You are on page 1of 4

19 August 2002

Chemical Physics Letters 362 (2002) 181–184


www.elsevier.com/locate/cplett

Spectroscopic detection of CO dissociation on defect sites on


Ru(1 0 9): implications for Fischer–Tropsch
catalytic chemistry q
Tykhon Zubkov, Gregg A. Morgan Jr., John T. Yates Jr.*
Department of Chemistry, Surface Science Center, 234 Chevron Science Center, University of Pittsburgh,
Pittsburgh, PA 15260, USA

Received 11 March 2002; in final form 29 May 2002

Abstract

Dissociative chemistry of CO on an atomically stepped single crystal Ru(1 0 9) surface was studied using the com-
bination of vibrational spectroscopic and thermal desorption measurements. CO molecules on stepped Ru are by
desorption and dissociation together upon heating to 480 K. Contrary to the atomically smooth Ru(0 0 1) surface, the
step defect sites on the Ru(1 0 9) surface facilitate CO dissociation. Adsorbed atomic O and C then recombinatively
desorb near 520 K. This low-temperature mobile carbon species is likely to be an intermediate in the Fischer–Tropsch
synthesis on Ru. Ó 2002 Elsevier Science B.V. All rights reserved.

Metallic Ru surfaces are known to effectively mixing techniques to investigate the behavior of
catalyze the production of linear alkanes from CO the CO reactant molecule. Using the Ru(1 0 9)
and H2 in the Fischer–Tropsch (FT) reaction [1,2], single crystal surface, which contains periodic
and it is very likely that a primary step in this re- double Ru atom height step defect sites, separated
action is the dissociation of the CO bond in the by 10 atom wide smooth Ru(0 0 1) terraces, we
chemisorbed CO species. The formation of carbon have determined that the CO dissociation reaction
on the surface is followed by reaction with chem- preferentially takes place at relatively low tem-
isorbed H atoms produced by H2 dissociation on peratures (480 K) on the atomic step defect sites.
the surface. We have combined reflection-absorp- Fig. 1 shows the structure of the Ru(1 0 9) surface,
tion infrared spectroscopy (RAIR) with tempera- as determined by low energy electron diffraction
ture programmed desorption (TPD) and isotopic measurements [3]. This crystalline surface was
cleaned by conventional methods involving Arþ
bombardment, followed by annealing in oxygen
q gas, and in ultrahigh vacuum. Fig. 2 shows the
Work supported by the Department of Energy, Office of
Basic Energy Sciences.
RAIR spectrum of 13 C18 O on this surface where
*
Corresponding author. Fax: +1-412-624-6003. the C–O vibrational frequency at low coverage is
E-mail address: jyates@imap.pitt.edu (J.T. Yates Jr.). observed at 1884-1891 cm (for 12 C16 O at 1973–

0009-2614/02/$ - see front matter Ó 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 0 0 9 - 2 6 1 4 ( 0 2 ) 0 0 8 9 5 - 3
182 T. Zubkov et al. / Chemical Physics Letters 362 (2002) 181–184

absorbance and shifts the terrace-CO frequency


downward which reflects the decreasing vibra-
tional coupling on the terraces than on the steps)
[5]. Upon heating to about 480 K, RAIR shows
that both CO is completely removed. Fig. 3a
shows the TPD spectrum of CO, where surpris-
ingly it is seen that some CO desorbs above 480 K.
This CO desorption process above 480 K must
result from the recombination of C(a) and O(a),
since the RAIR experiment indicates that undis-
sociated CO is not present above 480 K. To check
Fig. 1. Schematic of the Ru(1 0 9) surface showing step and that CO dissociation is followed closely by
terrace sites. The region of the step sites is circled.
C(a) + O(a) recombination, we chemisorbed
equimolar amounts of 13 C18 O and 12 C16 O, and
1980 cm1 ). By comparison with other RAIR
studies of the atomically smooth close-packed
Ru(0 0 1) surface [4], we find that the bonding of
CO is similar and corresponds to a single coordi-
nated on-top species. Thermal depletion of ad-
sorbed CO causes the infrared band to decrease in

Fig. 3. (a) TPD of CO from Ru(1 0 9). The initial coverage is


1:3  1014 CO=cm2 ; (b) fraction of isotopic mixing, fm , between
CO isotopomers, as determined by multiplex mass spectrometry
during TPD:
X13 C16 O
Fig. 2. RAIR detection of chemisorbed 13 C18 O species on fm ¼ ;
ðX13 C16 O þ X13 C18 O ÞðX13 C16 O þ X12 C16 O Þ
Ru(1 0 9) following heating in ultrahigh vacuum. The rate of
heating to the designated temperature is dT/dt ¼ 2 K/s. The where Xi is the mole fraction of the ith isotopomer in the de-
initial coverage of CO is 5:3  1013 CO=cm2 . sorbing CO. In both cases, dT/dt ¼ 1.9 K/s.
T. Zubkov et al. / Chemical Physics Letters 362 (2002) 181–184 183

