You are on page 1of 14

Corrosion Science 73 (2013) 18–31

Contents lists available at SciVerse ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Early stage oxidation of Ni–Cr binary alloy (1 1 1), (1 1 0) and (1 0 0)


surfaces: A combined density functional and quantum chemical
molecular dynamics study
Nishith Kumar Das a,⇑, Tetsuo Shoji b
a
Nuclear Safety Research Center, Japan Atomic Energy Agency, 2-4 Shirakata-shirane, Tokai-mura, Naka-gun, Ibaraki-ken 319-1195, Japan
b
Frontier Research Initiative, New Industry Creation Hatchery Center, Tohoku University, 6-6-10, Aoba, Aramaki, Aoba-ku, Sendai 980-8579, Japan

a r t i c l e i n f o a b s t r a c t

Article history: First-principles and tight-binding quantum chemical molecular dynamics were used in this study. The
Received 5 November 2012 chemisorption energies of O and OH on the Ni–Cr (1 1 0) surface are lower than those of other surfaces.
Accepted 12 March 2013 The oxygen 2p orbitals hybridise with Ni 3d, 4s and small amounts of p orbitals for the (1 0 0) surface
Available online 3 April 2013
while Ni p orbitals have no contribution for the (1 1 0) surface, which might reduce the adsorption energy.
Additionally, oxygen acquires the maximum depth into the Ni–Cr (1 1 0) surface. Applied strain increases
Keywords: the oxygen diffusivity. This study reveals that the Ni–Cr (1 1 0) surface is easier for oxygen diffusion
A. Alloy
accordingly oxidation accelerates.
B. Modelling studies
C. Oxidation
Ó 2013 Elsevier Ltd. All rights reserved.
C. Interfaces
C. Stress corrosion

1. Introduction Nickel-based alloys are widely used in the energy industry,


especially for structural components, because they have excellent
The interaction of water with solid surfaces is of great impor- corrosion resistance and good mechanical properties. Ni-based al-
tance in chemistry, materials science, electrochemistry, corrosion loys contain several alloying elements, which in turn may influence
and biophysics. The dissociation process of water on metal pro- the oxidation behaviour of the alloy, but it might be difficult to sim-
vides the basis of an understanding of these fundamental reactions. ulate the oxidation behaviour of the alloy using computational
Therefore, water adsorption and dissociation has been studied methods. In this alloy, chromium is essential to assure oxidation
extensively on metal surfaces using various techniques [1–4]. resistance. Therefore, the binary alloy or ternary alloy might be
The interaction of water with metal in high temperature environ- the better option to understand the initial stage of the oxidation
ments leads to the dissolution of metallic atoms and the diffusion mechanism. Several studies have been conducted on Ni-based al-
of elements. Subsequently, oxidation takes place at the surface. loys to understand and quantify their corrosion resistance [7–10].
Oxidation is an indispensable process of alloys that can occur prior For example, Jeng et al. performed an Auger electron spectroscopy
to crack initiation. In relation to the role of the oxide film in stress (AES) and ion-scattering spectroscopy (ISS) study of single and
corrosion cracking (SCC), Engell observed that almost all metal– polycrystalline Ni–Cr alloy surfaces at 500 °C [11]. They found that
electrolyte systems with SCC form films [5]. SCC is composed Cr migrated to the near-surface region, where it reacted with NiO to
two parts: initiation and propagation. A crack may propagate into form Cr oxide. In addition, a X-ray photoelectron spectroscopy
a material when it is initiated. Thus, metal surface oxidation is an (XPS) study of the polycrystalline Ni–Cr alloy showed that Cr is oxi-
important step for the early stage of crack initiation. The process dised preferentially to Cr2O3 initially but that Ni in the alloy is oxi-
requires a supply of oxygen and metal atoms with sufficiently high dised more rapidly to NiO than Ni metal [12]. Moreover, numerous
mobilities on the surface [6]. The oxygen–metal interaction is an studies have been conducted on H2O/Ni (1 1 1), Ni (1 1 0) and Ni
essential process for surface oxidation initiation, and it is funda- (1 0 0) and O/Ni (1 1 1), Ni (1 1 0) and Ni (1 0 0) systems by first prin-
mentally very important. ciples, molecular dynamics and experimental methods [13–21].
These studies show that the water molecule forms a weak bond
with the metal surfaces while atomic oxygen makes a strong bond.
An electron energy loss spectroscopy (EELS) study of H2O/Ni (1 0 0)
⇑ Corresponding author. Tel.: +81 29 282 6473. reports water dissociation at 200 K. In addition, a thermal desorp-
E-mail address: ndasbuet@yahoo.com (N.K. Das). tion spectroscopy study showed that water molecules are adsorbed

0010-938X/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.corsci.2013.03.020
N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31 19

molecularly on the Ni (1 1 1) surface from 165 K at low coverage to numerical functions (DNP), was employed; it is comparable in
170 K at saturation, and the surface can absorb up to 260 K when quality to a Gaussian 6-31G basis set [32]. A 2  2 unit cell was
oxygen-predosed on the surface [19]. Several theoretical and exper- employed. The (1 0 0) and (1 1 0) metal surfaces were modelled by
imental studies have been conducted to determine the adsorption a periodical array of four layer slabs and separated by a 12 Å vac-
energy, the geometrical features, the electronic structures and the uum region, where the top three layers allowed free interaction
bonding mechanism. Carrasco et al. demonstrated the two orbital and the bottom layers were fixed. The atomic oxygen, OH and
overlap between water and metal surfaces [22]. The metal s state water molecule interactions were not expected beyond the first
is found to hybridise with the d state, resulting in hybridised s–d three layers. We calculated a five layer slab model. The calculated
orbitals projecting on water 3a1 and 1b1 orbitals. However, most adsorption energy difference was 9 and 7 meV for oxygen adsorp-
studies focused on the interaction between a clean Ni surface or tion on the Ni (1 0 0) and Ni (1 1 0) surfaces, respectively. Thus, the
oxygen pre-adsorbed Ni surface with other molecules or atoms. It choice of four layers is believed to be quite adequate. Each layer
is well known that the dilute impurities alter the geometry and consists of four nickel atoms. Brillouin zone sampling was per-
electronic structures of bulk system. One theoretical study found formed using a Monkhorst–Pack grid. We performed several k-
that the chemisorption energy of carbon is altered by substitution- points calculations, which showed that a 5  5  1 k-points mesh
ally embedded Cu atoms on the Ni (1 1 1) surface [23]. The Cu sub- for surfaces was sufficient. The calculated energy differences be-
stitute nickel surface also increased the activation energy barrier for tween 5  5  1 and 6  6  1 and 7  7  1 k-points meshes were
CH. The substitution of chromium into the nickel surface increased less than 10 meV. The selected set of k-points calculations is suffi-
the oxygen and water chemisorption energies, which might reduce cient for the present study. A previous study is also in good agree-
the oxygen mobility on the surfaces [24]. Moreover, a multiscale ment with this calculation [24]. The self-consistent field
modelling study found that the stress intensity significantly in- convergence tolerance was permitted up to 1.0  105 and a max-
creases the oxygen diffusivity on the Ni and Ni–Cr surfaces; subse- imum force of 0.04 Ha/Å. The double optimisation of the unit cell
quently, it can intensify the primary stage of oxidation initiation was performed. First, the structure and atomic positions were fully
[25]. Therefore, surface oxidation is a process of that is a precursor optimised, and then the optimised structure opted for further cal-
to the early stages of cracking initiation. In order to address this is- culations. For the bimetallic Ni–Cr surfaces, one Ni atom was re-
sue, it is desirable to get better description and deeper insight into placed by a Cr atom in the top layer, and two nickel atoms in the
the basic features of chemical reaction of water with metal surface. top and second layers were replaced by Cr atoms. Fig. 1 was con-
Moreover, the oxygen diffusivity and the nature of oxygen bonding sidered for the geometrical optimisation of the surfaces. Initially,
with the elements of the metal surface are important in studying water molecules were situated parallel to the surface, and OH mol-
the corrosion and passivation phenomena. Computational chemis- ecules were placed parallel and perpendicular to the surface. The
try is a very powerful and reliable tool in obtaining detailed infor- OH perpendicular to the surface was more preferred than OH par-
mation on atomic scale for understanding the oxidation initiation allel to the surface. Thus, the OH parallel to the surface was not
process. The present theoretical study has been undertaken to shed considered for further study. The oxygen atoms, OH and water
light in the interaction between oxygen, OH and H2O and the Cr- molecules were placed on top, bridge, or hollow sites.
doped Ni (referred to as Ni–Cr here) surfaces as a first step for The adsorption energies for atomic oxygen, hydroxyl and water
understanding the early stage of oxidation mechanism. At present, molecules were calculated with the following formula:
it is a demanding task to simulate by a computational approach. In
this paper, the density functional theory (DFT) method is consid- Ead ¼ EðNi=Ni—Cr surfacesÞ þ EðO=OH=H2 OÞ
ered for nickel and Cr-doped nickel surfaces with oxygen, OH and  EðNi=Ni—Cr surfaces þ O=OH=H2 OÞ
H2O interactions, and a tight-binding quantum chemical molecular
dynamics (QCMD) method is employed to understand the early where E(Ni/Ni–Cr) is the total energy of a similar supercell repre-
stage oxidation of the Ni–Cr (1 1 1), (1 1 0) and (1 0 0) surfaces. senting the Ni or Ni–Cr clean surfaces; E(Ni/Ni–Cr + O/OH/H2O) is
the calculated total energy of the O/Ni, OH/Ni, H2O/Ni, O/Ni–Cr,
OH/Ni–Cr, or H2O/Ni–Cr systems; and E(O/OH/H2O) is the energy
2. Computational details of atomic oxygen or OH or H2O molecules in a cube with dimen-
sions 12  12  12 Å3.
The DFT and tight-binding QCMD methods were used in this For tight-binding QCMD, we designed a set of parameters to ana-
study. DFT is an accurate method to calculate the geometry, the lyse the reaction dynamics of binary or ternary nickel-based alloys or
structure, the energetics and the orbitals interaction. These proper- austenitic stainless steels. Diatomic calculation performed followed
ties are very important to understand the fundamental phenome- by the bulk systems by considering several parameters such as bind-
non of alloy surfaces and different interfaces. The DFT method ing energy, equilibrium bond distance, cohesive energy and lattice
considers a small atomic model, but it can give us fruitful informa- constant. The set of parameters were quite consistent with the
tion from the fundamental point of view. For instance, the higher experimental and first principles calculated data; for example, the
chemisorption of atomic oxygen on a metal surface may lead to calculated cohesive energy of nickel is 4.42 eV, and the experimental
less movement on the surface. In contrast, the tight-binding QCMD value is 4.44 eV. More details about the parameterisation can be seen
method is a tool that can consider reaction dynamics and large elsewhere [33]. The tight-binding QCMD code ‘colors’ developed by
models at high temperature. These two methods were used to ana- Tohoku University was employed for the water dissociation and
lyse the oxidation initiation phenomenon at the atomic level. adsorption on the Ni–Cr (1 1 1), Ni–Cr (1 1 0) and Ni–Cr (1 0 0) binary
The DFT study intends to investigate the adsorption energy and alloy surfaces in high temperature environments. This code is five
detailed bonding nature of clean and oxygen, OH and water ad- thousand times faster than conventional DFT and can handle the
sorbed Ni (1 1 1) and Ni–Cr (1 1 1) surfaces; all calculations were reaction dynamics of metal surfaces efficiently. Ni–Cr surface crystal
performed using DFT [26] within the generalised gradient approx- structure was created by using the software Material Studio. First,
imation (GGA), as implemented in the DMol3 module of the Mate- DFT calculations were performed for the optimisation of a small
rials Studio packages [27–29]. The exchange–correlation energy Ni–Cr surface; then, we built the slab surface based on the optimised
was described using the spin polarisation form of the Perdew surface. Fig. 2 shows the models of the Ni–Cr binary alloy slab sur-
and Wang 91 functional (PW91) [30,31]. A numerical basis set, faces. The models consist of 84 atoms (71 of Ni and 13 of Cr atoms)
which describes the orbitals in the valence shell with double with a monolayer of four water molecules embedded on the sur-
20 N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31

