Sie sind auf Seite 1von 12

Theoretical studies on the adsorption and decomposition

of H
2
O on Pd(111) surface
Yilin Cao, Zhao-Xu Chen
*
Lab of Mesoscopic Chemistry, Institute of Theoretical and Computational Chemistry, School of Chemistry and Chemical Engineering,
Nanjing University, Nanjing 210093, Peoples Republic of China
Received 27 April 2006; accepted for publication 14 July 2006
Available online 10 August 2006
Abstract
To provide information about the chemistry of water on Pd surfaces, we performed density functional slab model studies on water
adsorption and decomposition at Pd(111) surface. We located transition states of a series of elementary steps and calculated activation
energies and rate constants with and without quantum tunneling eect included. Water was found to weakly bind to the Pd surface. Co-
adsorbed species OH and O that are derivable from H
2
O stabilize the adsorbed water molecules via formation of hydrogen bonds. On the
clean surface, the favorable sites are top and bridge for H
2
O and OH, respectively. Calculated kinetic parameters indicate that dehydro-
genation of water is unlikely on the clean regular Pd(111) surface. The barrier for the hydrogen abstraction of H
2
O at the OH covered
surface is approximately 0.20.3 eV higher than the value at the clean surface. Similar trend is computed for the hydroxyl group
dissociation at H
2
O or O covered surfaces. In contrast, the OH bond breaking of water on oxygen covered Pd surfaces,
H
2
O
ad
+ O
ad
!2OH
ad
, is predicted to be likely with a barrier of 0.3 eV. The reverse reaction, 2OH
ad
!H
2
O
ad
+ O
ad
, is also found
to be very feasible with a barrier of 0.1 eV. These results show that on oxygen-covered surfaces production of hydroxyl species is highly
likely, supporting previous experimental ndings.
2006 Elsevier B.V. All rights reserved.
Keywords: Density functional calculations; Adsorption; Decomposition; Water; Palladium
1. Introduction
Water is one of the most popular substances on our pla-
net, it involves many physical and chemical processes such
as corrosion, catalysis, electrochemistry, photo-conversion,
material science, tribology and membrane science, and
plays important roles in hydrogen production, fuel cells,
biological sensors and so on [1,2]. During the past several
decades, tremendous eorts have been devoted to water
chemistry on surface [1,2]. As a mild oxidant, water always
determines the product distribution in many chemical pro-
cesses [39]. We are, in particular, interested in water
adsorption and dissociation on Pd surfaces because water
is involved in several important reaction systems. For
example, in the indirect route of partial oxidation of meth-
ane, reaction CH
4
+ H
2
O !CO + 3H
2
is suggested to be
an important step over Pd to produce syngas, the feedstock
for large volume processes of methanol production and
FisherTorpsch synthesis [7,8]. Water can also eliminate
the carbon deposition on palladium through the reactions
H
2
O + CO !CO
2
+ H
2
and C + H
2
O !CO + H
2
in the
production of propylene through propane decomposition
[5,6]. In the presence of water, propylene can be selectively
oxidized to acrolein and acrylic acid while only CO
2
is pro-
duced in the absence of water [9]. Undoubtedly, the knowl-
edge about how water participates in these processes,
molecularly or in the form of decomposed species, is help-
ful and prerequisite for raveling the mechanisms of these
reactions. However, fewer studies have been conducted to
address the behavior of water, and some foundational
0039-6028/$ - see front matter 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.susc.2006.07.028
*
Corresponding author. Fax: +86 25 83686553.
E-mail address: zxchen@nju.edu.cn (Z.-X. Chen).
www.elsevier.com/locate/susc
Surface Science 600 (2006) 45724583
questions remain to be answered. For example, how do
water and species derivable from water adsorb on Pd sur-
faces? Is water readily to be dissociated on the surfaces?
If yes, what is the microscopic picture of the OH bond
activation?
Concerning adsorption of water adsorption on Pd sur-
faces, experiments indicate that water binds weakly to Pd
surfaces with the desorption temperatures lower than
200 K, and the watersubstrate interaction is found to be
enhanced by the co-adsorbed oxygen [10]. EELS (electron
energy loss spectroscopy) experiments [11] further show
that water adsorbs via the oxygen atom with its molecular
axis signicantly tilted relative to the surface normal on
Pd(100) surface. STM (scanning tunneling microscopy)
measurements suggest that isolated water molecules ad-
sorbed on the top site of Pd atom at low temperatures
and coverage [12]. Formation of water clusters induced
by the molecular collision and hopping is expected. The
particularly stable structure is found to be the cyclic
hexamer structure [12]. Dissociative adsorption has not
been detected on clean Pd(111) surface [13]. At variance,
formation of OH group from water on oxygen pre-covered
Pd(111) is observed with thermal desorption spectroscopy
(TDS) [13]. Wolf et al. suggest that water can be activated
by the co-adsorbed oxygen atom and the formation of
OH group on Pd(111) surface is through the reaction
H
2
O + O !2OH [14,15]. This assumption is in agreement
with the isotope exchange experiments [16]. This reaction,
according to the TPD measurements, starts at approxi-
mately 175 K on polycrystalline Pd-lms [17] and the
formed OH group immediately recombines with another
OH group to yield H
2
O. The experimental conclusion that
water interacts weakly with the Pd surface is consistent
with the small adsorption energy and slight deformation
of molecular geometry upon adsorption predicted theoret-
ically [18]. Slab model and density functional calculations
further reveal that water molecule preferentially adsorbs
on the top site at Pd(111) surface with the molecular plane
almost parallel to the surface [12,19]. Studies of water
adsorption at higher coverage show that the H-down con-
guration is favored by 0.040.16 eV/H
2
O over the H-up
one (for the denition of H-down and H-up congurations,
see Section 3.1.1 for details) [19,20].
To the best of our knowledge, theoretical studies of
water on Pd surfaces are mainly focused on the adsorption.
Fewer investigations of co-adsorption and dissociation of
water on Pd surfaces have been reported. In this paper,
aiming to address the OH bond activation and the possi-
ble co-adsorbate eects on the adsorption and dehydroge-
nation of water, we performed a density functional study
using the ideal Pd(111) model system. The whole paper
is arranged as follows. Following this introduction, we de-
scribe the computational method and models we used in
this study (Section 2). Then we present our results and dis-
cussions for adsorption of H
2
O and OH on various sur-
faces, and stepwise decomposition of water in dierent
situations (Section 3). Thermodynamic and kinetic
comparisons of each elementary step are also presented
in Section 3. Final remarks are given in Section 4.
2. Computational details and models
Total energy calculations were performed within the
density functional theory (DFT) framework using the
VASP code [2123]. Exchange and correlation eects were
described by the PerdewWang in 1991 generalized gradi-
ent approximation [24]. The projector augmented-wave
method was used to represent the inner cores [25], and
the electron states were expanded in a plane wave truncated
at a cuto energy of 400 eV. Pd(111) surface was modeled
with a periodic array of 4-layer slab, separated by a vac-
uum region equivalent to 7 layer thickness. Our calculated
binding energy of water at atop site at 1/3, 1/4 and 1/9 cov-
erages diers at most by 0.06 eV. Therefore, in this paper
we choose a p(2 2) surface unit cell which corresponds
to 1/4 coverage for a single adsorbate to model separate
adsorption and decomposition. (By separate adsorption,
we mean there is only one adsorbate in one p(2 2) surface
unit cell. Co-adsorption implies that there are two species,
A and B, in the cell. In the case of 1/2 coverage of H
2
O or
OH, A and B denote the same species H
2
O or OH.) Monk-
horstPack [26] meshes with 5 5 1 k-point sampling
within the surface Brillouin zone were adopted using Meth-
fesselPaxton method with a smearing width of 0.15 eV
[27]. Adsorbates are placed on one side of the slab at the
bulk-truncated geometry.
The binding energy (E
A
ads
) for the separate adsorption of
an adsorbate (A) on the substrate was calculated with (1)
E
A
ads
E
slab
E
A
E
A=slab
1
where E
A/slab
denotes the total energy of the slab with the
adsorbate on it, E
A
is the energy of the adsorbate in the
gas phase, and E
slab
represents the energy of the clean slab.
In the case of co-adsorption of A with B, we used (2) to
estimate the binding energy of A
E
A
ads
E
B=slab
E
A
E
AB=slab
2
E
A
ads
and E
A
in (2) have the same meaning as stated above.
E
A+B/slab
is the total energy of the slab with A and B co-
adsorbed on it, and E
B/slab
refers to the energy of the slab
covered with B in the co-adsorption conguration.
To estimate quantitatively the inuence of co-adsor-
bates, we decomposed the binding energy into following
two components, E
AB
and E
A
adssub
. The former reects
the interaction between adsorbates A and B, which is calcu-
lated with (3). The latter measures the interaction of the
adsorbate A with the substrate, which is estimated using (4)
E
AB
E
A
E
B
E
A=B
3
E
A
adssub
E
A
ads
E
AB
4
E
A/B
in (3) is the total energy of the adsorbates A and B in
the gas phases at their co-adsorbed structures; E
A
and E
B
are the energies of A and B in the gas phase, respectively.
Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583 4573
Transition states were located using the climbing image
nudged elastic band method [28], and veried by vibra-
tional mode analyses. (There is only one imaginary fre-
quency in all the transition states corresponding to the
forming and breaking of the OH bond.) The reaction rate
constant k was evaluated according to the transition state
theory based on harmonic approximation [29]:
k
k
B
T
h
q