measured the fraction of isotopic mixing between In summary we have shown that Ru atomic step
the species during TPD. These results are shown in defect sites exhibit enhanced CO dissociation ac-
Fig. 3b, where it may be seen that isotopic mixing tivity compared to smooth (0 0 1) terrace sites, and
occurs extensively only during the high-tempera- also that C and/or O atomic mobility near 500 K
ture CO desorption process. For comparison, no leads to recombination and desorption of CO.
high-temperature CO desorption feature and no Fischer–Tropsch chemistry most likely involves
isotopic mixing were observed on a smooth the interaction of C(a) and O(a) species with
Ru(0 0 1) surface [6,7]. These results demonstrate adsorbed H to form linear alkane products and
conclusively that the sites associated with the Ru water.
atomic steps are selectively responsible for CO
dissociation near 480 K. Therefore, the well-
known catalytic activity of Ru for the FT reaction References
is likely due to the special behavior of Ru atoms
with low Ru coordination numbers and/or due to [1] H.H. Storch, N. Golumbic, R.B. Anderson, in: The
the special steric environment at the step vicinity Fischer–Tropsch and Related Syntheses, Wiley, Chapman
& Hall, New York, London, 1951, p. 309.
that facilitates the CO dissociation. Similar pref- [2] R.B. Anderson, in: The Fischer–Tropsch Synthesis, Aca-
erential behavior has been seen for NO dissocia- demic Press, New York, 1984, p. 110.
tion and N2 dissociative chemisorption on step [3] T.S. Zubkov, G.A. Morgan Jr., John T. Yates Jr.,
sites on Ru surfaces [8–10]. This is consistent with O. K€ uhlert, M. Lisowski, R. Schillinger, D. Fick, H.J.
J€ansch (in preparation).
the known electronic properties of low coordina-
[4] H. Pfn€ur, D. Menzel, F.M. Hoffmann, A. Ortega, A.M.
tion number sites on metal surfaces leading to Bradshaw, Surf. Sci. 93 (1980) 431.
enhanced reactivity [11]. If the CO dissociation is [5] In a separate experiment, as CO coverage on the Ru
rate-limiting, the above considerations imply that surface increased, the CO frequency shifted consistently
the FT reaction is structure-sensitive, as reported upward.
for supported Ru catalysts [12–16]. No evidence of [6] H. Pfn€ur, P. Feulner, D. Menzel, J. Chem. Phys. 79 (1983)
4613.
structure sensitivity was reported for methanation [7] E. Shincho, C. Egawa, S. Naito, K. Tamaru, Surf. Sci. 149
over two different Ru single crystal faces, (0 0 1) (1985) 1.
and (1 1 0) [17]. However both surfaces may lack [8] T. Zambelli, J. Wintterlin, J. Trost, G. Ertl, Science 273
the Ru ensembles of the proper geometry for being (1996) 1688.
[9] S. Dahl, A. Logadottir, R.C. Egeberg, J.H. Larsen,
active.
I. Chorkendorff, E. T€ ornqvist, J.K. Nørskov, Phys. Rev.
In addition, on the early transition metals Lett. 83 (1999) 1814.
[18–26], CO has been observed to dissociate, but [10] S. Dahl, E. T€ornqvist, I. Chorkendorff, J. Catal. 192 (2000)
C(a) + O(a) recombination processes occur at 381.
much higher temperatures than on the stepped Ru [11] B. Hammer, O.H. Nielsen, J.K. Nørskov, Catal. Lett. 46
surface. Since atom recombination must involve (1997) 31.
[12] M. Boudart, M.A. McDonald, J. Phys. Chem. 88 (1984)
surface mobility, it is possible that the special FT 2185, and references therein.
catalytic activity of Ru may be due to low tem- [13] A.K. Datye, J. Schwank, J. Catal. 93 (1985) 256.
perature CO dissociation coupled with C(a) or [14] K.J. Smith, R.C. Everson, J. Catal. 99 (1986) 349.
O(a) atom mobility at temperatures near 500 K. [15] J.C. Kelzenberg, T.S. King, J. Catal. 126 (1990) 421, and
By forcing CO dissociation on a stepped Ru sur- references therein.
[16] M.R. Prairie, A. Renken, J.G. Highfield, K.R. Thampi,
face with high CO exposures at 501 K, Tamaru M. Gr€atzel, J. Catal. 129 (1991) 130, and references therein.
et al. [7] have also observed effects attributed to C– [17] R.D. Kelley, D.W. Goodman, in: D. King, D.P. Woodruff
O bond scission at the steps. However, their TPD (Eds.), The Chemical Physics of Solid Surfaces and
data differ considerably from ours, and their ul- Heterogeneous Catalysis, vol. 4, Elsevier, Amsterdam,
1982, p. 427.
traviolet photoelectron spectra (UPS) mainly in-
[18] N.D. Shinn, T.E. Madey, Phys. Rev. Lett. 53 (1984) 2481.
dicate that CO can be forced to dissociate at high [19] N.D. Shinn, T.E. Madey, J. Chem. Phys. 83 (1985) 5928.
temperatures and CO exposures, blocking CO [20] N.D. Shinn, T.E. Madey, J. Vac. Sci. Technol. A 3 (1985)
adsorption at the step sites. 1673.
184 T. Zubkov et al. / Chemical Physics Letters 362 (2002) 181–184

[21] C. Benndorf, B. Kr€ uger, F. Thieme, Surf. Sci. 163 (1985) [24] F. Zaera, E. Kollin, J.L. Gland, Chem. Phys. Lett. 121
L675. (1985) 464.
[22] D.W. Moon, S.L. Bernasek, J.-P. Lu, J.L. Gland, P.J. [25] J.G. Chen, M.L. Colaianni, W.H. Weinberg, J.T. Yates
Dusyer, Surf. Sci. 184 (1987) 90. Jr., Chem. Phys. Lett. 177 (1991) 113.
[23] J.P. Lu, M.R. Albert, S.L. Bernasek, Surf. Sci. 217 (1989) [26] M.L. Colaianni, J.G. Chen, W.H. Weinberg, J.T. Yates
55. Jr., J. Am. Chem. Soc. 114 (1992) 3735.

You might also like