Fig. 1. (a) Side view (different colours indicate the layer), (b) top view of the possible adsorption sites of Ni (1 1 0), (c) one chromium atom doped, (d) two chromium atom
doped Ni (1 1 0), (e) side view (different colours indicate the layer), (f) top view of the possible adsorption sites of Ni (1 0 0) and (g) one chromium atom doped and (h) two
chromium atom doped Ni (1 0 0) surfaces for DFT calculations. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

faces. Thirteen nickel atoms, at chosen random, were substituted by velocities need to be carried at any one time. The ensemble for this
chromium atoms for all slab models. Water molecules were placed simulation was constant volume and constant temperature (NVT),
3 Å away from the surfaces. The metal surfaces were modelled by a and the temperature was controlled by using scaling the atom veloc-
periodical array of seven layer slabs where the top layers allowed ities. The temperature considered was 325 °C, which represents the
free interaction and the bottom layer was fixed. One percent (1%) PWR condition, and atmospheric pressure was used.
uniaxial tensile strain was applied along the x-axis on the z-plane
of the slab model. The bulk length along the x-axis was extended 3. Results and discussion
1% over the initial length. The applied strain for the perfect crystal
surface should be elastic. A total of 8000 simulation steps were car- 3.1. DFT results
ried out, where each step was 0.2 femto second (fs), with the Verlet
algorithm. The advantage of this algorithm is that it requires less Previously, we studied the oxygen and water adsorption on Ni
computer memory because only one set of positions, forces and (1 1 1) and Ni–Cr (1 1 1) surfaces [24]. Here a combined DFT and
N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31 21

Terrace site Water

Ni

Cr

(a) (b) (c)


Fig. 2. Slab model surfaces for simulation: (a) Ni–Cr (1 1 1), (b) Ni–Cr (1 1 0) and (c) Ni–Cr (1 0 0).

tight-binding QCMD methods have been considered. The DFT atoms have slightly (2%) outward movement between the first
method applied to study atomic oxygen, OH, and water adsorption and second layer.
on Ni and Ni–Cr (1 1 0) and (1 0 0) surfaces and OH adsorption on Ni
(1 1 1) and Ni–Cr (1 1 1) surfaces. 3.1.2. Energetics and geometries of oxygen, OH and H2O adsorbed on
Ni and Cr-doped Ni (1 1 0) surfaces
3.1.1. Ni (1 1 1) and Ni–Cr (1 1 1) surfaces We investigated the site preference of O, OH and H2O on Ni
The calculated results found that the highest absorption energy (1 1 0) and Ni–Cr (1 1 0) surfaces. The calculated adsorption ener-
for oxygen was 6.86 eV for the Ni–Cr (1 1 1) surface, followed by gies are shown in Tables 1–3. Oxygen was found to adsorb on short
the hcp site, whereas the absorption energy was 5.56 eV for the bridge (SB) site of the unreconstructed p(2  2) Ni (1 1 0) surface
Ni (1 1 1) surface [24]. The result concerning the Ni–Cr surface im- with an adsorption energy of 5.22 eV. Long bridge (LB) site was
plied that the bond distances were 1.93–1.95 Å and 1.75 Å for Ni–O the next most stable site with an adsorption energy of 5.10 eV.
and Cr–O, respectively. The oxygen atoms preferred to bond with Oxygen chemisorption on Ni (1 1 0) surface has been extensively
Cr rather than Ni atoms. However, we calculated OH adsorption studied by various methods. One theoretical study concluded that
properties on the Ni (1 1 1) and Ni–Cr (1 1 1) surfaces. The OH the short bridge site is the most preferential site for oxygen
adsorption energies were 3.28 eV and 3.73 eV for the Ni and Ni– adsorption. Moreover, a low-energy electron diffraction (LEED)
Cr surfaces, respectively. The most stable adsorption sites were study demonstrated that the adsorbed oxygen atoms were located
fcc hollow and on top Cr site for the Ni and Ni–Cr surfaces, respec- in the short-bridge sites of the unreconstructed Ni (1 1 0) surfaces
tively. The OH adsorption energy was significantly increased due to [36]. The Ni–O bond distance was found to be 1.77 Å, which is
the substituted Cr on top of the nickel surface. The previous theo- the same as the value found in the previous study [37]. The oxygen
retical study found an OH adsorption energy on the Ni (1 1 1) sur- adsorption energies were increased to 6.55 and 6.60 eV for one and
face of 3.71 eV [34]. They considered the cluster model of the Ni two chromium atom doped Ni (1 1 0) surfaces, respectively. Fur-
(1 1 1) surface, which might give us a different energy value. The thermore, the adsorption energy was significantly decreased for
bond distance of the nickel and the oxygen of the OH was 2.01 Å the surface with a chromium atom doped into the second and third
for the pure nickel surface, which is very close to the value found layers. For instance, the first and third layer chromium doped sur-
in the previous study [34,35]. In the case of the Ni–Cr surface, face is 1.2 eV less stable than the one chromium atom doped Ni
the chromium and oxygen bond distance was 1.79 Å. The oxygen (1 1 0) surface. A previous DFT study found that the oxygen adsorp-
of OH preferentially bonds with chromium. The adsorbed OH was tion energy increased for the Ni–Cr (1 1 1) surface, which agrees
approximately 5° tilted relative to the surface normal, and the O–H very well with the present study [24]. Chromium substituted on
bond length was 0.97 Å. The results are in agreement with the the top layer had a considerable effect on the oxygen adsorption
observations of Yang et al. [35]. The OH adsorbed surface metallic energy while the effect is insignificant for the second and third lay-

Table 1
Oxygen adsorption on Ni (1 1 0) and Cr-doped Ni (1 1 0) surfaces.

Surface Sites Ead (eV) Ni–O distance Cr–O distance


Ni (1 1 0) Hollow 5.11 2.10
Long-bridge 5.05, 5.29a 1.93
Short-bridge 5.22, 5.35a 1.77
Top 4.14 1.66
Ni-top1Cr (1 1 0) Hollow 6.05 1.92 1.68
Long-bridge 6.05 1.96 1.69
Short-bridge 6.55 1.91 1.68
Top 6.38 1.59
Ni-top2Cr (1 1 0) Hollow 6.10 2.3 1.89
Long-bridge 6.01 1.94 1.70
Short-bridge 6.60 1.92 1.68
Top 6.33 1.59
a
Ref. [35].
22 N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31

Table 2
OH adsorption on Ni (1 1 0) and Cr-doped Ni (1 1 0) surfaces.