q
e
E

E=RT
A
0
e
Ea=RT
5
Here h, R and k
B
are the Planck constant, the universal gas
constant and the Boltzmann constant, respectively. E
*
and
E are the energies of the transition state (TS) and initial
state (IS), respectively. q
*
and q are the partition function
corresponding to the TS and IS, respectively. For more
information about the calculations of q
*
and q, please refer
to Ref. [30]. E
a
is the activation energy including the clas-
sical or quantum zero point energy (ZPE) correction
[31,32]. The transmission ecient of quantum tunneling ef-
fect, C, is estimated using the Wigner correction scheme
[33]:
C 1
1
24
ht

k
B
T

2
1
1
6
pT
c
T

2
6
where T
c
= ht
*
/2pk
B
is crossover temperature, t
*
is the
imaginary frequency of the transition state.
3. Results and discussions
Adsorption structures are prerequisite for our subse-
quent kinetic investigations of water and hydroxyl group
decomposition. Thus, we rst studied H
2
O adsorption on
clean Pd(111) as well as oxygen or OH covered Pd(111)
surfaces. Then we investigated the adsorption of OH group
on clean Pd(111) and oxygen covered Pd(111). These sys-
tems are relevant to water and hydroxyl group decomposi-
tion in dierent situations considered in this paper. Usually
a system has several adsorption structures which are local
minima on the potential energy surfaces. Unless otherwise
stated, only the results for the most favorable structures are
reported here.
3.1. Adsorption of H
2
O and OH on various Pd(111)
surfaces
3.1.1. Adsorption of H
2
O on clean Pd(111) surface
STM experiments indicate that water adsorbs at a top
site on Pd(111) via the oxygen atom [12]. The angle be-
tween the water molecular plane and the substrate surface,
dened as a in this paper, is theoretically predicted to be
7 [18] and 20 [20]. We further scanned the potential
energy surface (PES) with respect to the angle a at the
optimized geometry of water on the top position. Our ob-
tained a value at the top site is 5.5, close to 7 as reported
in Ref. [18]. As revealed in Fig. 1, the PES around 5.5 is
rather at. Hence, most likely the adsorbed water mole-
cules adopt various orientations with a around 5.5. We
also calculated the a values at the bridge and three fold hol-
low fcc (hcp) sites to be 0 and 27 (27), respectively,
indicating that the molecular plane is parallel to the sub-
strate surface at the bridge site whereas at the hollow sites
the H ends tilt down towards the substrate surface. The cal-
culated binding energies (E
ads
) are: 0.22 (top) >0.13
(bridge) >0.11 eV (fcc, hcp), in accordance with the previ-
ous theoretical nding that the most favorable site is the
top position [20]. The E
ads
at the most favorable (top) site,
0.22 eV, is somewhat smaller than the previous results
ranging from 0.29 to 0.33 eV [18,20,34]. Our test calcula-
tions demonstrated that the discrepancy between the previ-
ous and present values is mainly due to the surface
coverage, 1/4 in our case vs. 1/9 adopted in the previous
papers [18,20,34]. Based on the TPD experimental data
[17], Redhead analysis [35] yields a binding energy of
0.45 eV which is twice as large as our result of 0.22 eV.
We conjectured this discrepancy to be due to surface
defects. To verify it, we calculated the binding energy at
the step edge of Pd(221) surface. The estimated result,
0.51 eV, is very close to the experimental value of
0.45 eV. Note, even the largest binding energy (0.22 eV at
atop site) is less than 0.3 eV, and comparable to the
strength of hydrogen bonding. Hence water adsorption
on Pd(111) is a physisorption process, and clustering or
formation of network via hydrogen bonding can be ex-
pected. Consistent with the weak adsorbate-substrate inter-
action, the geometry of adsorbed water changes very
slightly: the OH bond length and the bond angle between
the two OH bonds, \
HOH
, are computed to be 0.98 A

and
105.1 at the top site (Fig. 2a), respectively, compared to
0.97 A

and 104.6 in the gas phase (Table 1). The OPd dis-
tance is found to be 2.45 A

. These results are in good agree-


ment with the reported theoretical results: 0.98 A

for the
OH distance, 105 for the \
HOH
, and 2.42 A

for the
OPd distance [18,20].
Fig. 1. Energy variations with the angle (a) between the water molecular
plane and the substrate Pd(111) surface.
4574 Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583
As mentioned above, the binding energy of water at low
coverage is comparable to the hydrogen bonding strength,
and water molecules may form a series of structures linked
by hydrogen bonds at high coverage. Indeed, STM investi-
gations found that the hexagonal ring is the most stable
structure [12]. When the surface coverage increases, the
hexagonal ring grows into the so-called ordered honey-
comb structure with

