger while the increase
3606 A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609

Fig. 15. A set of 50 /50 nm2 STM images (2.0 V, 1.0 nA) of TiO2(110) /(1/1) with different Au coverage (uAu): (A) 0.10 ML, (B) 0.25 ML, (C) 0.50
ML, (D) 1.0 ML, (E) 2.0 ML and (F) 4.0 MLE. With increasing coverage, Au clusters grow and gradually cover the surface [103].

in cluster density is minimal. However, even at uAu /4.0 1). Three-dimensional Au clusters, imaged as bright
ML, some portions of the TiO2 substrate are still visible. protrusions, have average diameters of /2.6 nm and
In Fig. 16A, the constant current STM micrograph of heights of /0.7 nm (corresponding to 2/3 atoms thick)
Au (uAu /0.25 ML) deposited onto single crystal and are preferentially nucleated at the step edges. Quasi-
TiO2(110) /(1 /1) [55,103] is shown. The metal deposi- two-dimensional clusters are characterized by heights of
tion was performed at 300 K, followed by annealing to 1 /2 atomic layers. Previous studies have shown that the
850 K for 2 min to stabilize the clusters. In Fig. 16A only Au clusters upon annealing form large microcrystals
the Ti cations are visible; whereas the O2 anion are not with well-defined hexagonal shapes.
seen. The inter-atomic distance between the 001 rows Fig. 16B shows STS taken over various clusters on the
is /0.65 nm, which can be observed along the terraces surface, where the tunneling current (I) as a function of
corresponding to the length of the unit cell along the bias voltage (V) across the STM tip is measured. The I/
[110] direction of the unreconstructed TiO2(110) /(1 / V curves correlate with the Au cluster size on the TiO2
A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609 3607

Fig. 17. CO oxidation TOFs as a function of the Au cluster size


supported on TiO2. (A) The Au/TiO2 catalysts were prepared by a
precipitation method, and the average cluster size was measured by
TEM, 300 K. (B) The Au/TiO2 catalysts were prepared by vapor-
deposition of Au on planner TiO2 films on Mo(100). The CO /O2
mixture was 1:5 at a total pressure of 40 Torr, 350 K [55].

Fig. 16. (a) A CCT /STM image of a Au (uAu /0.25 ML) deposited
onto TiO2(110) /(1/1) prepared just prior to a CO:O2 reaction. The
sample had been annealed to 850 K for 2 min; (b) STS data acquired
for Au clusters of varying sizes on the TiO2(110) /(1/1). An STS of
TiO2 substrate, having a wider band-gap than the Au cluster, is also
shown as a point of reference [55]. Fig. 18. The specific activity for CO conversion as a function of
reaction time at 300 K on a model Au/TiO2/Mo(100) catalyst. The Au
coverage (uAu) was 0.25 ML, corresponding to an average cluster size
surface. The length of the observed plateau at the zero
of /2.4 nm [103].
tunneling current is a measure of band-gap (along the
bias voltage axis) of electrons tunneling between the
GaAs(110) [104]. Very small clusters then are non-
valence and conduction band of the cluster and tip. The
metallic and exhibit electronic and chemical properties
electronic character of these clusters vary between that
unlike those of the corresponding bulk metal.
of a metal and a non-metal depending on their size.
With an increase in size the clusters gradually exhibit
4.2.2. Particle size effects
metallic character with an enhanced density of states at Interestingly, a marked correlation between the clus-
the Fermi level. Note that clusters of 2.5 /0.7 nm2 size ter size and catalytic activity has been observed for CO
have a larger band gap than that for a cluster 5.0 /2.5 oxidation over Au/TiO2 system [54,55,60 /63]. Studies
nm2 in size. Smaller clusters have a non-metallic have been carried out on Au/TiO2/Mo(100) as well as on
character resulting in significant band-gap and a re- Au/TiO2(110) /(1 /1) for comparison. Fig. 17a and b
duced density of states near the Fermi level. A similar show plots of CO oxidation activity (TOF) at 350 K as a
metal to non-metal transition with respect to cluster size function of Au cluster size supported on TiO2(110) /
has also been observed for Fe clusters deposited on (1 /1) and TiO2 /Mo(100) substrates, respectively.