Surface Sites Ead (eV) Ni–O distance Cr–O distance


Ni (1 1 0) Hollow 2.87 2.28
Long-bridge 3.36 1.99
SB (inclined) 3.66 1.93
SB (perp.) 3.39 1.90
Ni-top1Cr (1 1 0) Hollow 3.46 2.25 2.09
Long-bridge 3.68 2.13 1.91
SB (inclined) 4.30 2.01 1.89
SB (perp.) 4.08 1.97 1.88
Ni-top2Cr (1 1 0) Hollow 3.50 2.38 2.16
Long-bridge 3.72 2.08 1.96
SB (inclined) 4.26 2.01 1.88
SB (perp.) 4.03 1.99 1.87

tance was 1.87 Å while Ni–O bond distance was 1.99 Å, which re-
Table 3
H2O adsorption on Ni (1 1 0) and Cr-doped Ni (1 1 0) surfaces.
veals that the Ni–O bond distances were slightly increased due to
the chromium doped into the surfaces. The Cr–O bond distances
Surface Sites Ead (eV) Ni–O distance Cr–O distance were smaller than those of Ni–O, corresponding to the preferential
Ni (1 1 0) Hollow 0.48 binding between the oxygen of the OH molecule and the Cr of the
Long-bridge 0.37 Cr-doped alloy surface.
Short-bridge 0.48
For H2O/Ni (1 1 0), the calculated absorption energy was 0.70 eV,
Top 0.70 2.14
which is the highest among all low-index nickel surfaces. The top
Ni-top1Cr (1 1 0) 0.99 2.135
site was the most stable site. The value is slightly higher than
Ni-top2Cr (1 1 0) 0.98 2.133 the previous results (0.58 eV) [44]. However, surface roughness in-
creased the adsorption energy. Experimental observation demon-
ers. It is evident that chromium atoms on the top layer are very ac- strated that the more open (1 1 0) surface adsorbs H2O more
tive for oxygen adsorption. The Cr–O and Ni–O bond distances strongly [41,45]. Our findings are in agreement with the previous
were 1.68 and 1.92 Å for Ni–Cr (1 1 0) surfaces, respectively. A clus- theoretical and experimental results. The calculated O–H bond dis-
ter model study found a bond distance of 1.61 Å for Cr–O in Cr2O3 tance and HOH angle were 0.98 Å and 103.5°, respectively. Water
[38]. The X-ray adsorption near-edge structure (XANES) structure was molecularly adsorbed via the oxygen atom, and the orienta-
showed a bond distance of 1.60 Å for Cr–O in Cr(IV) [39]. The short- tion of the hydrogen atoms was parallel to the surface. Several
est distance between chromium and oxygen corresponds to strong experimental studies observed that H2O is bound through the oxy-
binding. Oxygen bonds favourably with chromium, which might gen and that the hydrogen atom is parallel to the Ni (1 1 0) surface
increase the adsorption energy. [42]. The Ni–O bond distance was 2.14 Å. The X-ray absorption
Hydroxyl species are often formed on metal surfaces. The LEED fine-structure technique (SEXAFS) analysis showed that the water
analysis showed the OH formation on an oxygen preadsorbed Ni molecules occupied top sites with Ni–O bond lengths of
(1 1 0) surface below 360 K [40]. A angle resolved ultraviolet photo- 2.06 ± 0.03 Å [46]. In addition, a previous DFT calculation (2.20 Å)
emission spectroscopy (UPS) and ESDIAD study also found OH on and surface X-ray diffraction observation (2.24 Å) found identical
the Ni (1 1 0) surface at 350 K [41]. The most preferential site was values [24,47,48]. With the doping of chromium into the surfaces,
SB of OH inclined to the normal of the surface, and the oxygen the adsorption energy increased to 0.98 eV for the one and two
was always at the bottom. The adsorption energy of the most sta- chromium doped surfaces. The water adsorption energy was not
ble structure was 3.66 eV. The adsorption energy of OH on SB site increased due to the doping of two chromium atoms. The tendency
with perpendicular to the surface was 3.39 eV, which is smaller is very similar to that of OH adsorption. The O–H bond distance
than that of the most stable SB site. One experimental study dem- and the HOH angle of the adsorbed water molecule were 0.98 Å
onstrated that OH molecules were inclined with respect to the sur- and 103°, respectively, which is close to those of the gas phase
face normal [42]. The calculated adsorption energy is marginally water molecule. The calculated Cr–O bond distance of 2.14 Å is
higher than that of the previous ab initio study [43]. The small size compatible with the previous study [24].
of the cluster and the different method may affect the value. The The interlayer spacing was significantly modified for the bare Ni
DFT study of OH adsorption on Ni28 cluster gave 3.77 eV for the (1 1 0) surface. The present study demonstrated a 9% contraction in
Ni (1 1 1) surface [34]. The O–H bond distance was found to be the first interlayer and a 4% expansion in the second interlayer dis-
0.98–0.99 Å, which is very close to the calculated gas phase value. tance. The dynamical LEED study confirmed that the first interlayer
It indicates that OH is molecularly adsorbed on the surface and that spacing is contracted by 9.8 ± 1.8% and that the second interlayer
the adsorption is weaker than that of atomic oxygen, with a Ni–O spacing is increased by 3.8 ± 1.8% [49]. Another LEED study demon-
bond distance 1.93 Å. The absence of reconstruction of the Ni (1 1 0) strated an 8.7 ± 0.5% contraction and a 3 ± 0.6% expansion in the
surface upon OH adsorption is attributed to the weaker bonding first and second interlayer spacing, respectively [50]. Several
interaction between the O and Ni atoms in the case of OH chemi- experimental and theoretical studies reported the contraction of
sorption compared to O chemisorption [40]. The adsorption energy the first interlayer spacing and the expansion of the second inter-
of the Ni–Cr surface was increased to 4.30 eV. The most preferable layer spacing [51]. The Cr-doped surfaces showed 11.5% inward
site was SB with inclined to the surface normal. It appears that the and 5.25% outward movement of the first and second interlayer
doping of chromium did not change the adsorption sites. The OH spacing, respectively. The doping of chromium slightly increased
adsorption energy on the two chromium atom doped surface did the surface structure modification compared to that of the pure
not increase compared to that of the single chromium doped sur- nickel surface. The adsorption of oxygen on the Ni (1 1 0) surface
face. The higher chromium concentration on the Ni (1 1 0) surface led to expansion (4%) between the first and second layers and
had less effect on the OH adsorption energy. The Cr–O bond dis- contraction (1.11%) between the second and third layers. Gunn
N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31 23

et al. previously noticed a 2.66% expansion between the first and an adsorption energy of 3.37 eV. The OH was adsorbed perpendic-
second layers and a 0.24% expansion between the second and third ular to the surface with a Ni–O bond distance of 2.14 Å. The calcu-
layers [52]. Adsorption sites may be attributed to this difference. lated distance between the oxygen and the nearest nickel was
The OH adsorbed surface had a 2.5% expansion and a 2% contrac- found to be 2.16 Å by Yang et al. [66]. They found that the adsorp-
tion between first and second and second and third layers, respec- tion energy is slightly higher than that in the present study. Differ-
tively, while the H2O adsorbed surface had no significant ent methods might provide different results while the adsorption
expansion between first and second inter layer. The adsorption of site and OH adsorption axis to the surface is consistent. The Ni–
oxygen on the Cr-doped surface led to a 2.2% expansion between Cr bridge site was found to be the most stable adsorption site for
the first and second layers. There is a negligible contraction be- one Cr-doped nickel surface, and the hollow site was 0.03 eV less
tween the second and third interlayer spacing. Additionally, the stable. The hollow site adsorption energy is very competitive.
OH and H2O adsorbed surfaces showed a very small (1.5%) expan- The OH adsorption energy on the hollow site was 4.15 eV for the
sion between the first and second layers. two Cr-doped Ni–Cr (1 0 0) surface. The energy of the Ni–Cr bridge
site was 0.20 eV smaller than that of the hollow site. It seems that
hollow site is the most stable for OH and oxygen adsorption. The
3.1.3. Energetics and geometries of oxygen, OH and H2O adsorbed on
Cr–O bond distances were smaller than the Ni–O bond distance
Ni and Cr-doped Ni (1 0 0) surfaces
for one and two chromium doped Ni–Cr (1 0 0) surfaces. The OH
Tables 4–6 show the calculated adsorption energies. The
distances varied from 0.97 to 0.98 Å. Water molecules were ad-
adsorption energy was calculated with the spin polarisation in
sorbed on the top site of the Ni (1 0 0) surface and remained parallel
the 2  2 unit cell with a coverage of 0.25 ML. The oxygen adsorp-
to the surface with an adsorption energy of 0.56 eV. The O–H bond
tion energy on the Ni (1 0 0) surface was 5.69 eV, and the most
distance was 0.98 Å, and the HOH angle was 103.46°. It reveals that
favourable adsorption site was the fourfold hollow site. One exper-
the water was absorbed molecularly on the bimetallic alloy sur-
imental study measured a value of 5.64 eV, which is in good agree-
faces. The previous DFT calculated adsorption energy of water on
ment with the present study [55]. A previous DFT study found a
the Ni (1 0 0) surface was 0.51 eV, which is compatible with this
wide range of adsorption energies from 4.81 eV to 6.7 eV [56]. Nev-
study [44]. A few theoretical and experimental studies on the Ni
ertheless, the adsorption energy of this study is in agreement with
(1 1 1) surface showed that the water absorption energy was
the experimental results. However, the four fold hollow site is the
0.52 eV and 0.43 eV, respectively [24,67]. The adsorption energy
most stable site for oxygen adsorption on the Ni (1 0 0) surface [57].
increased to 0.90 eV for one chromium doped on the top layer,
The Cr-doped surfaces increased the adsorption energies. The max-
but the adsorption energy decreased to 0.67 eV for the two chro-
imum adsorption energy was attained by the surface with two
mium atom doped alloy surface. The geometrical structure showed
chromium atoms doped on the Ni (1 0 0) surface with 6.82 eV.
that the Cr–O bond distances were 2.14 and 2.17 Å for one and two
The fourfold hollow site was the most favourable while the top site
chromium atom doped surfaces, respectively, and the hydrogen
was the most unfavourable. The bridge site was the next most sta-
atoms of the water molecules remained on the hollow site while
ble site. The oxygen adsorption energy of the Ni–CrNi–Cr (1 0 0)
one hydrogen atom was on the bridge and another one was on
surfaces were higher than that of the pure nickel surfaces, as
the hollow site. These two features may affect the adsorption en-
shown in Fig. 3, which reveals that oxygen bonds strongly with
ergy. The molecularly adsorbed water made a weak bond with me-
Ni–Cr surfaces. The oxygen adsorbed surfaces did not show recon-
tal surfaces.
struction. This observation is in agreement with the previous stud-
The clean Ni (1 0 0) surface showed an approximately 1% inward
ies [58–60]. The oxygen adsorption energy on the (1 0 0) surface is
relaxation between the first and second layers, and the value is
the highest among the low-index nickel surfaces. A higher adsorp-
perfectly consistent with the experimental observation [59]. The
tion energy indicates the lower oxygen mobility on the surface. As
clean surface is well-known to be unreconstructed. Top layer Ni
a result, the process might decelerate the oxygen movement on the
atoms moved slightly outward (0.08 Å) to adsorb the oxygen atom.
Ni (1 0 0) surface. The calculated Ni–O bond distance was 1.96 Å.
The LEED analysis also found a 0.07 Å elongation for the p(2  2)
The value is exactly same as that according to the surface extended
structures [68]. The top layer nickel atoms had a small (>1%) out-
X-ray absorption fine structure (SEXAFS), LEED and normal photo-
ward relaxation for the OH and H2O absorption. The Ni–Cr surface
electron diffraction (NPD) studies and in fair agreement with the
showed a slightly higher (2%) elongation between first interlayer
high-resolution electron energy loss spectroscopy (HREELS,
compared to the nickel surface due to oxygen adsorption. The Cr–O
1.98 Å) studies [61–65]. The Ni–O bond distances were increased
preferential bond might help to increase the metallic bond. The OH
to 2.05 Å in the presence of Cr on the Ni–Cr surface while the
and H2O adsorption had a small effect (1%) on the Ni–Cr bond dis-
Cr–O bond distances varied from 1.80 to 1.82 Å. The shorter Cr–O
tance. Molecular adsorption had less effect on the interlayer
bond distance implies that the oxygen has some affinity to chro-
spacing.
mium. For OH, the fourfold hollow site was the most stable with