3
p

3
p
R30

periodicity relative
to the Pd(111) surface [12]. In such structures (bilayer
structures) there are two sorts of water molecules classied
by dierent heights of the oxygen atoms. One, denoted as
H
2
O
L
hereafter, is closer to the substrate surface and inter-
acts with the substrate via the oxygen atoms. The molecu-
lar planes of H
2
O
L
are basically parallel to the substrate
surface. Another type of water molecules (referred to as
H
2
O
H
) sits a bit far away from the substrate with the
molecular planes perpendicular to the substrate surface.
There are three hydrogen bonds formed between each pair
of H
2
O
H
and H
2
O
L
. For H
2
O
L
, all the two H atoms are in-
volved in hydrogen bonding while in H
2
O
H
only one H
atom participates in the forming of a hydrogen bond.
The remaining H atom of H
2
O
H
can either direct away
from or point to the substrate surface, which are called
H-up or H-down congurations (Fig. 2b), respectively.
At 2/3 coverage, the H-down structure is calculated to be
preferred over the H-up conguration by 0.05 eV, with
the mean binding energy of 0.55 eV for the former
[19,20]. (The mean value of binding energy is dened as
E
avr
0:5E
H
2
OH
2
O=Pd
E
Pd
2E
H
2
O
, where E
H
2
OH
2
O=Pd
,
E
Pd
and E
H
2
O
are the energies of adsorbed system, clean
Fig. 2. Adsorption of H
2
O and OH on Pd (111) surface (a) H
2
O/Pd(111), (b) H
2
O + H
2
O/Pd(111), (c) H
2
O + O/Pd(111), the most favorable co-
adsorption structure. (d) H
2
O + O/Pd(111), the next most favorable co-adsorption structure. (e) H
2
O + OH/Pd(111), (f) OH/Pd(111), (g) OH + OH/
Pd(111), (h) OH + O/Pd(111). Red sphere: O atom; white sphere: H atom; yellow sphere: Pd atoms on the top layer; Golden sphere: Pd atoms in the
second layer. Inlets in (b) and (e) show the cyclic structures. (For interpretation of the references in colour in this gure legend, the reader is referred to the
web version of this article.)
Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583 4575
slab and water in gas phase, respectively.) This averaged
value is close to our result of 0.52 eV at the same coverage.
Both H
2
O
H
and H
2
O
L
are approximately on the top sites
with the heights of the oxygen atoms being 2.58 A

in
H
2
O
L
and 2.96 A

in H
2
O
H
, slightly smaller than 2.66 A

and 3.18 A

in Ref. [20]. Because of the hydrogen bonds,


the tilting angle a is computed to be 28 for H
2
O
L
, 22.5
larger than the one at 1/4 coverage.
We also studied the H-down and H-up congurations at
1/2 coverage. The averaged binding energy is estimated to
be 0.43 eV for the former and 0.36 eV for the latter, giving
the same stability sequence as at 2/3 coverage. The mean
binding energy at 1/2 coverage is 0.1 eV smaller than at
2/3 coverage, due to, obviously, the weakened hydrogen
bonding as indicated by the longer hydrogen bonds:
2.12 A

and 2.37 A

at 1/2 coverage and 1.85 A

and 1.84 A

at 2/3 coverage. In the H-down structure (Fig. 2b), H


2
O
L
sits at a top site and H
2
O
H
at a bridge position. The oxygen
heights are 2.58 A

and 2.97 A

, respectively. The OH bond


lengths are 0.98 A

in H
2
O
L
and 0.99 A

in H
2
O
H
, essentially
the same as at the 1/4 coverage, 0.98 A

. It should be men-
tioned that other water cluster patterns composed of
mainly at lying water molecules arranged in planar hexa-
gons with the similar stability to the bilayer structure on
Pd(111) have been reported [36]. Based on the density
functional calculations and STM simulations, a novel
mechanism of water overlayer growth has been presented,
which explains the experimentally observed patterns very
well [36].
3.1.2. H
2
O adsorption on O/Pd(111)
When adsorbed separately, H
2
O prefers atop site (see
above). Dierent site preferences of atomic O on Pd(111)
are reported [37,38]. Steltenpohl and the co-worker sug-
gested atomic oxygen favors an hcp site whereas Rose
et al. argued that O atom mainly occupied an fcc position
with a diusion barrier of about 0.40.5 eV to other sites.
We have scrutinized various co-adsorption congurations
with water on a top site and oxygen atom at the fcc, hcp,
bridge and top sites, respectively. The most favorable co-
adsorption state is the structure with water on the atop site
and oxygen atom on the fcc position (Fig. 2c). In this struc-
ture, the oxygen atom of water is 2.37 A

above the surface.


The OH bond length and the \
HOH
are 0.98A

and 105.9,
respectively. The tilting angle a is 10.4. The co-adsorbed O
atom (denoted as O
B
) sits at the fcc site with the PdO
B
bond length of 2.00 A

. The binding energy of water in this


structure is calculated to be 0.34 eV (Table 1). In the next
most favorable co-adsorption complex, H
2
O is still on a
top site and the O
B
atom occupies an hcp sit (Fig. 2d). This
structure is only 0.2 eV less stable than the most favorable
one. The calculated E
ads
of H
2
O in this system H
2
O + O/
Pd(111), 0.42 eV, is about twice as large as that on clean
Pd(111) (0.22 eV) (Table 1), in agreement with the higher
desorption temperatures of water on oxygen covered
Pd(111) than at the clean surface [10]. The estimated
adsorbateco-adsorbate and adsorbate-substrate interac-
tion, E
AB
and E
adssub
, are 0.17 and 0.25 eV, respectively,
showing that H
2
O is mainly stabilized by its interaction
Table 1
Energetic properties and geometrical parameters of various systems at Pd(111) surfaces
a
System A + B
b
H
2
O
c
H
2
O H
2
O + O
d
H
2
O + O
e
H
2
O
L
+ H
2
O
H
f
d
A
OH
=d
B
OH
0.97/ 0.98/ 0.98/ 0.98/ 0.98/0.99
d
OAHB
/d
OBHA
/ / /2.59 /2.60 2.12/2.37
d
OAPd
/d
OBPd
/ 2.45/ 2.39/2.04 2.37/2.00 2.59/3.31
a
A
/a
B
/ 5.5/ 2.8/ 10.4/ 22.0/90.0
\
A
HOH
=\
B
HOH
104.6/ 105.1/ 106.2/ 105.9/ 105.4/102.2
E
A
adssub
=E
B
adssub
/ 0.22/ 0.25/4.16 0.23/4.40 0.26/0.17
E
A
ads
=E
B
ads
/ 0.22/ 0.42/4.33 0.34/4.51 0.79/0.70
E
AB
0.17 0.11 0.53
OH
c
OH OH + O OH + OH H
2
O + OH
d
A
OH
=d
B
OH
0.99/ 0.98/ 0.98/ 0.99/0.99 0.98/0.99
d
OAHB
/d
OBHA
/ / -/2.98 2.01/2.01 2.25/2.47
d
OAPd
/d
OBPd
/ 2.14/ 2.15/1.99 2.15/2.15 2.57/2.22
a
A
/a
B
/ 25.4/ 34.3/ 25.5/25.5 0/27.2
\
A
HOH
=\
B
HOH
/ / / / 106.1/
E
A
adssub
=E
B
adssub
/ 2.41/ 1.73/3.68 2.34/2.34 0.16/2.29
E
A
ads
=E
B
ads
/ 2.41/ 1.95/3.90 2.66/2.66 0.57/2.70
E
AB
0.22 0.32 0.41
a
d
i
OH
: the OH bond length in adsorbate i; a
i
: tilting angle of the molecular plane (i = H
2
O) or the bond axis (i = OH) relative to the substrate surface;
\
i
HOH
: bond angle formed by two OH bonds if i is water; E
i
ads
and E
i
adssub
: binding energy of adsorbate i and interaction energy of the adsorbate with the
substrate, respectively, d
OiHj
: the OH distance between the O of i and the H of j; d
OiPd
: the shortest OPd distance between the O atom of i and the Pd
atom; E
AB
: interaction energy between A and B. All energies are in the unit of eV, distance in A

and angles in degree.