3608 A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609

These results show similarities in the structure sensitivity optimum reaction temperature, deactivation, etc. Fi-
of CO oxidation with a maximum activity evident at /3 nally it has been shown that ultra-small gold particles,
nm Au cluster size on both TiO2 supports. For each unlike bulk gold metal, are promising catalysts for low
catalyst, the activity and the selectivity of the supported temperature CO oxidation.
Au clusters are markedly size-dependent. Although the
TiO2 supported Au catalysts exhibit a high activity for
the low-temperature CO oxidation, rapid deactivation Acknowledgements
was observed as a function of reaction time. Fig. 18
shows a plot of TOF versus time for CO oxidation at The support of this work by the Department of
300 K on a Au(uAu /0.25 ML)/TiO2/Mo(100) model Energy, Office of Basic Energy Sciences, Division of
catalyst. The model catalyst, which exhibited a high Chemical Sciences, and the Robert A. Welch Founda-
initial activity, deactivated after a CO/O2 (1:5) reaction tion is gratefully acknowledged.
of /120 min at 40 Torr. This deactivation is due to
agglomeration of the Au clusters with reaction time, as
detailed by post-reaction STM measurements. The STM References
data clearly demonstrate [54,55,60 /63] that under reac-
tion conditions the Au clusters ripen via an Ostwald [1] J. Szanyi, D.W. Goodman, J. Phys. Chem. 98 (1994) 2972.
mechanism; i.e. large clusters grow at the expanse of [2] J. Szanyi, W.K. Kuhn, D.W. Goodman, J. Phys. Chem. 98
(1994) 2978.
small ones leading to a bimodal size distribution. This [3] J. Szanyi, D.W. Goodman, J. Catal. 145 (1994) 508.
ripening mechanism depends upon the strength of the [4] J. Szanyi, D.W. Goodman, Stud. Surf. Sci. Catal. 75 (1993) 1599.
cluster/support interaction as well as gas pressure. The [5] J. Szanyi, D.W. Goodman, Catal. Lett. 21 (1993) 165.
TOF of CO oxidation, which maximizes with respect to [6] J. Szanyi, X.P. Xu, D.W. Goodman, Abs. Am. Chem. Soc. 206
cluster size, correlates with a metal to non-metal (1993) 83.
[7] J. Szanyi, D.W. Goodman, Catal. Lett. 14 (1992) 27.
transition at a particle size of /3 nm, as revealed by [8] J.A. Rodriguez, D.W. Goodman, Surf. Sci. Rep. 14 (1991) 1.
a detailed STM /STS investigation. This behavior has [9] C.H.F. Peden, D.W. Goodman, M.D. Weisel, F.M. Hoffmann,
been discussed in the previous section and there is no Surf. Sci. 253 (1991) 44.
unified theory to understand the catalytic activity of the [10] P.J. Berlowitz, C.H.F. Peden, D.W. Goodman, J. Phys. Chem.
92 (1988) 5213.
small gold particles at the moment. However, the fact
[11] C.H.F. Peden, D.W. Goodman, D.S. Blair, P.J. Berlowitz, G.B.
that CO oxidation activity increases with decreasing Fisher, S.H. Oh, J. Phys. Chem. 92 (1988) 1563.
cluster size as long as the particles are metallic suggests [12] D.W. Goodman, C.H.F. Peden, J. Phys. Chem. 90 (1986) 4839.
that an overall catalytic activity that depends on both [13] C.H.F. Peden, D.W. Goodman, J. Vac. Sci. Technol. A 3 (1985)
electronic as well as geometric factors. Electronic factor 1558.
means the size-induced changes in the electronic levels of [14] A. Von Oertzen, H.H. Rotermund, A.S. Mikhailov, G. Ertl, J.