3.1.4. Density of states analysis


Table 4 Mulliken and Hirshfeld charge analysis suggested that oxygen
Oxygen adsorption on Ni (1 0 0) and Cr-doped Ni (1 0 0) surfaces. atoms were receiving electrons from the metal surfaces while me-
Surface Sites Ead (eV) Ni–O distance Cr–O distance tal atoms were donating electrons from their outer orbitals. The
oxygen atoms were receiving 0.58 and 0.66e from the nickel and
Ni (1 0 0) Hollow 5.69, 5.63a, 5.00b 1.96
Bridge 5.12 1.78 Ni–Cr surfaces, respectively. We attempted to elucidate the nature
Top 3.98 1.67 of bonding between the bare metal and the oxygen, OH and H2O
Ni-top1Cr (1 0 0) Hollow 6.33 2.05 1.80 adsorbed surfaces. The top most layer nickel and the adsorbed oxy-
Bridge 6.31 1.94 1.67 gen, OH and water molecules were projected to calculate the par-
Top 6.15 1.60 tial density of states (pDOS), as shown in Fig. 4. The pDOS of the Ni
Ni-top2Cr(1 0 0) Hollow 6.82 2.37 1.82 (1 1 1) surface has been published elsewhere [24]. The Ni (1 0 0) and
Bridge 6.22 1.91 1.69 Ni (1 1 0) bare surfaces showed that the 3d orbitals were highly
Top 6.04 1.59 localised at the Fermi level. The Ni 4s orbitals had a small contribu-
a
Ref. [53]. tion below the Fermi level. There was a small overlapping between
b
Ref. [54]. the Ni s and p orbitals near the Fermi energy, as shown in Figs. 4a
24 N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31

Table 5
OH adsorption on Ni (1 0 0) and Cr-doped Ni (1 0 0) surfaces.

Surface Sites Ead (eV) Ni–O distance Cr–O distance


Ni (1 0 0) Hollow 3.37 2.14
Bridge 3.22 1.92
Top 2.97 1.81
Ni-top1Cr (1 0 0) Hollow 3.88 2.25 1.91
Bridge 3.91 1.98 1.88
Top 3.79 1.79
Ni-top2Cr (1 0 0) Hollow 4.15 2.28 2.06
Bridge 3.95 1.95 1.92
Top 3.82 1.79

Table 6
H2O adsorption on Ni (1 0 0) and Cr-doped Ni (1 0 0) surfaces.

Surface Sites Ead (eV) Ni–O distance Cr–O distance


Ni (1 0 0) Hollow 0.36 3.17
Bridge 0.39 2.98
Top 0.56, 0.51a 2.21
Ni-top1Cr (1 0 0) Top 0.90 2.15
Ni-top2Cr (1 0 0) Top 0.67 2.17
a
Ref. [40].

was observed between the Ni p and oxygen 2p orbitals. Earlier the-


8
oretical studies on O/Ni (1 0 0) demonstrated similar bonding be-
tween Ni and oxygen [70]. Soft X-ray emission spectroscopy
Oxygen adsorption energy, Ead (eV)

Two Cr-doped Two Cr-doped (SXES) analysis demonstrated the hybridisation between the oxy-
7 (100) One Cr-doped (110) gen 2p and Ni 3d states [71]. The oxygen 2p states were also ob-
One Cr-doped (110) served to interact with the metal s-p states. The bonding nature
(100) changed when Cr was doped onto the nickel surfaces. The Cr 3d
6 orbitals were more pronounced for the two chromium doped sur-
Ni (100)
face, as shown in Fig. 4j. The strong hybridisation between the Cr
Ni (110) 3d and oxygen 2p orbitals was clearly observed while a small
amount of donation was due to the Ni d and s, and Cr s and p orbi-
5
tals. A single chromium doped nickel surface demonstrated the
interaction between the oxygen 2p and Cr 3d, Ni 3d and 4s orbitals.
The pDOS analysis suggests the strong hybridisation between the
4 oxygen 2p and chromium 3d orbitals; concurrently, metal s–p orbi-
Surfaces tals donated a small amount electrons, which might increase the
adsorption energy. Upon OH adsorption, the electrons were trans-
Fig. 3. Comparison of the oxygen adsorption energies between two different
ferred from the metal surface to the oxygen of the OH. The OH 3r
surfaces.
and 1p orbitals were found below the Fermi energy. The OH ad-
sorbed on Ni surfaces showed the bonding between Ni d and s
and b. These peaks are less pronounced for the top layer atoms. It and oxygen 2p orbitals. A clear small Ni p peak was also found at
seems that the 3d orbitals are mainly dominant in the bonding 5.86 eV. The hydrogen 1s orbitals were not interacting with the
while the s and p orbitals contribute a little for bare metal surfaces. metal atoms. The Ni 3d and 4s orbitals were the main contributors
The DFT study of nickel clusters demonstrated the similar orbitals while the p orbitals were less significant. Moreover, the Cr 3d and
interactions [69]. The Ni–Cr surfaces showed a strong interaction Ni 3d and 4s orbitals interacted with the oxygen 2p orbitals for the
between the Ni 3d and Cr 3d orbitals and the Ni s small peak hybri- OH adsorbed surfaces. UPS studies have led to the assignment of
dised with the Cr 3d orbitals. With further increase in the chro- levels 5.5 and 9.2 eV below the Fermi energy as p-bonding and
mium concentration on the top layer, the orbital interactions r-bonding levels for OH adsorption on Ni (1 1 0) [42]. Avdeev
changed slightly due to the increasing contribution from the Cr d et al. found the admixture of Ni 3d and 4s with oxygen 2p orbitals
and s orbitals. However, the Cr-doped surface bonds were mainly [72]. We also found the hybridisation between the oxygen 2p and
influenced by 3d orbitals. The pDOS of Ni and Ni–Cr surfaces were Ni 3d and 4s orbitals. Additionally, a small Ni p orbitals peak was
modified to adsorb atomic oxygen. The oxygen 2p orbitals and Ni found in this study. A small amount of charge was transferred from
3d and 4s orbitals were interacting for the Ni (1 0 0) surface while the water molecule to the metal surface. Water donated a small
a strong interaction was observed between the oxygen 2p and Ni amount of charge (0.07e) to the surface. A DFT study demon-
3d orbitals of Ni (1 1 0). A small Ni p peak appeared at 5.9 eV strated that the 0.03e charge was transferred from water to the
for the oxygen adsorbed Ni (1 0 0) surface, as shown in Fig. 4i. In surfaces [22]. Water adsorption on the Ni and Ni–Cr surfaces had
the case of Ni (1 1 0), the Ni p orbitals had no effect on the bonding three peaks below the Fermi energy, corresponding to the 1b2,
process. These orbitals bonding might affect the oxygen chemi- 3a1 and 1b1 orbitals, respectively. The pDOS curves for the nickel
sorption energy. However, the Ni 3d orbitals were hybridised with surfaces were slightly modified after adsorbing water molecules.
the oxygen 2p orbitals, and the Ni s orbitals also showed a peak be- The Ni 4s and 3d orbitals were the main contributors in the bond-
low the Fermi energy at 5–6 eV. In addition, a small interaction ing process. In addition, Ni p orbitals also showed a little contribu-
N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31 25

250 250
Ni of 4s Ni of 4s
Ni of 3p Ni of 3p
Ni of 3d Ni of 3d
200 200

pDOS (electrons/eV)

pDOS (electrons/eV)
150 150

100 100

50 50

0 0
-10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0 -10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0
Energy (eV) Energy (eV)
(a) (b)
Ni of 4s
Ni of 4s

pDOS (electrons/eV)
Ni of 3p
pDOS (electrons/eV)

Ni of 3p
Ni of 3d
40 Ni of 3d
40 Cr of 4s
Cr of 3p
Cr of 3d

20 20

250 250
-7.5 -5.0 -2.5 0.0 Ni of 4s -7.5 -5.0 -2.5 0.0 Ni of 4s
Ni of 3p Ni of 3p
Energy (eV) Ni of 3d Energy (eV) Ni of 3d
200
pDOS (electrons/eV)

200
pDOS (electrons/eV)

Cr of 4s
Cr of 3p
Cr of 3d
150 150

100 100

50 50

0 0
-10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0 -10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0
Energy (eV) Energy (eV)
(c) (d)
Ni of 4s Ni of 4s
pDOS (electrons/eV)
pDOS (electrons/eV)

Ni of 3p Ni of 3p
Ni of 3d Ni of 3d
40 Cr of 4s 40
Cr of 3p
Cr of 3d

20 20

250 250
-7.5 -5.0 -2.5 0.0 Ni of 4s -7.5 -5.0 -2.5 0.0 Ni of 4s
Ni of 3p Ni of 3p
Energy (eV) Ni of 3d Energy (eV) Ni of 3d
200 200
pDOS (electrons/eV)

pDOS (electrons/eV)