b
A and B refer to the adsorbates contained in a p(2 2) surface unit cell.
c
In the gaseous phases.
d
The structure with water on atop site and O on an hcp position.
e
The structure with water and O on atop and fcc sites, respectively.
f
H
2
O
L
and H
2
O
H
denote water molecules close to and far away from the substrate surface, respectively. See Section 3.1.1.
4576 Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583
with the substrate. The OPd distance of 2.39 A

in
H
2
O + O/Pd(111) is 0.06 A

shorter than the one in H


2
O/
Pd(111) (Table 1), consistent with the larger E
adssub
(0.25 eV) of the former than 0.22 eV of the latter. The
O
B
-H distance (d
OBHA
) is 2.59 A

, 0.220.47 A

longer than
the corresponding values in H
2
O + H
2
O/Pd(111), in
accordance with the smaller E
AB
(0.17 eV), compared to
0.53 eV of the latter. Owing to the existence of O
B
, the
molecular axis rotates by 23 (clockwise) around the O cen-
ter so that one of H atoms directly points to O
B
, and the a
value decreases to 2.8, compared to 5.5 in H
2
O/Pd(111).
The OH bond remains to be 0.98 A

as in H
2
O/Pd(111)
while the \
HOH
increases from 105.1 to 106.2 (Table 1).
3.1.3. Co-adsorption of H
2
O with OH on Pd(111)
In the favorable co-adsorption conguration, H
2
O and
OH form a hexagonal ring structure (Fig. 2e). This top
view structure looks like the bilayer structures of H
2
O
L
+ -
H
2
O
H
/Pd(111) where H
2
O
H
is replaced by an OH group.
Each H
2
O molecule forms three hydrogen bonds with the
neighboring OH groups with the bond distances being
2.25 and 2.47 A

, respectively (Table 1). The oxygen atoms


denoted as O
A
for that of H
2
O and O
B
for the one of OH
group are in dierent heights from the substrate, and be-
cause OH groups interact more strongly with the substrate
than H
2
O molecules, the height for O
B
, 1.70 A

, is lower
than 2.54 A

for O
A
(not shown in Table 1). d
OAPd
and
d
OBPd
are 2.57 A

and 2.22 A

, respectively, which are


0.12 A

and 0.08 A

longer than the corresponding values


in H
2
O/Pd(111) and OH/Pd(111) (Table 1). The \
HOH
of H
2
O is 106.1, 1.0 larger than that in H
2
O/Pd(111)
and comparable to the one in H
2
O + O/Pd(111), 106.2.
The hydrogen bond distances (d
OAHB
and d
OBHA
) are
2.25 A

and 2.47 A

, 0.1 A

longer than those in


H
2
O + H
2
O/Pd(111) at 1/2 coverage, in line with the smal-
ler E
AB
of 0.41 eV for H
2
O + OH/Pd(111) than 0.53 eV
for H
2
O + H
2
O/Pd(111) (Table 1). The adsorption energy
of water, 0.57 eV, is 0.35 eV higher than that of water on
clean surface (Table 1). Owing to the hydrogen bond which
links the H
A
with the low lying O
B
, the angle a of H
2
O de-
creases to 0, compared to 5.5 in H
2
O/Pd(111). The esti-
mated interaction of water with the substrate, (E
adssub
=
0.16 eV), is the smallest among all the systems studied in
this paper (Table 1). On the other hand, the adsorption en-
ergy of the OH group is calculated to be 2.70 eV, 0.29 eV
higher than at the clean Pd(111) surface (Table 1). Accord-
ing to our energy decomposition scheme, the interaction
of OH with the substrate decreases slightly (from 2.41 eV
in OH/Pd(111) to 2.29 eV in H
2
O + OH/Pd(111))
(Table 1). However, the adsorbateco-adsorbate interac-
tion contributes 0.41 eV to the adsorption energy of OH,
which makes the OH in H
2
O + OH/Pd(111) more stable.
3.1.4. Adsorption of OH on clean Pd(111) surface
EEL experiments suggest that hydroxyl group binds to
Pd(100) surface through the Oatom with the molecular axis
signicantly tilted relative to the substrate surface normal
[39]. We studied the adsorption of OH via the oxygen end
at the top, bridge, hcp and fcc sites. The optimal tilting angle
(a) of the OH bond axis with respect to the substrate surface
are 19.4 (top), 25.4 (bridge), 42.8 (fcc) and 59.0 (hcp),
respectively. The corresponding binding energies are calcu-
lated to be 2.22 eV (top), 2.41 eV (bridge), 2.27 eV (fcc) and
2.31 eV (hcp), showing that the most favorable position is
the bridge site which is 0.100.19 eV more stable than the
other three sites. At variance, cluster model calculations
suggested the site preference to be the top site with a binding
energy of 2.02 eV [40]. This disagreement is likely owing to
the nite size eect suered in the cluster model. It is worthy
to mention that the binding energy dierence of OH group is
quite small (< 0.20 eV). Hence the potential energy surface
of OH adsorption on Pd(111) is rather at, implying a large
mobility of OH on the surface. At the most favorable bridge
site (Fig. 2f), the OH bond is slightly shortened by 0.01 A

,
and the OPd contact is found to be 2.14 A

(Table 1).
We also studied the OH adsorption at 1/2 coverage
(Fig. 2g). In the most favorable structure the two OH
groups are identical (Table 1) and sit at bridge sites. The tilt-
ing angles a are found to be 25.5. d
OPd
is 2.15 A

. d
OAHB
and d
OBHA
are the same, 2.01 A

. The binding energy of


the hydroxyl group (E
B
ads
) is found to be 2.66 eV, 0.25 eV
higher than that for OH/Pd(111), but essentially the same
as the one in H
2
O + OH/Pd(111) (Table 1). E
adssub
is
estimated to be 2.34 eV, 0.07 eV smaller than that of
OH/Pd(111) (Table 1).
3.1.5. OH adsorption on O/Pd(111)
Fig. 2h shows the most favorable conguration of OH
co-adsorbed with O where the OH group occupies a bridge
site and the co-adsorbate O atom sits at an fcc site. The
shortest distance between the O atom of hydroxyl group
(O
A
) and the Pd atom is 2.15 A

, essentially the same as


in OH/Pd(111), 2.14 A

. The calculated tilting angle a of


the OH is 34.3, about 9 larger than that in the OH/
Pd(111)(Table 1). The closest contact between the co-ad-
sorbed O atom (O
B
) and the hydrogen of OH group,
d
OBHA
, is 2.98 A