Phys. Chem. B 104 (2000) 3155.
the small clusters namely electronegativity, charge state [15] P. Strasser, M. Lubke, F. Raspel, M. Eiswirth, G. Ertl, J. Chem.
and metallicity leading to changes in the cluster /CO Phys. 107 (1997) 979.
and /O2 interaction strength thereby the overall reac- [16] P. Strasser, M. Eiswirth, G. Ertl, J. Chem. Phys. 107 (1997) 991.
tion itself. On the other hand, geometric factor leads to [17] S. Volkening, J. Wintterlin, J. Chem. Phys. 114 (2001) 6382.
changes in the shape with respect to the changes in the [18] J. Dicke, H.H. Rotermund, J. Lauterbach, Surf. Sci. 454 (2000)
352.
particles size leading to change in the number of steps, [19] M.G. Moula, S. Wako, Y. Ohno, M.U. Kislyuk, I. Kobal, T.
size of the terraces and facetes. Matsushima, Phys. Chem. Chem. Phys. 2 (2000) 2773.
[20] K. Bleakley, P. Hu, J. Am. Chem. Soc. 121 (1999) 7644.
[21] C. Stampfl, M. Scheffler, Surf. Sci. 435 (1999) 119.
[22] G.L. Dong, J.G. Wang, Y.B. Gao, S.Y. Chen, Catal. Lett. 58
5. Conclusions
(1999) 37.
[23] M. Menon, B.C. Khanra, Ind. J. Chem. A 37 (1998) 1070.
An approach that combines both surface science and [24] D.C. Skelton, R.G. Tobin, D.K. Lambert, C.L. Dimaggio, G.B.
traditional catalytic methodologies has been shown to Fisher, J. Phys. Chem. B 103 (1999) 964.
be extremely advantageous in bridging the material and [25] G.A. Somorjai, Appl. Surf. Sci. 121 (1997) 1.
pressure ‘gaps’ between ‘real world catalysis’ and ‘sur- [26] X.C. Su, P.S. Cremer, Y.R. Shen, G.A. Somorjai, J. Am. Chem.
Soc. 119 (1997) 3994.
face science’. Although, it appears that the CO oxida- [27] S. Akhter, J.M. White, Surf. Sci. 171 (1986) 527.
tion reaction follows a simple Langmuir-Hinshelwood [28] N.W. Cant, D.E. Angove, J. Catal. 97 (1986) 36.
mechanism, many complicating factors influence the [29] E.M. Stuve, R.J. Madix, C.R. Brundle, Surf. Sci. 146 (1984) 155.
overall activity. The example of Pd shows that the [30] J.T. Kiss, R.D. Gonzalez, J. Phys. Chem. 88 (1984) 892.
[31] J.T. Kiss, R.D. Gonzalez, J. Phys. Chem. 88 (1984) 898.
knowledge acquired from single crystal data can be used
[32] T. Matsushima, J. Phys. Chem. 88 (1984) 202.
to understand the results for model supported catalysts, [33] M. Kawai, T. Onishi, K. Tamaru, Appl. Surf. Sci. 8 (1981) 361.
including structure sensitivity, reaction order determina- [34] W.H. Weinberg, D.E. Ibbotson, J.L. Taylor, J. Vac. Sci.
tion, nature of the bonding of adsorbed molecules, the Technol. 18 (1981) 620.
A.K. Santra, D.W. Goodman / Electrochimica Acta 47 (2002) 3595 /3609 3609

[35] C.T. Campbell, G. Ertl, H. Kuipers, J. Segner, J. Chem. Phys. 73 [72] G. Rupprechter, T. Dellwig, H. Unterhalt, H.J. Freund, J. Phys.
(1980) 5862. Chem. B 105 (2001) 3797.
[36] J.L. Taylor, D.E. Ibbotson, W.H. Weinberg, Surf. Sci. 90 (1979) [73] G. Rupprechter, T. Dellwig, H. Unterhalt, H.J. Freund, Top.
37. Catal. 15 (2001) 19.
[37] J.L. Taylor, D.E. Ibbotson, W.H. Weinberg, J. Chem. Phys. 69 [74] Y.Y. Yeo, L. Vattuone, D.A. King, J. Chem. Phys. 106 (1997)
(1978) 4298. 392.