Cr of 4s
Cr of 3p
Cr of 3d
150 150

100 100

50 50

0 0
-10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0 -10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0
Energy (eV) Energy (eV)
(e) (f)
Fig. 4. Partial density of states of (a) all atoms of Ni (1 0 0), (b) all atoms of Ni (1 1 0), (c) top layer atoms of Ni (1 0 0), (d) top layer atoms of one Cr-doped Ni (1 0 0), (e) top layer
atoms of two Cr-doped Ni (1 0 0), (f) top layer atoms of Ni (1 1 0), (g) top layer atoms of one Cr-doped Ni (1 1 0), (h) top layer atoms of two Cr-doped Ni (1 1 0), (i) projected
atoms of oxygen adsorbed Ni (1 0 0), (j) projected atoms of oxygen adsorbed two Cr-doped Ni (1 0 0), (k) projected atoms of OH adsorbed Ni (1 1 0), (l) projected atoms of OH
adsorbed two Cr-doped Ni (1 1 0), (m) projected atoms of H2O adsorbed Ni (1 1 0), and (n) projected atoms of H2O adsorbed two Cr-doped Ni (1 1 0) surfaces.
26 N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31

250 250
Ni of 4s Ni of 4s
Ni of 3p Ni of 3p
Ni of 3d Ni of 3d
200 200

pDOS (electrons/eV)
pDOS (electrons/eV) Cr of 4s Cr of 4s
Cr of 3p
Cr of 3p
Cr of 3d Cr of 3d
150 150

100 100

50 50

0 0
-10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0 -10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0
Energy (eV) Energy (eV)
(g) (h)

Ni of 4s Ni of 4s

pDOS (electrons/eV)
pDOS (electrons/eV)

Ni of 3p Ni of 3p

40 Ni of 3d
O of 2s
40 Ni of 3d
Cr of 4s
O of 2p Cr of 3p
Cr of 3d
O of 2s
O of 2p

20 20

250 250
-7.5 -5.0 -2.5 0.0 Ni of 4s -7.5 -5.0 -2.5 0.0 Ni of 4s
Ni of 3p Ni of 3p
Energy (eV) Ni of 3d
Energy (eV) Ni of 3d
200 200
pDOS (electrons/eV)
pDOS (electrons/eV)

O of 2s Cr of 4s
O of 2p Cr of 3p
Cr of 3d
150 150 O of 2s
O of 2p

100 100

50 50

0 0
-10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0 -10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0
Energy (eV) Energy (eV)
(i) (j)
250 250
Ni of 4s Ni of 4s
Ni of 3p Ni of 3p
Ni of 3d Ni of 3d
200 200
pDOS (electrons/eV)
pDOS (electrons/eV)

O of 2s Cr of 4s
O of 2p Cr of 3p
H of 1s Cr of 3d
150 150 O of 2s
O of 2p
H of 1s
100 100

50 50

0 0
-10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0 -12.5 -10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0
Energy (eV) Energy (eV)

(k) (l)
Fig. 4. (continued)

tion, which may not be negligible. The H2O/Ni (1 1 1) theoretical tals peak was slightly decreased with increased chromium concen-
investigation demonstrated that the Ni p states were involved in tration while the Cr s–p–d orbitals increased the interaction with
the bonding with oxygen 2p states, which is in line with the pres- water molecules. This study reveals that the transitional metal d
ent study [73]. Moreover, several studies suggested that surface s– and s orbitals were essentially in bonding with water while the p
d hybridisation is a general feature in water metal interaction [74]. orbitals showed little contribution. The metal surface and water
The Ni d along with the Cr d, s, and p orbitals interacted with water interactions were found to be very weak, resulting in a small
1b1 orbitals for one chromium doped nickel surface. The Ni s orbi- adsorption energy.
N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31 27

Ni of 4s Ni of 4s
Ni of 3p Ni of 3p

pDOS (electrons/eV)

pDOS (electrons/eV)
Ni of 3d Ni of 3d
40 H2O 40 Cr of 4s
Cr of 3p
Cr of 3d
H2O

20 20

250 250
-7.5 -5.0 -2.5 0.0 Ni of 4s -7.5 -5.0 -2.5 0.0 Ni of 4s
Ni of 3p Energy (eV) Ni of 3p
Energy (eV) Ni of 3d Ni of 3d
200 200
pDOS (electrons/eV)

pDOS (electrons/eV)
H2O Cr of 4s
Cr of 3p
Cr of 3d
150 150 H2O

100 100

50 50

0 0
-12.5 -10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0 -12.5 -10.0 -7.5 -5.0 -2.5 0.0 2.5 5.0 7.5 10.0
Energy (eV) Energy (eV)
(m) (n)
Fig. 4. (continued)

3.2. Tight-binding QCMD analysis of the water/metal interfaces layers of the metal surface. The diffusivity of oxygen into Ni–Cr
(1 1 0) surface is the highest among the all surfaces. One of the dis-
The tight-binding self-consistent field is a worthy method to sociated oxygen atoms reached the bottom of the Ni–Cr (1 1 0) sur-
describe the reaction dynamics of metal/solution interfaces at high face. We have measured the diffusivity of oxygen along the z-axis,
temperature. as shown in Fig. 6a and b. The penetration depth was measured
from the initial position of the top layer metal atoms. One of the
3.2.1. Metal/water interface reaction dynamics oxygen atoms did not diffuse into the surface, which was not con-
In the case of the adsorption of water molecules, OH and atomic sidered for this plot. The travelling time of the oxygen atoms to ac-
oxygen, different sites for the Ni (1 1 1), Ni (1 1 0) and Ni (1 0 0) sur- quire the maximum distances were 980 fs, 1020 fs and 1200 fs for
faces are the energetically most stable sites. Previous studies dem- the Ni–Cr (1 1 1), Ni–Cr (1 1 0), and Ni–Cr (1 0 0) surfaces, respec-
onstrated that threefold hollow, short bridge and fourfold hollow tively. The oxygen penetration depth was slightly increased due
sites are the most stable for atomic oxygen adsorption on Ni to applied uniaxial strain on the surfaces. The applied strain gener-
(1 1 1), Ni (1 1 0) and Ni (1 0 0) surfaces, respectively [21,37,75]. ates more space in the lattice, promoting the penetration of oxygen
OH preferred to stay on fcc hollow, SB and four fold hollow sites into the slab. The oxygen atom was trapped around the chromium
for (1 1 1), (1 1 0) and (1 0 0) surfaces, respectively while the top site and created a preferential bond that could help to reduce its diffu-
was the most favourable for water molecule adsorption [34,67]. At sivity, which indicates that 1% uniaxial strain increased the oxygen
the early stage, water molecules rotate from their original position penetration depth into the structure.
in which hydrogen atoms are directed toward the surface. Fig. 5 Fig. 6 clearly shows that the depth of the penetration of oxygen
shows the calculated morphology of different Ni–Cr surfaces as a in the Ni–Cr (1 1 0) surface is larger than that of the other surfaces.
function of time. The calculated results for other surface morphol- The oxygen penetrated a short distance into the Ni–Cr (1 1 1) sur-
ogies are similar to those in the figure. Water molecules are placed face. The larger diffusivity of oxygen reveals that the surface is easy
at randomly on surface and are free to rotate. At the initial stage, to oxidise by the oxygen atom; as a result, the metal surface can be
water molecules were adsorbed by the surfaces, and some of them damaged. The analysis of the Ni–Cr (1 1 1) surface shows that the
dissociated into OH and H ions. Subsequently some of the OH ions oxygen atom is trapped around the chromium atom; correspond-
produce O and H. The metal–hydrogen bond strength is stronger ingly, the process reduces the atomic diffusivity. However, the
than that of metal–water but weaker than those of metal–hydroxyl Ni–Cr (1 1 0) surface has the terrace site like atomic defect, where
and metal–oxygen. A lower bond strength helps to increase the dif- foreign atoms or molecules can be trapped easily. This defect en-
fusivity of hydrogen into the structure. Dissociated hydrogen is hances the oxygen diffusivity into the structure. Previous QCMD
small in atomic size and can diffuse into the surfaces through the results showed that the atomic defect increased the oxygen diffu-
hollow sites, and it can be trapped in interstitial sites. Some of sivity in the metal surface [76]. Additionally, the Ni–Cr (1 1 1) and
the hydrogen cannot diffuse due to coulombic repulsion with the Ni–Cr (1 0 0) surfaces are more dense than the Ni–Cr (1 1 0) surface,
surface metallic atoms at the beginning. The interaction of dissoci- which may reduce the oxygen diffusivity into the structure [77].
ated oxygen and hydroxyl with metal atoms weakens the metal– The dense structures do not have sufficient space to adsorb foreign
metal bond, leading to the initiation of possible dissolution. The atoms. Norton et al. demonstrated that the oxygen uptake rate was
reaction progress with respect to time showed that the top layer in the order (1 0 0) < (1 1 1) < (1 1 0) [78]. This experimental obser-
atoms moved upward, which produced extra space into the struc- vation reveals that oxygen can penetrate deeper into the Ni
ture. The process helps to diffuse oxygen or hydroxyl ions. The sim- (1 1 0) surface than other surfaces; subsequently, the process might
ulated results showed that the oxygen atom penetrated several increase the surface oxidation rate. Moreover, the DFT study found
28 N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31