. The binding energy of OH, 1.95 eV, is


0.46 eV smaller than that on clean Pd(111). According to
the calculated E
adssub
, this is due to the weakened interac-
tion of the OH group with the substrate, 1.73 eV, com-
pared to 2.41 eV at the clean Pd(111). On the other
hand, the estimated E
AB
is positive, 0.22 eV, implying that
the direct interaction between the adsorbate and the co-
adsorbate stabilizes the hydroxyl group.
In the above, we have discussed the adsorption of H
2
O
and OH on various surfaces. To sum up, the stability of
H
2
O is enhanced on the surfaces covered with O. Co-ad-
sorbed OH and H
2
O also facilitate the stabilization of
water at certain range of coverages due to the formation
of hydrogen bonds between adsorbates. Each hydrogen
bond contributes 0.10.2 eV to the adsorption energy.
Co-adsorption tends to weaken the adsorbate-substrate
interaction, which makes the OH group less stable on oxy-
gen covered Pd(111) surface than on the clean surface.
Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583 4577
3.2. Water decomposition on Pd(111) surface
In this section, we report the water decomposition on
the clean and oxygen covered Pd(111) surfaces, reactions
(1)(4) (see below). Inuences of co-adsorbed OH and
H
2
O on the OH bond breaking of H
2
O and OH, respec-
tively, are examined by studying the corresponding OH
bond cleavage in a complex system of H
2
O + OH. With
this system, two OH bond breaking processes can be envi-
sioned: H
2
O
ad
+ OH
ad
!2OH
ad
+ H
ad
(5) and OH
ad
+
H
2
O
ad
!O
ad
+ H
ad
+ H
2
O
ad
(6), which represent the OH
bond scission of H
2
O and OH, respectively. In all the cases
considered here, we assume that the reactions follow the
LangmuirHinshelwood mechanisms. All the initial and
nal states are characterized to be local minima on poten-
tial energy surface by vibrational analysis (no imaginary
frequency).
1 H
2
O
ad
!OH
ad
H
ad
2 OH
ad
!O
ad
H
ad
3 H
2
O
ad
O
ad
!2OH
ad
4 OH
ad
O
ad
!2O
ad
H
ad
5 H
2
O
ad
OH
ad
!2OH
ad
H
ad
6 OH
ad
H
2
O
ad
!O
ad
H
ad
H
2
O
ad
3.2.1. Stepwise dehydrogenation of water
on clean Pd(111) surface
This includes reactions (1) and (2). Reaction (1) stimu-
lates the OH bond breaking of H
2
O at low coverage.
The initial state (IS) of reaction (1) has been described in
Section 3.1.1 (Fig. 3a). The activation energy of reaction
(1) (as well as (5), see Section 3.2.3) is much higher than
the adsorption energy of water (See Section 3.3). Therefore,
water will desorb, rather than dissociate in these cases. For
the sake of comparison, in the following we still present the
geometries concerning reactions (1) and (5). At the begin-
ning of the reaction, the dissociated H atom (denoted as
H
d
hereafter) moves towards an fcc hollow site, leading to
an elongated OH
d
bond. In the transition state (TS), the
OH
d
bond is extended from 0.98 A

in the IS to 1.68 A

.
The O atom shifts slightly away from the top site in the
IS to an o-top position, with the OPd distance (d
OPd
)
of 2.07 A

. The H
d
atom resides practically on the fcc site
with the shortest H
d
Pd bond length (d
HdPd
) being
1.81 A

, 0.85 A

shorter than in the IS, manifesting an en-


hanced interaction between H
d
and the Pd atom (Fig. 3a).
After the TS, the OH group shifts towards a bridge site.
In the nal state (FS), The H
d
atom occupies the fcc posi-
tion, and the H
d
Pd bond length, 1.79 A

, is slightly shrunk
(by 0.02 A

), compared to that in the TS. The produced OH


group is located at the bridge site with the shortest OPd
contact being 2.15 A

. We have also checked the OH bond


breaking of H
2
O at 1/2 coverage. It is found that the reac-
tion barrier (1.10 eV) is close to that at 1/4 coverage. How-
ever, it is thermodynamically less favorable, due to the
hydrogen bonds formed between water molecules.
The subsequent step of water stepwise dehydrogenation
is reaction (2). The most favorable adsorption congura-
tion of OH group is chosen to be the IS (see Section
3.1.4 and Fig. 3b). The reaction starts with the tilting of
the OH bond to a Pd atom, which shortens the distance
between the H and the Pd (d
HdPd
), and thus enhances
the interaction between these two atoms and simulta-
neously elongates and weakens the OH bond. Simulta-
neously, the O atom moves from the previous bridge site
towards an fcc site. In the TS, the O atom resides at the
fcc site at a height of 1.35 A

. The OPd bond length is


2.06 A

. The H atom occupies the o-top site with a height


of 1.46 A

. The OH and the HPd bond lengths are 1.46 A

Fig. 3. Schematic illustration (top view) of the OH bond breaking of H


2
O and OH on clean Pd(111) surface. (ab) correspond to reaction (1) and (2),
respectively. Red sphere: O atom; white sphere: H atom; yellow sphere: Pd atoms on the top layer; golden sphere: Pd atoms in the second layer. (For
interpretation of the references in colour in this gure legend, the reader is referred to the web version of this article.)
4578 Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583
and 1.71 A

, respectively (Table 2). After the TS, the H


atom moves further from the o-top position to a neigh-
boring fcc site. The O remains at the fcc site. In the FS,
the OPd and HPd bond distances reach 1.99 A

and
1.77 A

, respectively. The broken OH bond is elongated


to 2.82 A

, 1.36 A

longer than in the TS.


3.2.2. Stepwise dehydrogenation of water on O/Pd(111)
surface
This includes reactions (3) and (4). Two initial states for
reaction (3) are examined. The rst IS is the most favorable
co-adsorption structure of water and atomic oxygen where
H
2
O is on a top site and O atom on an fcc position (Fig. 2c,
Section 3.1.2). With this IS, the barrier is calculated to be
0.59 eV, 0.25 eV higher than the E
ads
of water, 0.34 eV.
Thus this reaction channel is less competitive to water
desorption. Therefore, we discard further discussions of
this reaction path. On the other hand, the relatively low
diusion barrier, 0.40.5 eV on the clean Pd(111) [38], ren-
ders it feasible for atomic O to move easily on the surface.
In light of this fact, we studied reaction (3) with the second
IS, i.e. the next most favorable water and atomic oxygen
co-adsorption system (H
2
O and O are on atop and hcp
sites, respectively, Fig. 2d). The hydrogen atom directing
to the co-adsorbed oxygen atom (O
B
) is chosen to be the
dissociated H (H
d
) (Fig. 4a). In the course of the reaction,
the H
d
atom and the O
B
atom get close to each other,
which gradually shortens d
OBHd
and elongates the d
HdO
.
In the TS, d
HdO
increases from 0.98 A

in the IS to
1.36 A

(Table 2). The O


B
sits at a bridge site with the
d
OBHd
decreased from 2.59 A

in the IS to 1.10 A

in the
TS. The O atom of water almost remains at the initial
top site (Fig. 4a), with the d
OPd
of 2.11 A

, compared to
2.39 A

in the IS. The d


OBPd
is essentially unchanged as
in the IS (2.00 A

) while the height of the O


B
is 1.53 A

, in-
creased by 0.28 A

. After the TS, the OH group derived


from H
2
O moves towards a bridge site. In the FS, all the
OH groups reside on the bridge positions (Fig. 4a). Key
parameters about the FS are: d
HdPd
= 2.55 A

; d
OPd
=
2.15 A

and h
O
= 1.64 A

, respectively (Table 2).