[38] N.W. Cant, P.C. Hicks, B.S. Lennon, J. Catal. 54 (1978) 372. [75] K.J. Lyons, J. Xie, W.J. Mitchell, W.H. Weinberg, Surf. Sci. 325
[39] D.I. Hagen, B.E. Nieuwenhuys, G. Rovida, G.A. Somorjai, (1995) 85.
Surf. Sci. 57 (1976) 632. [76] J.T. Grant, Surf. Sci. 18 (1969) 228.
[40] T. Matsushima, C.J. Mussett, J.M. White, J. Catal. 41 (1976) [77] F.M. Hoffmann, Surf. Sci. Rep. 3 (1983) 107.
397. [78] R. Burch (Ed.), Catal. Today 9 (1991) R9.
[41] T. Matsushima, J.M. White, J. Catal. 40 (1975) 334. [79] R.J. Behm, K. Christmann, G. Ertl, M.A. Vanhove, J. Chem.
[42] D.M. Nicholas, Y.T. Shah, I.A. Zlochower, Ind. Eng. Chem. Phys. 73 (1980) 2984.
Prod. Res. Dev. 15 (1976) 29. [80] H. Pfnur, D. Menzel, F.M. Hoffmann, A. Ortega, A.M.
[43] T. Matsushima, J.M. White, J. Catal. 39 (1975) 265. Bradshaw, Surf. Sci. 93 (1980) 431.
[44] T. Matsushima, D.B. Almy, D.C. Foyt, J.S. Close, J.M. White, [81] R.J. Behm, K. Christmann, G. Ertl, M.A. Vanhove, P.A. Thiel,
J. Catal. 39 (1975) 277. W.H. Weinberg, Surf. Sci. 88 (1979) L59.
[45] R.L. Palmer, J.N. Smith, J. Chem. Phys. 60 (1974) 1453. [82] A.M. Bradshaw, F.M. Hoffmann, Surf. Sci. 72 (1978) 513.
[46] S.E. Voltz, C.R. Morgan, D. Liederma, S.M. Jacob, Ind. Eng. [83] J.C. Tracy, P.W. Palmberg, J. Chem. Phys. 51 (1969) 4852.
Chem. Prod. Res. Dev. 12 (1973) 294. [84] R.L. Park, H.H. Madden, Surf. Sci. 11 (1968) 188.
[47] I. Langmuir, Trans. Faraday Soc. 17 (1922) 672. [85a] W.K. Kuhn, J. Szanyi, D.W. Goodman, Surf. Sci. 274 (1992)
[48] T. Engel, G. Ertl, Adv. Catal. 28 (1979) 1. L611.
[49] T. Engel, G. Ertl, Chem. Phys. Solid Surf. Het. Catal. 4 (1982) [85b] J. Szanyi, W.K. Kuhn And, D.W. Goodman, J. Vac. Sci.
73.
Technol. A 11 (1993) 1969.
[50] R.R. Ford, Adv. Catal. 21 (1970) 51.
[86] D.R. Rainer, M.C. Wu, D.I. Mahon, D.W. Goodman, J. Vac.
[51] J.T. Kummer, J. Phys. Chem. 90 (1986) 4747.
Sci. Technol. A 14 (1996) 1184.
[52] J.N. Armor, Appl. Catal. A 176 (1999) 159.
[87] J. Wintterlin, S. Volkening, T.V.W. Janssens, T. Zambelli, G.
[53] J.R. Rostrup-Neilsen, Catal. Steam Reform. Sci. Technol. 5
Ertl, Science 278 (1997) 1931.
(1984) 1.
[88] T. Zambelli, J.V. Barth, J. Wintterlin, G. Ertl, Nature 390 (1997)
[54] M. Valden, S. Pak, X. Lai, D.W. Goodman, Catal. Lett. 56
495.
(1998) 7.