Cr-O/OH bond Diffused O surface. One can conclude from this study that Ni (1 1 0) is the
weakest surface for oxidation initiation. In addition, the calculated
surface morphologies showed the high oxygen concentration
around the chromium atom, which indicates that chromium is
selectively oxidised during the reaction progress and the oxygen
favourably forms a bond with chromium. This bonding initiates
the localised passive film formation on the metal surface. One of
the experimental studies showed that Cr2O3 is a major product
even for alloys containing as little as 0.1% Cr [79]. The present
study is not be able to quantify the oxide film thickness due to
the limitations of time and the length scale, but the simulated re-
sults qualitatively observe the initiation of passive film formation
on the surface. A short period oxidation study has been carried
200 fs 800 fs 1600 fs out on alloys 600 and 690 [80]. The results showed that a layer
(a) of oxide film is formed very quickly on the bare metal surface.
The oxide film is composed of chromium oxide and hydroxide
O/OH adsorbed on terrace site Diffused O and is covered with a small amount of nickel hydroxide. The XPS
study of Ni–Cr (1 1 0) polycrystalline demonstrated that Cr is oxi-
dised preferentially to Cr2O3 initially [11]. In addition, scanning
electron microscopy (SEM) analysis of alloy 600 showed that chro-
mium and iron are selectively dissolved and nickel remains at the
surface of the matrix [81]. The present results are qualitatively in
good agreement with the experimental observation. The diatomic
energy of Cr–O is higher than the Ni–O energy; therefore, it can
help oxygen to bond preferentially with chromium. The calculated
surfaces found that the chromium had moved upward to the sur-
200 fs 800 fs 1600 fs
face because it dissoluted faster than nickel atoms. The Cr–O bond
(b) helps to move the atoms in the upward direction. In addition, the
cohesive energy of chromium is less than that of nickel, which
Dissociated H2O H remain on the surface might help to dissolute chromium atoms from the surface. The
oxygen chemisorption process weakens the metallic bond; as a
consequence, metal–metal bond breakage and initiates the surface
oxidation. The simulated surfaces showed that the hydrogen is
trapped in the structure. It can diffuse faster than oxygen or any
other species due to its small atomic size. Hydrogen can be ad-
sorbed by the surface as atomic hydrogen or OH. The adsorption
of OH on the metal surface generates metal hydroxides such as
chromium hydroxide or nickel hydroxide [7]. Hydrogen can be
trapped in the interstitial sites of fcc Ni–Cr alloys. Ni–Cr alloys have
numerous octahedral and tetrahedral sites to adsorb hydrogen.
These interstitial hydrogen has detrimental effect on material
200 fs 800 fs 1600 fs properties such as hydrogen embrittlement and hydrogen oxida-
(c) tion. The simulated surfaces demonstrated that the dissociated
hydrogen travelled ahead of the oxygen or OH. These atoms can
Fig. 5. The calculated surface morphologies: (a) Ni–Cr (1 1 1), (b) Ni–Cr (1 1 0) and be able to lessen the strength of metallic bonds. Thus, they can help
(c) Ni–Cr (1 0 0). to increase the oxygen penetration depth in the metal.

that the oxygen chemisorption energy is 5.69 eV, 5.56 and 5.22 eV 3.2.2. Atomic dissolution and diffusion path
for Ni (1 0 0), Ni (1 1 1) and Ni (1 1 0), respectively. It also indicates The dynamical feature of atoms is depicted in Fig. 7. Oxygen, in
that the atomic oxygen make the weakest bond on the Ni (1 1 0) the form of H2O, OH or O, diffuses through the hollow sites. The

-7.0 -7.0
-6.0 111 Ni-Cr(111) 980 fs 111 Ni-Cr(111) 980 fs
-6.0
110 Ni-Cr(110) 1020 fs 110 Ni-Cr(110) 1060 fs
-5.0 Ni-Cr(100) 1200 fs -5.0
100 100 Ni-Cr(100) 960 fs
-4.0 -4.0
Depth (Å)

Depth (Å)

-3.0 -3.0
-2.0 -2.0
-1.0 -1.0 O1
O1
0.0 0.0
O3 O2 O3
O2
1.0 1.0
2.0 2.0
(a) (b)
Fig. 6. Maximum penetration depth of oxygen for different surfaces: (a) surfaces without applied strain and (b) surface with applied uniaxial strain.
N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31 29

experimental and theoretical studies showed that the hollow site is Negatively charged H
energetically the most preferential for oxygen adsorption on Ni
and Ni–Cr surfaces [18,24]. Dissociated oxygen or OH was ad-
sorbed on the hollow sites and formed a strong bond with metal
atoms. The process helps to dissolute metal atoms. However, the
trajectory of oxygen clearly showed that atoms preferred to remain
around the chromium, and the bonding decelerates further diffu-
sion, as we have described in the previous section. Chromium is
selectively oxidised at the initial stage of the oxidation process
and can form a protective passive film. In the case of metal atom 200 fs 600 fs
dissolution, chromium atoms attempted to move upward from Cr-O/OH bond
their initial position, whereas the nickel segregation rate was slow. +0.968
The present results confirm us that the chromium atoms selec-
tively dissolute and the nickel remain at the surface of alloys at
0.000
the early stage of oxidation. The AES study of Ni–Cr alloys showed
that chromium depletion occurred in gaseous atmosphere at high
temperature [82,83]. The calculated results are in good agreement -0.968
with the AES observation. The process indicated the transport of Cr
to inject vacancies into the base metal, caused by the growth of the 1600 fs
oxide layer.
Fig. 8. Atomic charge changes with respect to time for the Ni–Cr (1 1 0) surface.

3.2.3. Atomic charge analysis


The degree of electron transfer is an important issue to under- Furthermore, the DFT study confirmed that the oxygen 2p orbitals
stand the bonding nature of atoms. The positively charged atoms hybridised strongly with the nickel 3d orbitals [37]. In the case of
can a donate higher number of electrons to the acceptor, whereas chromium, the orbital bonding nature is slightly different. The
the highly negatively charged atoms can receive electrons from the DFT study demonstrated that the charge transfer occurs from the
donors. The process helps to form a strong bond between highly Cr 3d and 4s orbitals to the oxygen 2p orbitals. The Cr–O orbitals
positively charged and highly negatively charged atoms. The atom- interaction enhances the electron transfer between them and
ic charge with diffusion is calculated by using Mulliken population makes a strong bond. The top layer atoms played an active role
analysis method, as shown in Fig. 8. The calculated results for other in the reaction process at the beginning stage, which increased
surfaces agree with this figure, and thus it is not necessary to show the electron transfer. The hydrogen atom on top of the surfaces
them here. The water–metal bond has been interpreted as a Lewis exhibited positively charged, whereas the penetrated hydrogen
acid–base interaction, with the H2O in the role of the electron do- was negatively charged. Negatively charged hydrogen in metal is
nor [84]. The electron transfer from water to the nickel substrate is interesting but not surprising, and it is quite consistent with our
only 0.1e [34]. Our calculations considered the surface reaction previous results [33,76]. The calculated average charge of the dif-
dynamics at high temperature. The water molecules are adsorbed fused hydrogen atom was 0.20e. the first principles study shows
by the surface for a very short period of time. Thus, it is difficult that a charge transfer of 0.15–0.20e to the hydrogen atom from the
to observe this phenomenon. However, at the elementary stage, nickel surface occurs [87]. However, the positively charged hydro-
chromium and hydrogen were positive in charge, and oxygen gen is well known in the literature and causes decohesion. The
atoms were negative in charge; in addition, nickel atoms were hydrogen decohesion mechanism causes embrittlement. According
not designated to any remarkable charges. The atomic charge to the anionic model, hydrogen receives electrons from the metal
was modified with respect to time evolution. The negative oxygen and changes to negatively charged [33]. Therefore, the deeply dif-
atoms increased in value by taking electrons from metals, and me- fused hydrogen interacts with metallic atoms for a comparatively
tal atoms became positively charged. In particular, chromium longer period of time and takes electrons from the metal. The elec-
atoms were positive in charge. The calculated oxygen atom average tron transfer process activates the surface chemically for further
charge was 0.48e. The net electron flow, as calculated by the reaction, which makes it easier for oxygen to diffuse, and thus oxi-
topological method, was from the nickel surface to the oxygen dation accelerates.
atoms, amounting to 0.86e per ad-atom [85]. Another theoretical
study found that the oxygen atom received 0.52e from the nickel
surface [86]. Our results agree fairly well with the previous study. 4. Conclusions

O diffusion path Cr dissolution Ni dissolution We have studied the adsorption characteristics of atomic oxy-
gen, OH, and water molecules on different Ni and Ni–Cr surfaces
by DFT calculations. The results show that the oxygen adsorption
energy on Ni (1 0 0) is higher than that on the Ni (1 1 1) and unre-
constructed Ni (1 1 0) surfaces. The Cr substituted nickel surface in-
creases the oxygen adsorption energies. The maximum adsorption
energy is attained when two chromium atoms are doped on top of
the Ni (1 1 1) surface. The oxygen and OH adsorption energies of Ni
(1 1 0) and Ni–Cr (1 1 0) surfaces are smaller than those of the other
surfaces. Thus the (1 1 0) surface makes a weak bond with the ad-
sorbed oxygen and OH. The most favourable oxygen adsorption
sites are the fcc hollow, fourfold hollow and SB sites of (1 1 1),
(1 0 0) and (1 1 0) surfaces, respectively. The OH is adsorbed molec-
ularly on the hollow site of (1 1 1) and (1 0 0) and the SB site of
Fig. 7. Atomic trajectories. (1 1 0) surfaces. The OH adsorption energies of the Ni–Cr surfaces
30 N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31