Fig. 4b shows the structures of IS, TS and FS for reac-
tion (4) and Table 2 lists some key geometrical parameters
of them. The process of this reaction resembles that of (2).
Table 2
Key geometrical parameters of the initial state (IS), transition state (TS)
and nal state (FS) for reactions (1)(6)
Reaction (1) (3) (5)
IS TS FS IS TS FS IS TS FS
d
HdO
0.98 1.68 3.09 0.98 1.36 2.01 0.98 1.68 3.47
d
HdPd
2.66 1.81 1.79 2.66 2.42 2.55 2.82 1.66 1.77
d
OPd
2.45 2.07 2.15 2.39 2.11 2.15 2.57 2.09 2.00
h
Hd
2.47 1.04 0.80 2.48 1.95 2.06 2.56 1.22 0.79
h
O
2.45 2.01 1.63 2.39 2.04 1.64 2.54 2.06 2.00
d
OBHd
2.59 1.10 0.99 2.41 2.48 3.06
d
OBPd
2.02 2.03 2.15 2.20 2.10 2.10
h
OB
1.25 1.53 1.64 1.70 1.61 1.58
(2) (4) (6)
d
HdO
0.98 1.46 2.82 0.98 1.51 2.79 0.99 1.46 2.84
d
HdPd
2.66 1.71 1.77 2.65 1.61 1.78 2.60 1.60 1.76
d
OPd
2.14 2.06 1.99 2.15 2.02 2.00 2.20 2.01 2.02
h
Hd
2.17 1.46 0.79 2.17 1.44 0.79 2.15 0.76 0.76
h
O
1.61 1.35 1.18 1.62 1.27 1.19 1.70 1.30 1.23
d
OBHd
2.98 2.50 2.82 2.25 3.01 2.54
d
OBPd
1.99 1.96 1.98 2.57 2.53 2.52
h
OB
1.15 1.17 1.19 2.54 2.44 2.49
d
HdO
: bond distance of the OH bond to be cleavaged; d
HdPd
: the
shortest distance between the dissociated H atom (H
d
) and a Pd atom;
d
OPd
: the shortest bond distance between the O atom linked to the H
d
in
IS and a Pd atom; h
Hd
and h
O
: height of the H
d
atom and the O atom
attached to it in IS; d
OBHd
: the shortest distance between the O atom of
the co-adsorbate (B) and the H
d
. d
OBPd
: the shortest distance between the
O atom of B and a Pd atom. h
OB
: height of the O atom of B. E
a
: energy
barrier. All the bond distances are in angstrom and energies in eV.
Fig. 4. Schematic illustration (top view) of the OH bond breaking of H
2
O and OH at O covered surface. (ab) correspond to reaction (3) and (4),
respectively. Atom colouring are the same as in Fig. 3.
Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583 4579
To be concise and avoid repetition, we omit the description
of the geometries.
3.2.3. H
2
O
ad
+ OH
ad
!2OH
ad
+ H
ad
The initial state of this reaction is chosen to be the co-
adsorption structure of H
2
O and OH discussed in Section
3.1.3. This reaction begins with the displacement of the dis-
sociated H
d
atom towards an hcp site, which elongates the
H
d
O bond (Fig. 5a). In the TS, the H
d
O bond distance
increases from 0.98 A

to 1.68 A

(Table 2). The H


d
atom
is around a top site with the height of 1.22 A

, compared
to 2.56 A

in the IS, and the shortest H


d
Pd distance de-
creases from 2.82 A

in the IS to 1.66 A

, showing an en-
hanced interaction of H
d
with the substrate. The O atom
of H
2
O roughly occupies a top position with the OPd
bond length of 2.09 A

, 0.30 A

shorter than in the IS. The


co-adsorbed OH group remains at the bridge site in the
TS with a slight decrease of 0.09 A

in the height. After


the TS, the H
d
moves towards an hcp site. In the FS,
the H
d
resides on the hcp site with the height of the H
d
and the shortest H
d
Pd bond length being 0.79 A

and
1.77 A

, respectively. The OH
d
bond is completed broken
(d
HdO
= 3.47 A

). The product fragment, OH group, is lo-


cated on a top site, and the OPd bond length is calculated
to be 2.00 A

. The co-adsorbed OH group still stays on the


bridge site with the OPd bond length of 2.10 A

, 0.10 A

shorter than in the IS.


3.2.4. OH
ad
+ H
2
O
ad
!O
ad
+ H
ad
+ H
2
O
ad
The IS structure is the same as for reaction (5). Starting
from the IS, the O atom of the OH group slides to a nearby
fcc site with a tilting down of the OH
d
bond towards the
substrate surface, which weakens the OH
d
bonding and
facilitates the interaction of H
d
with substrate. In the TS,
the O atom of the OH resides at the fcc site. The shortest
OPd bond length is 2.01 A

while the OH
d
bond
(d
HdO
) reaches 1.46 A

(Table 2). The H


d
atom is basically
at an hcp site, 0.76 A

(2.15 A

in the IS) above the substrate


surface. The shortest H
d
Pd bond length is 1.60 A

. Because
the H
d
and the O atom of the OH group share the two Pd
atoms in the TS, a strong bonding competition [42] is ex-
pected, which drives the H
d
atom to a neighboring fcc site.
After the TS, the H
d
atom moves from the hcp site to the
nearby fcc position (Fig. 5b). In the FS, the OH
d
bond ex-
tends to 2.84 A

. The height of the H


d
atom remains to be
0.76 A

as in the TS. The produced oxygen atom occupies


the fcc site, with the height decreased from 1.30 A

in the
TS to 1.23 A

in the FS. The co-adsorbed H


2
O tends to shift
away from the previous top site to an fcc position before
the TS. After the TS it retreats back to the initial top site
(Fig. 5b).
The above analyses show that the TS structures tend to
keep the hydrogen bonding which is obviously helpful for
stabilizing the TS (Fig. 5).
3.3. Thermodynamics and kinetics of water decomposition
on Pd(111) surfaces
We have described the geometrical evolvement of water
dehydrogenation at various surfaces. Now we focus on the
thermodynamics and kinetics of the six elementary steps.
Table 3 contains the calculated reaction barriers, reaction
heats, pre-exponential factors and reaction rate constants
at 300 K for these steps. The classical corrected barriers
(E
cc
) are lower than the quantum corrected barriers E
qm
,
indicating that the zero-point vibrational energy is overcor-
rected in the classical scheme (Table 3). The dierence be-
tween the classical and quantum correction methods lies
between 0.01 eV and 0.05 eV. Tunneling eect increases
the rate constant. The higher the crossover temperature
T
c
, the larger is the transmission coecient. In all the cases
we studied here, the largest transmission coecient is
Fig. 5. Schematic illustration (top view) of the OH bond breaking of H
2
O in the presence of OH and OH bond scission of OH in the presence of H
2
O.
(ab) correspond to reaction (5) and (6), respectively. Atom colouring are the same as in Fig. 3.
4580 Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583
found with reaction (2), 2.7, whereas the smallest one is cal-
culated with reaction (3), 1.1. Although dierent correction
schemes produce dierent barriers or rate constants, they
predict the same trends. In the following, unless explicitly
stated, we adopt the kinetic parameters derived from E
qm
to discuss the kinetics of each reaction.
Fig. 6 illustrates the energy prole for the OH breaking
of H
2
O (reactions (1), (3) and (5), Fig. 6a) and of OH (reac-
tions (2), (4) and (6), Fig. 6b). Let us rst discuss the OH
bond scission of H
2
O on the clean surface (1), oxygen cov-
ered Pd(111) (3) and OH co-adsorbed Pd(111) surface (5).
As revealed in Fig. 6 and Table 3, the three reactions are
computed to be endothermic by 0.59 eV (1), 0.19 eV (3)
and 0.78 eV (5), respectively. On the clean surface, the acti-
vation energy for the OH bond scission is 0.92 eV (1);
when co-adsorbed with OH group, the barrier increases
to 1.23 eV (5), indicating that dehydrogenation of water be-
comes more dicult. This is mainly because the reactant is
stabilized by the hydrogen bonds formed between H
2
O and
OH (see Section 3.1.3). As shown in Fig. 6a the binding
energies of H
2
O in the IS of (1) and (5), are notably lower
than the activation energies. This implies that H
2
O will des-
orb from the surface, rather than decompose into OH. On
the other hand, the water adsorption energy of the reactant
in (3) is essentially the same in magnitude as the corre-
sponding barrier. In this case, the OH bond breaking will
strongly compete with water desorption. In fact, the activa-
tion energy of (3) is predicted to be very small, 0.34 eV.
Such a low barrier means that water decomposition is
kinetically very favorable. Indeed, the largest rate constant
(k
qm
) reaches 4.0 10
6
for (3), which is 10
10
and 10
15
larger
than those of (1) and (5), respectively. Why the activation
energy of (3) is so small? We have compared and analyzed
the adsorbate-substrate and adsorbateco-adsorbate inter-
action in the ISs and TSs of (1) and (3). It is found that the
reactant (H
2
O) of (3) is 0.20 eV more stable than that of (1)
(Table 1). In addition, mainly because of the longer dis-
tance of H
d
from the surface, 1.95 A