[89] J. Wintterlin, R. Schuster, G. Ertl, Phys. Rev. Lett. 77 (1996)
[55] M. Valden, X. Lai, D.W. Goodman, Science 281 (1998) 1647.
123.
[56] D.R. Rainer, M. Koranne, S.M. Vesecky, D.W. Goodman, J.
[90] M.J.P. Hopstaken, J.W. Niemantsverdriet, J. Chem. Phys. 113
Phys. Chem. B 101 (1997) 10769.
[57] D.W. Goodman, Chem. Rev. 95 (1995) 523. (2000) 5457.
[58] D.W. Goodman, Surf. Sci. 300 (1994) 837. [91] J.Z. Xu, J.T. Yates, J. Chem. Phys. 99 (1993) 725.
[59] X.P. Xu, D.W. Goodman, J. Phys. Chem. 97 (1993) 7711. [92] C.E. Tripa, J.T. Yates, Nature 398 (1999) 591.
[60] M. Haruta, S. Tsubota, A. Ueda, H. Sakurai, Stud. Surf. Sci. [93] C.E. Tripa, C.R. Arumaninayagam, J.T. Yates, J. Chem. Phys.
Catal. 77 (1993) 45. 105 (1996) 1691.
[61] M. Haruta, S. Tsubota, T. Kobayashi, A. Ueda, H. Sakurai, M. [94] V.I. Savchenko, G.K. Boreskov, A.V. Kalinkin, A.N. Salanov,
Ando, Stud. Surf. Sci. Catal. 75 (1993) 2657. Kinet. Catal. 24 (1983) 983.
[62] M. Haruta, Catal. Today 36 (1997) 153. [95] C.H.F. Peden, D.W. Goodman, J. Phys. Chem. 90 (1986) 1360.
[63] M. Valden, D.W. Goodman, Isr. J. Chem. 38 (1998) 285. [96] S.H. Oh, G.B. Fisher, J.E. Carpenter, D.W. Goodman, J. Catal.
[64] D.W. Goodman, R.D. Kelley, T.E. Madey, J.T. Yates, J. Catal. 100 (1986) 360.
63 (1980) 226. [97] T. Matsushima, Surf. Sci. 127 (1983) 403.
[65] D.W. Goodman, C.T. Campbell, Rev. Sci. Instrum. 63 (1992) [98] K.L. Kostov, P. Jakob, D. Menzel, Surf. Sci. 377 (1997)
172. 802.
[66] J.F. Jia, J.N. Kondo, K. Domen, K. Tamaru, J. Phys. Chem. B [99] K.H. Allers, H. Pfnur, P. Feulner, D. Menzel, J. Chem. Phys. 100
105 (2001) 3017. (1994) 3985.
[67] V.P. Zhdanov, B. Kasemo, J. Chem. Phys. 114 (2001) 5351. [100] K. Krischer, M. Eiswirth, G. Ertl, J. Chem. Phys. 96 (1992) 9161.
[68] A. Baraldi, L. Gregoratti, G. Comelli, V.R. Dhanak, M. [101] T. Fink, J.P. Dath, R. Imbihl, G. Ertl, Surf. Sci. 251 (1991) 985.
Kiskinova, R. Rosei, Appl. Surf. Sci. 99 (1996) 1. [102] M. Eiswirth, P. Moller, G. Ertl, J. Vac. Sci. Technol. A 7 (1989)
[69] G. Blyholder, J. Phys. Chem. 68 (1964) 2772. 1882.
[70] L.W.H. Leung, J.W. He, D.W. Goodman, J. Chem. Phys. 93 [103] X. Lai, T.P. St Clair, M. Valden, D.W. Goodman, Prog. Surf.
(1990) 8328. Sci. 59 (1998) 25.
[71] T. Matsushima, H. Akiyama, A. Lesar, H. Sugimura, G.E.D. [104] P.N. First, J.A. Stroscio, R.A. Dragoset, D.T. Pierce, R.J.
Torre, T. Yamanaka, Y. Ohno, Surf. Sci. 386 (1997) 24. Celotta, Phys. Rev. Lett. 63 (1989) 1416.

Вам также может понравиться