are also increased. The (1 1 0) bare surface demonstrates a higher [7] G.B. Hoflund, W.S. Epling, Oxidation study of a polycrystalline Ni/Cr alloy II,
Chem. Mater. 10 (1998) 50–58.
percentage of contraction and expansion between the interlayers
[8] S.E. Ziemniak, M. Hanson, Corrosion behavior of NiCrFe alloy 600 in high
compared to the (1 1 1) and (1 0 0) surfaces while the O, OH and temperature, hydrogenated water, Corros. Sci. 48 (2006) 498–521.
H2O adsorbed metal surfaces show the outward expansion be- [9] F. Huang, J.Q. Wang, E.H. Han, W. Ke, Short-time oxidation of alloy 690 in high-
tween the first and second layers and a slight contraction between temperature and high-pressure steam and water, J. Mater. Sci. Technol. 28
(2012) 562–568.
the second and third layers. The bond distances of Cr–O are smaller [10] M. Juez-Lorenzo, V. Kolarik, W. Stamm, H. Fietzek, Oxidation of nickel-based
than the Ni–O distances for oxygen, OH, and H2O adsorption, which alloys in dry and water vapour containing air, Mater. High Temp. 29 (2012)
implies that the oxygen atom has some affinity to chromium. The 229–234.
[11] S.P. Jeng, P.H. Holloway, D.A. Asbury, G.B. Hoflund, Changes induced at Ni/Cr
Cr 3d orbitals bond strongly with the oxygen 2p orbitals; concur- alloy surfaces by annealing and oxygen exposure, Surf. Sci. 235 (1990) 175–
rently, the Ni 3d and 4s orbitals bond with the oxygen 2p orbitals. 185.
A small Ni p peak appears at 5.9 eV for the oxygen adsorbed Ni [12] S.P. Jeng, P.H. Holloway, C.D. Batich, Surface passivation of Ni/Cr alloy at room
temperature, Surf. Sci. 227 (1990) 278–290.
(1 0 0) surface. In the case of O/Ni (1 1 0), the Ni p orbitals have [13] C.W. Bauschlicher Jr., H2O/Ni(1 0 0) and NH3/Ni(1 0 0): a computational
no effect on the bonding process. The bonding of those orbitals approach, J. Chem. Phys. 83 (1985) 3129–3133.
might the cause of the reduction of the oxygen chemisorption en- [14] D. Sebastiani, L.D. Site, Adsorption of water molecules on flat and stepped
nickel surfaces from first principles, J. Chem. Theor. Comput. 1 (2005) 78–82.
ergy. The Cr 3d and the Ni 3d and 4s orbitals interact with the oxy- [15] M. Nakamura, M. Tanaka, M. Ito, O. Sakata, Water adsorption on a p(2  2)-
gen 2p orbitals on the OH adsorbed surfaces. The water adsorbed Ni(1 1 1)–O surface studied by surface X-ray diffraction and infrared
surfaces show that the metallic d and s orbitals bond essentially reflection absorption spectroscopy at 25 and 140 K, J. Chem. Phys. 122
(2005) 224703.
with water while the p orbitals show little contribution. The metal
[16] C. Taylor, R.G. Kelly, M. Neurock, First-principles calculations of the
surfaces interact very weakly with the water molecules, resulting electrochemical reactions of water at an immersed Ni(1 1 1)/H2O interface, J.
in a small adsorption energy. Electrochem. Soc. 153 (2006) E207–E214.
The QCMD results show that the depth of penetration of oxygen [17] G.C. Wang, S.X. Tao, X.H. Bu, A systematic theoretical study of water
dissociation on clean and oxygen-preadsorbed transition metals, J. Catal. 244
in the Ni–Cr (1 1 0) surface is the highest among all surfaces. The ter- (2006) 10–16.
race like atomic defects on the Ni–Cr (1 1 0) surface help to trap oxy- [18] S. Yamagishi, S.J. Jenkins, D.A. King, First principles studies of chemisorbed O
gen or dissociated ions. Thus, oxygen permeates deeper into the on Ni {1 1 1}, Surf. Sci. 543 (2003) 12–18.
[19] T. Pache, H.-P. Steinrück, W. Huber, D. Menzel, The adsorption of H2O on clean
structure through these sites. Moreover, the Ni–Cr (1 1 1) and Ni– and oxygen precovered Ni(1 1 1) studied by ARUPS and TPD, Surf. Sci. 224
Cr (1 0 0) surfaces are denser than the Ni–Cr (1 1 0), which might (1989) 195–214.
be another reason to increase the oxygen diffusivity into the struc- [20] S. Garruchet, O. Politano, P. Arnoux, V. Vignal, Diffusion of oxygen in nickel: a
variable charge molecular dynamics study, Solid State Commun. 150 (2010)
ture. In addition, a 1% applied strain increases the oxygen diffusivity 439–442.
into the surfaces. The surface morphology shows the significantly [21] A.V. Kalinkin, G.K. Boreskov, V.I. Savchenko, K.A. Dadayan, N.N. Bulgakov,
high oxygen concentration around the chromium, which indicates React. Kinet. Catal. Lett. 18 (1981) 55–58.
[22] J. Carrasco, A. Michaelides, M. Scheffler, Insight from first principles into the
the early stage of passive film formation. The process helps to in- nature of the bonding between water molecules and 4d metal surfaces, J.
crease the chromium dissolution rate. Lower nickel segregation Chem. Phys. 130 (2009) 184707.
may help to form a nickel enriched inner oxide layer. In addition, [23] W. An, X.C. Zeng, C.H. Turner, First-principles study of methane
dehydrogenation on a bimetallic Cu/Ni(1 1 1) surface, J. Chem. Phys. 131
the charge analysis results show that the chromium atoms make
(2009) 174702.
a strong bond with highly negatively charged oxygen atoms, and [24] N.K. Das, T. Shoji, A density functional study of atomic oxygen and water
the deeply diffused hydrogen is negatively charged. The initially molecule adsorption on Ni(1 1 1) and chromium-substituted Ni(1 1 1) surfaces,
penetrated hydrogen weakens the metallic bond by taking electron, Appl. Surf. Sci. 258 (2011) 442.
[25] N.K. Das, I. Tirtom, T. Shoji, A multi-scale modeling study of a Ni–Cr(1 1 1)
and the process makes it easier for oxygen to further permeate the surface oxidation at different stress intensities, Mater. Chem. Phys. 122 (2010)
surface. Finally, this study shows that the atomic oxygen makes the 336–342.
weakest bond on the (1 1 0) surface; simultaneously, oxygen can [26] P. Hohenberg, W. Kohn, Inhomogeneous electron gas, Phys. Rev. B 136 (1964)
864–871.
permeate longer distance into the (1 1 0) surface. As a consequence, [27] B. Delley, An all-electron numerical method for solving the local density
the surface is susceptible to the oxidation initiation process. functional for polyatomic molecules, J. Chem. Phys. 92 (1990) 508–517.
[28] B. Delley, A scattering theoretic approach to scalar relativistic corrections on
bonding, Int. J. Quantum Chem. 69 (1998) 423–433.
[29] B. Delley, From molecules to solid with the DMol3 approach, J. Chem. Phys. 113
Acknowledgements
(2000) 7756–7764.
[30] J.P. Perdew, Y. Wang, Accurate and simple analytic representation of
DFT calculation has been performed by supercomputing re- the electron-gas correlation energy, Phys. Rev. B 45 (1992) 13244–13249.
[31] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made
sources. The authors gratefully acknowledge SR11000 supercom-
simple, Phys. Rev. Lett. 77 (1996) 3865–3868.
puting resources from the Center for Computational Materials [32] W.J. Hehre, L. Radom, P.V.R. Schleyer, J.A. Pople, Ab initio Molecular Orbital
Science of the Institute for Materials Research, Tohoku University. Theory, Wiley, New York, 1986.
We also would like express our sincere thanks to the crew of this [33] N.K. Das, T. Shoji, An atomic study of hydrogen effect on the early stage
oxidation on transition metal surfaces, Int. J. Hydrogen Energy. 38 (2013)
center. 1644–1656.
[34] H. Yang, J.L. Whitten, The adsorption of water and hydroxyl on Ni(1 1 1), Surf.
Sci. 223 (1989) 131–150.
References [35] H. Yang, J.L. Whitten, Adsorption of SH and OH and coadsorption of S, O and H
on Ni(1 1 1), Surf. Sci. 370 (1997) 136–154.
[36] S. Masuda, M. Nishijima, Y. Sakisaka, M. Onchi, Structures of oxygen-covered
[1] M. Nakamura, M. Ito, Monomer water structures of water adsorbed on
Ni (1 1 0) surfaces, Phys. Rev. B 25 (1982) 863–874.
p(2  2)-Ni(1 1 1)–O surface at 25 and 140 K studied by surface X-ray
[37] W.-B. Zhang, B.-Y. Tang, First-principles studies of the oxygen adsorption on
diffraction, Phys. Rev. Lett. 94 (2005) 035501.
unreconstructed and reconstructed Ni(1 1 0) surfaces, Surf. Sci. 603 (2009)
[2] L. Ollé, M. Salmerón, A.M. Baró, The adsorption and decomposition of water on
1002–1009.
Ni (1 1 0) studied by electron energy loss spectroscopy, J. Vac. Sci. Technol., A 3
[38] K.-H. Xiang, R. Pandey, J.M. Recio, E. Francisco, J.M. Newsam, A theoretical
(1985) 1866–1870.
study of the cluster vibrations in Cr2O2, Cr2O3, and Cr2O4, J. Phys. Chem. A 104
[3] M.E. Gallagher, S. Haq, A. Omer, A. Hodgson, Water monolayer and multilayer
(2000) 990–994.
adsorption on Ni (1 1 1), Surf. Sci. 601 (2007) 268–273.
[39] M.L. Peterson, A.F. White, G.E. Brown, G.A. Parks, Surface passivation of
[4] C. Benndorf, C. Nöbl, M. Rusenberg, F. Thieme, H2O interaction with clean and
magnetite by reaction with aqueous Cr(VI): XAFS and TEM results, Environ.
oxygen precovered Ni (1 1 0), Surf. Sci. 111 (1981) 87–101.
Sci. Technol. 31 (1997) 1573–1576.
[5] R.W. Staehle, in: J.C. Scully (Ed.), The Theory of SCC in Alloys, NATO, Brussels,
[40] C.D. Roux, H. Bu, J.W. Rabalais, Structure of the Ni{1 1 0}-p(2  1)–OH
1971, p. 86
surface from time of flight scattering and recoiling spectrometry, Surf. Sci.
[6] H. Over, A.P. Seitsonen, Oxidation of metal surfaces, Science 297 (2002) 2003–
279 (1992) 1.
2005.
N.K. Das, T. Shoji / Corrosion Science 73 (2013) 18–31 31