of h
Hd
in (3) compared
to 1.04 A

in (1) (see Table 2), the transition state complex


(HO H) of (3) is 0.7 eV less stable than the one of (1).
These two factors disfavor the barrier reduction of (3).
On the other hand, the transition state complex in (3) is sta-
bilized by 2.05 eV by the co-adsorbed oxygen O
B
at the ex-
pense of stability decrease of 0.5 eV of the oxygen. The net
reduction of the barrier is thus estimated to be 2.050.2
0.70.5 = 0.65 eV, close to the dierence of 0.58 eV (Table
3). This shows that the barrier reduction in (3) is owing to
the bonding interaction between O
B
and H
d
(d
OBHd
=
1.10 A

in the TS of (3)).
Our above results indicate that water decomposition on
clean Pd(111) is unlikely whereas in the presence of oxy-
gen, this process (H
2
O + O !2OH) is able to occur. These
conclusions are in accordance with the facts that dissocia-
tive adsorption of water on clean Pd(111) surface is not
observed [13], while hydroxyl groups are detected when
Table 3
Kinetic parameters and reaction heat for the reactions (1)(6)
E
a
E
cc
E
qm
DH A
0
k
cc
k
qm
k
qmt
T
c
(1) 1.09 0.87 0.92 0.59 5.9E11 1.6E3 2.2E4 3.0E04 142
(2) 1.30 1.13 1.18 0.20 7.5E12 8.0E7 9.9E8 2.7E07 304
(3) 0.43 0.33 0.34 0.19 1.8E12 4.5E6 4.0E6 4.4E06 70
(4) 1.48 1.33 1.37 0.57 3.3E12 1.3E10 2.9E11 6.8E11 270
(5) 1.37 1.20 1.23 0.78 9.0E11 7.3E9 2.1E9 2.8E09 136
(6) 1.57 1.42 1.46 0.43 2.5E12 3.3E12 6.7E13 1.4E12 242
E
a
: activation energy without zero-point energy correction; E
cc
(eV): classical zero point corrected barrier energy. E
qm
(eV): Wigner zero point corrected
barrier energy. DH: reaction heat (eV), positive values denote endothermic reactions. A
0
(s
1
): pre-exponential factor. k
cc
(s
1
): reaction rate constant at
300 K with classical zero point energy corrected barrier. k
qm
and k
qmt
(s
1
): reaction rate constant at 300 K from the quantum zero point energy corrected
barrier without and with tunneling correction. T
c
(K): crossover temperature for tunneling. aEb = a 10
b
.
-3.0
-2.5
-2.0
-1.5
-1.0
-0.5
0.0
0.5
1.0
1.5
FS
TS
IS
E
n
e
r
g
y

(
e
V
)
(2)
(4)
(6)
-0.8
-0.6
-0.4
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
FS
TS
IS
E
n
e
r
g
y