[41] C. Nöbl, C. Benndorf, T.E. Madey, H2O adsorption on Ni(1 0 0): evidence for [64] T.S. Rahman, D.L. Mills, J.E. Black, J.M. Szeftel, S. Lehwald, H. Ibach, Surface
oriented water dimers, Surf. Sci. 157 (1985) 29–42. phonons and the c(2  2) oxygen overlayer on Ni(1 0 0): theory and
[42] C. Benndorf, C. Nöbl, T.E. Madey, H2O adsorption on oxygen-dosed Ni(1 1 0): experiment, Phys. Rev. B 30 (1984) 589–603.
formation and orientation of OH(ad), Surf. Sci. 138 (1984) 292–304. [65] R.L. Strong, J.L. Erskine, Adsorbate structure determination using surface
[43] V.I. Avdeev, I.I. Zakharov, G.M. Zhidomirov, N.M. Neshev, E.I. Proinov, Ab-initio vibrational spectroscopy, Phys. Rev. Lett. 54 (1985) 346–348.
calculations for adsorbed state of OH group on Ni(1 1 0), React. Kinet. Catal. [66] H. Yang, J.L. Whitten, Energetics of hydroxyl and influence of coadsorbed
Lett. 45 (1) (1991) 61–66. oxygen on metal surfaces, J. Phys. Chem. B 101 (1997) 4090–4096.
[44] J.I. Siepmann, M. Sprik, Ordering of fractional monolayers of H2O on Ni(1 1 0), [67] T.E. Madey, F.P. Netzer, The adsorption of H2O on Ni(1 1 1); influence of
Surf. Sci. Lett. 279 (1992) L185–L190. preadsorbed oxygen on azimuthal ordering, Surf. Sci. 117 (1982) 549–560.
[45] B.W. Callen, K. Griffiths, R.V. Kasza, M.B. Jensen, P.A. Thiel, P.R. Norton, [68] W. Oed, U. Starke, K. Heinz, K. Müller, Adsorbate-induced reconstruction of
Structural phenomena related to associative and dissociative adsorption of O and S on Ni(1 0 0): a local model, J. Phys.: Condens. Matter 3 (1991) S223–
water on Ni(1 1 0), J. Chem. Phys. 97 (1992) 3760–3774. S230.
[46] N. Pangher, A. Schmalz, J. Haase, Structure determination of water [69] N.K. Das, T. Shoji, Geometry, orbital interaction, and oxygen chemisorption
chemisorbed on Ni(1 1 0) by use of X-ray absorption fine-structure properties of chromium-doped nickel clusters, J. Phys. Chem. C 116 (2012)
measurements, Chem. Phys. Lett. 221 (1994) 189–193. 13353–13367.
[47] M. Pozzo, G. Carlini, R. Rosei, D. Alfe, Comparative study of water dissociation [70] C.S. Wang, A.J. Freeman, Self-consistent electronic structure of chemisorption
on Rh(1 1 1) and Ni(1 1 1) studied with first principles calculations, J. Chem. bonding: c(2  2) O on Ni(0 0 1), Phys. Rev. B 19 (1979) 4930.
Phys. 126 (2007) 164706. [71] H. Tillborg, A. Nilsson, T. Wiell, N. Wassdahl, N. Mårtensson, J. Nordgren,
[48] M. Nakamura, M. Ito, Monomer structures of water adsorbed p(2  2)- Electronic structure of atomic oxygen adsorbed on Ni(1 0 0) and Cu(1 0 0)
Ni(1 1 1)–O surface at 25 and 140 K studied by surface X-ray diffraction, studied by soft-X-ray emission and photoelectron spectroscopies, Phys. Rev. B
Phys. Rev. Lett. 94 (2005) 035501. 47 (1993) 16464–16470.
[49] M.L. Xu, S.Y. Tong, Multilayer relaxation for the clean Ni(1 1 0) surface, Phys. [72] V.I. Avdeev, I.I. Zakharov, G.M. Zhidomirov, N.M. Neshev, E.I. Proinov,
Rev. B 31 (1985) 6332–6336. Electronic structure of OH adsorbed on nickel, J. Struct. Chem. 33 (2) (1992)
[50] D.L. Adams, L.E. Petersen, C.S. Sörensen, Oscillatory relaxation of the Ni(1 1 0) 179–185.
surface: a LEED study, J. Phys. C 18 (1985) 1753. [73] J. Li, S. Zhu, Y. Li, E.E. Oguzie, F. Wang, Electronic structure of monomeric water
[51] F. Mittendorfer, A. Eichler, J. Hafner, Structural, electronic and magnetic adsorption on Ni{1 1 1}: beyond the general model, J. Phys. Chem. C 112 (2008)
properties of nickel surfaces, Surf. Sci. 423 (1999) 1–11. 8301–8303.
[52] D.S.D. Gunn, S.J. Jenkins, Adsorbate modification of the structural, electronic, [74] Q.L. Tang, Z.X. Chen, Density functional slab model studies of water adsorption
and magnetic properties of ferromagnetic fcc 1 1 0 surfaces, Phys. Rev. B 83 on flat and stepped Cu surfaces, Surf. Sci. 601 (2007) 954.
(2011) 115403. [75] I. Panas, P. Siegbahn, U. Wahlgren, Model studies of the chemisorption of
[53] D. Brennan, D.O. Hayward, B.M.W. Trapnell, The calorimetric determination of hydrogen and oxygen on nickel surfaces, Theor. Chim. Acta 74 (1988) 167–
the heats of adsorption of oxygen on evaporated metal films, Proc. Roy. Soc. 184.
Lond. A 256 (1960) 81. [76] N.K. Das, K. Suzuki, Y. Takeda, K. Ogawa, T. Shoji, Quantum chemical molecular
[54] H. Conrad, G. Ertl, J. Kuppers, E.E. Latta, Interaction of NiO with a Ni(1 1 1) dynamics study of stress corrosion cracking behaviour for fcc Fe and Fe–Cr
surface, Surf. Sci. 50 (1975) 296. surfaces, Corros. Sci. 50 (2008) 1701–1706.
[55] W.F. Egelhoff Jr., Core-level binding – energy-shift analysis of adsorption and [77] L.H. Germer, A.U. MacRae, Surface reconstruction caused by adsorption, PNAS
dissociation, Phys. Rev. B 29 (1984) 3681–3683. 48 (1962) 997–1000.
[56] A. Goursot, F. Mele, N. Russo, D.R. Salahub, M. Toscano, Geometrical [78] P.R. Norton, R.L. Tapping, J.W. Goodale, A photoemission study of the
spectroscopic and magnetic properties of an oxygen atom adsorbed on the interaction of Ni(1 0 0), Ni(1 1 0) and Ni(1 1 1) surfaces with oxygen, Surf. Sci.
Ni(1 0 0) surface, Int. J. Quantum Chem. 48 (1993) 277–286. 65 (1977) 13–36.
[57] J.E. Demuth, D.W. Jepsen, P.M. Marcus, Chemisorption bonding of c(2  2) [79] T.D. Tribbeck, J.W. Linnett, P.G. Dickens, Oxidation of metals and alloys. Part 1.
chalcogen overlayers on Ni(0 0 1), Phys. Rev. Lett. 31 (1973) 540–542. – Oxidation of nickel containing small amounts of chromium, Trans. Faraday
[58] L. Wenzel, D. Arvanitis, W. Daum, H.H. Rotermund, J. Stöhr, K. Baberschke, H. Soc. 65 (1969) 890–895.
Ibach, Structural determination of an adsorbed-induced surface [80] P. Combrade, in: T.R. Allen, P.J. King, and L. Nelson (Eds.), Proceedings of the
reconstruction: p4g(2  2) versus c(2  2)O on Ni(1 0 0), Phys. Rev. B 36 12th International Conference on Environmental Degradation of Materials in
(1987) 7689–7692. Nuclear Power System – Water Reactors, TMS, 2005.
[59] W. Oed, H. Lindner, U. Starke, K. Heinz, K. Müller, Adsorbate-induced [81] T. Terachi, N. Totsuka, T. Yamada, T. Nakagawa, H. Deguchi, M. Horiuchi, M.
relaxation and reconstruction of c(2  2)O/Ni(1 0 0): a reinvestigation by leed Oshitani, J. Nucl. Sci. Technol. 40 (2003) 509–516.
structure analysis, Surf. Sci. 224 (1989) 179–194. [82] F. Carrette, F.L. Guinard, B. Pieraggi, in: Proceedings of the 9th International
[60] A.B. Hayden, P. Pervan, D.P. Woodruff, K-resolved inverse photoemission study Conference on Water Chemistry of Nuclear Reactor Systems, Avignon, France,
of half-monolayer structure of O, C and N on Ni(1 0 0): a fingerprint of SFEN, 2002.
adsorbate-induced restructuring?, Surf Sci. 306 (1994) 99–113. [83] M. Seo, N. Sato, Int. Congress Metal. Corros. 2 (1984) 362.
[61] J. Stohr, R. Jaeger, T. Kendelewicz, Structure of p(2  2) and c(2  2) oxygen on [84] P.A. Thiel, T.E. Madey, The interaction of water with solid surfaces:
Ni(1 0 0): a surface extended-X-ray absorption fine-structure study, Phys. Rev. fundamental aspects, Surf. Sci. Rept. 7 (1987) 211–385.
Lett. 49 (1982) 142–146. [85] R.F.W. Bade, Atoms in Molecules: A Quantum Theory, Clarendon Press, Oxford,
[62] S.Y. Tong, W.M. Kang, D.H. Rosenblatt, J.G. Tobin, D.A. Shirley, Photoelectron- 1990.
diffraction analysis of the structure of c(2  2)O on Ni(0 0 1), Phys. Rev. B 27 [86] J.E. Kirsch, S. Harris, Electronic structure studies of Ni(1 0 0) surface
(1983) 4632–4636. reconstructions resulting from carbon, nitrogen, or oxygen atom adsorption,
[63] C.R. Brundle, J.Q. Broughton, in: D.A. King, D.P. Woodruff (Eds.), The Chemical Surf. Sci. 522 (2003) 125–142.
Physics of Solid Surfaces and Heterogeneous Catalysis, vol. 3A, Elsevier, [87] G. Kresse, J. Hafner, First-principles study of the adsorption of atomic hydrogen
Amsterdam, 1990 (Chapter 3). on Ni(1 1 1), Ni(1 0 0) and Ni(1 1 0), Surf. Sci. 459 (2000) 287–302.

You might also like