(
e
V
)
(1)
(3)
(5)
Fig. 6. Energy proles for the OH bond breaking in H
2
O (a) and OH (b)
in various systems. Energy zero point refers to H
2
O or OH and the
innitely separated clean slab (for (1) and (2)), or the slab covered with O
(for (3) and (4)), with OH (for (5)) or with H
2
O for (6). IS, TS and FS
denote initial states, transition states and nal states, respectively.
Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583 4581
water adsorbs on oxygen pre-covered Pd(100) [39]. Note
the calculated barrier of (3), 0.3 eV, is quite low, implying
that OH formation through this path is still feasible, even
on the most inactive Pd(111) surface. Similar nding has
been reported concerning O assisted dissociation of water
on the (111) of Pt (a group VIII metal with very similar
surface chemistry to Pd), where the reaction barrier is cal-
culated to be 0.33 eV [41]. Therefore, we speculate that the
produced OH groups from adsorbed water on oxygen cov-
ered Pd surface in Ref. [39] comes from the path: H
2
O
ad
+
O
ad
!2OH
ad
. Ref. [39] also reported that at 215 K the hy-
droxyl groups recombine to produce adsorbed oxygen and
a water desorption state on Pd(100), and they suggested
that water is formed via the reaction OH + OH !
H
2
O + O [39]. The low barrier we calculated, 0.2 eV, pre-
sents a strong support for the proposed mechanism. Fur-
thermore, our results may shed light on the question
raised in Ref. [10]: which of the following two reactions
is responsible for the water formation on Pd surface:
OH + H !H
2
O or OH + OH !H
2
O + O? These two
steps are the reverse processes of (1) and (3), denoted as
(1
0
) and (3
0
), respectively. From the calculated barriers
and reaction energies in Table 3, one gets the following
activation energy: 0.33 eV for (1
0
) and 0.15 eV for (3
0
).
Both are very small, especially that of (3
0
). Thus, we con-
clude that both (1
0
) and (3
0
) may play role in the water for-
mation process with a dominant contribution from (3
0
) at
lower temperatures. At higher temperatures, (1
0
) becomes
more important.
Finally we discuss the hydroxyl decomposition via (2),
(4) and (6) channels. All the forward reactions are also cal-
culated to be endothermic by at least 0.20 eV (2) (Fig. 3b).
At clean Pd(111) surface the barrier is computed to be
1.18 eV (2) (Table 3) while at the O covered surface, the
barrier increases a bit to 1.37 eV (4). The co-adsorbed
H
2
O makes the hydroxyl decomposition even harder with
a barrier of 1.46 eV (6). The largest rate constant (k
qm
) is
found with (2), 9.9 10
8
, which is about 10
3
to 10
5
higher
than (4) and (6), respectively. However, even the barrier of
(2) is too high for the reaction to take place noticeably at
room temperatures. It is worthy to mention that our calcu-
lated barrier for OH formation through the reverse reac-
tion of (2), O
ad
+ H
ad
!OH
ad
, is 0.98 eV, much higher
than the experimental value of 0.32 eV [43] or <0.4 eV
[10]. This indicates that formation of OH group are unli-
kely via (2
0
) on the ideal Pd(111) surface. STM experi-
ments [10] clearly show that the formation of OH occurs
at the step edge, which is more active than sites at the reg-
ular planar surfaces, thus a lower barrier is expected. We
note our calculated barrier of (2
0
) is higher than (1
0
) (Table
3), in line with the deduction based on the tting of the ki-
netic model to the experimental data [44]. It is notable that
formation of OH group from atomic O and H via the re-
verse process of (4), 2O
ad
+ H
ad
!O
ad
+ OH
ad
, needs to
overcome a barrier of 0.80 eV, which is 0.2 eV smaller
than the one of (2
0
), implying that co-adsorbed oxygen
has a notable inuence on the association of atomic O
and H, investigations of the mechanism of the OH forma-
tion should take the eect of co-existed oxygen into ac-
count. Relevant work is under way.
4. Conclusions
We have studied the adsorption and reaction of water
on Pd(111) surfaces using periodic slab models at density
functional PW91 level. Water adsorption was found to be
physi-sorption. It preferably binds to the top site through
the O atom with the molecular plane almost parallel to
the substrate surface. Hydroxyl group favors the bridge site
with the OH bond axis tilted relative to the surface normal
by 75. Water molecules were stabilized through hydrogen
bonding or the co-adsorption of O atom. Hydrogen
abstraction in H
2
O and OH was found dicult on the clean
Pd (111) surface. All the dehydrogenation steps are com-
puted to be endothermic. We found that co-adsorbed OH
increases the activation energy of the OH bond scission
of H
2
O by 0.3 eV, compared to that on the clean surface.
The co-adsorbed H
2
O and O also inhibit the hydroxyl
decomposition by promoting the energy barrier of
0.19 eV and 0.28 eV, respectively. In contrast, the energy
barrier for the OH bond breaking of H
2
O is greatly re-
duced from 0.92 eV (1) to 0.34 eV (3) in the presence of
O. The corresponding barrier of the reverse reaction of
(3) is also predicted to be very small, 0.2 eV. Such low
barriers show that production of OH group by dosing
water on oxygen covered Pd surface, or reversibly, forma-
tion of H
2
O from the combination of OH groups is feasi-
ble. These results are in agreement with the experimental
observations. Our calculations support the hypothesis [39]
that hydroxyl formation is via the reaction channel
O + H
2
O !2OH. Our results also indicate that catalytic
formation of water is likely dominated by 2OH !
O + H
2
O at low temperatures whereas reaction HO +
H !H
2
O becomes more important at higher tem-
peratures.
Acknowledgments
Financial support from a Foundation for the Author of
National Excellent Doctoral Dissertation of PR China
(No. 200123) and the natural science foundation of China
(NSFC2003CB615804) are greatly acknowledged. Part of
calculations is carried out on SGI3800 supercomputer of
the computing center of Nanjing University.
References
[1] P.A. Thiel, T.E. Madey, Surf. Sci. Rep. 7 (1987) 211.
[2] M.A. Henderson, Surf. Sci. Rep. 46 (2002) 1.
[3] V.V. Galvita, V.D. Belyaev, V.A. Semikolenov, P. Tsiakaras, A.
Frumin, V.A. Sobyanin, React. Kinet. Catal. Lett. 76 (2002) 343.
[4] V.V. Galvita, G.L. Semin, V.D. Belyaev, V.A. Semikolenov, P.
Tsiakaras, V.A. Sobyanin, Appl. Catal. A 220 (2001) 123.
[5] G. Karagiannakis, S. Zisekas, C. Kokkotis, M. Stoukides, Catal.
Lett. 104 (2005) 219.
4582 Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583
[6] G. Karagiannakis, S. Zisekas, C. Kokkotis, M. Stoukides, Appl.
Catal. A 301 (2006) 265.
[7] F. Von Looij, E.R. Stobbe, J.W. Gues, Catal. Lett. 50 (1998) 59.
[8] M.A. Pen a, J.P. Go mez, J.L.G. Fierro, Appl. Catal. A 144 (1996) 7.
[9] J. Xie, Q. Zhang, K.T. Chung, Appl. Catal. A 220 (2001) 215.
[10] T. Mitsui, M.K. Rose, E. Fomin, D.F. Ogletree, M. Salmeron, J.
Chem. Phys. 117 (2002) 5855.
[11] S. Anderson, C. Nyberg, C.G. Tengstal, Chem. Phys. Lett. 104 (1984)
305.
[12] T. Mitsui, M.K. Rose, E. Fomin, D.F. Ogletree, M. Salmeron,
Science 297 (2002) 1850.
[13] M. Wolf, S. Nettesheim, J.M. White, E. Hasselbrink, G. Ertl, J.
Chem. Phys. 92 (1990) 1509.
[14] M. Wolf, S. Nettesheim, J.M. White, E. Hasselbrink, G. Ertl, J.
Chem. Phys. 94 (1991) 4609.
[15] X.Y. Zhu, J.M. White, M. Wolf, E. Hasselbrink, G. Ertl, J. Phys.
Chem. 95 (1991) 8393.
[16] E.M. Stuve, S.W. Jorgensen, R.J. Madix, Surf. Sci. 146 (1984) 179.
[17] J.M. Heras, G. Estiu, L. Viscido, Appl. Surf. Sci. 108 (1997) 455.
[18] A. Michaelides, V.A. Ranea, P.L. de Andres, D.A. King, Phys. Rev.
Lett. 90 (2003) 216102.
[19] A. Michaelides, A. Alavi, D.A. King, Phys. Rev. B 69 (2004) 113404.
[20] S. Meng, E.G. Wang, S. Gao, Phys. Rev. B 69 (2004) 195404.
[21] G. Kresse, J. Furthmu ller, Phys. Rev. B 54 (1996) 11169.
[22] G. Kresse, J. Furthmu ller, Comput. Mater. Sci. 6 (1999) 15.
[23] G. Kresse, J. Hafner, Phys. Rev. B 47 (1993) 558.
[24] J.P. Perdew, J.A. Chevary, S.H. Vosko, et al., Phys. Rev. B 46 (1992)
6671.
[25] P.E. Blo chl, Phys. Rev. B: Condens. Matter Mater. Phys. 50 (1994)
17953.
[26] H.J. Monkhorst, J.D. Pack, Phys. Rev. B 13 (1976) 5188.
[27] M. Methfessel, A.T. Paxton, Phys. Rev. B 40 (1989) 3616.
[28] G. Henkelman, B.P. Uberuaga, H. Jo nsson, J. Chem. Phys. 113
(2000) 9901.
[29] W.F.K. Wynne-Jones, H. Eyring, J. Chem. Phys. 3 (1935) 492.
[30] Z.-X. Chen, H.K. Lim, K.M. Neyman, N. Ro sch, J. Phys. Chem. B
109 (2005) 4568.
[31] G. Henkelman, A. Arnaldsson, H. Jo nsson, J. Chem. Phys. 124 (2006)
044706.
[32] E. Wigner, Trans. Faraday Soc. 34 (1938) 29.
[33] E. Wigner, Z. Phys. Chem. 19 (1932) 203.
[34] V.A. Ranea, A. Michaelides, R. Ramirez, P.L. de Andres, J.A.
Verges, D.A. King, Phys. Rev. Lett. 92 (2004) 136104.
[35] P.A. Redhead, Vacuum 12 (1962) 203.
[36] J. Cerda, A. Michaelides, M.-L. Bocquet, P.J. Feibelman, T. Mitsui,
M. Rose, E. Fomin, M. Salmeron, Phys. Rev. Lett. 63 (2004) 116101.
[37] A. Steltenpohl, N. Memmel, Surf. Sci. 443 (1999) 13.
[38] M.K. Rose, A. Borg, J.C. Dunphy, T. Mitsui, D.F. Ogletree, M.
Salmeron, Surf. Sci. 561 (2004) 69.
[39] C. Nyberg, C.G. Tengstal, J. Chem. Phys. 80 (1984) 3463.
[40] J. Kua, W.A. Goddard III, J. Am. Chem. Soc. 121 (1999) 10928.
[41] A. Michaelides, P. Hu, J. Am. Chem. Soc. 123 (2001) 4235.
[42] K. Bleakley, P. Hu, J. Am. Chem. Soc. 121 (1999) 7466.
[43] T.H. Engle, H. Kuipers, Surf. Sci. 90 (1979) 181.
[44] J. Fogelberg, L.G. Petersson, Surf. Sci. 350 (1996) 91.
Y. Cao, Z.-X. Chen / Surface Science 600 (2006) 45724583 4583

Das könnte Ihnen auch gefallen