You are on page 1of 8

Surface Science 603 (2009) 2210–2217

Contents lists available at ScienceDirect

Surface Science
journal homepage: www.elsevier.com/locate/susc

Nickel deposition on c-Al2O3 model catalysts: An experimental


and theoretical investigation
Francois Loviat, Izabela Czekaj *, Jörg Wambach, Alexander Wokaun
General Energy Department, Paul Scherrer Institut, 5232 Villigen PSI, Switzerland

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

Article history: Recently, surface modifications on a commercial Ni/c-Al2O3 catalyst during the production of methane
Received 19 January 2009 from synthesis gas were investigated by quasi in situ X-ray photoelectron spectroscopy (XPS) [I. Czekaj,
Accepted for publication 27 April 2009 F. Loviat, F. Raimondi, J. Wambach, S. Biollaz, A. Wokaun, Appl. Catal. A: Gen. 329 (2007) 68]. The conclu-
Available online 6 May 2009
sion was that the reactivity and the observed reaction mechanisms on the different Ni particles are influ-
enced directly by both the size and the composition of the particles on the c-Al2O3 support.
Keywords: In this investigation, Ni deposition and cluster growth on model catalyst samples (10 nm thick, poly-
Nickel catalyst
crystalline c-Al2O3 on Si(100)) were investigated by XPS. Several steps in the binding energy during Ni
Alumina support
XPS
deposition indicate changes in the cluster growth. The molecular structure of the catalyst was investi-
c-Al2O3 gated using Density Functional Theory calculations (StoBe) with a cluster model and non-local functional
DFT (RPBE) approach. An Al15O40H35 cluster was selected to represent the c-Al2O3(100) surface. Ni clusters of
Methanation different size were cut from a Ni(100) surface and deposited on the Al15O40H35 cluster in order to validate
Metal deposition the deposition model determined by XPS.
Metal–support interactions Ó 2009 Elsevier B.V. All rights reserved.

1. Introduction their size but also on their shape. Thus, the surface structure of
the particles is closely related to the chemisorption properties.
The activity and selectivity of supported metal catalysts are The presence of the support was recognised to play an important
strongly influenced by the amount of metal employed, the size of role in the control of the particle morphology. The intrinsic heter-
the dispersed metal particles, the composition of the support, ogeneity of the supported model catalysts has to be taken into ac-
and metal–support interactions. In the case of supported metal cat- count to understand in detail the catalytic reactions.
alysts, attention has focused on the effect of dispersion, which in- Three different theoretical growth modes during metal deposi-
volves both a surface size effect and a relation between reactivity tion have been suggested [6–9]: ‘‘Volmer–Weber” mode, where the
and the electronic properties of various systems. Nickel-based cat- deposited metal forms clusters immediately on the support;
alysts supported on c-Al2O3, TiO2, Pt, zeolites or SiO2 [2–9] are ‘‘Frank van der Merwe” mode describes a layer-by-layer growth;
interesting examples due to their usage in numerous industrial and ‘‘Stranski–Krastanov” mode, which includes transition from
processes [2–5]. Previously, it was shown that nickel particles an initial layer-by-layer to a consecutive three-dimensional cluster
change their morphology during catalytic reactions by cluster growth at a critical layer thickness. An additional model for the
growth processes, and that part of the active clusters are lifted growth mechanism was suggested by Jacobs et al. [4], who de-
from the support due to carbon deposition and carbon whisker for- scribe a complex growth mechanism of Ni on Al2O3 when atomic
mation [1]. The industrial application of these types of catalysts layer epitaxy is applied. Determining the surface composition with
[2,10] makes the investigation of the role of metal–support interac- LEIS and XPS, the authors conclude that in the initial stage nickel
tions for the Ni particle growth and detachment during methana- adsorbs as well dispersed atoms and interacts with the strongest
tion an important issue. However the complex structure of these binding sites on the alumina support. These Ni centres are the base
catalysts demands further exploration on model catalysts. for the subsequent growth of nickel particles. Some literature data
The structure, the electronic properties and the reactivity of [3,11–13] suggest the possibility that Ni ions migrate from the sur-
supported model catalysts have been studied applying a large face into the bulk forming nickel aluminate, NiAl2O4, where nickel
number of surface science techniques [1]. It was found that the appears in NiO form. The latter is extremely stable and difficult to
electronic properties of small metal particles depend not only on reduce to metallic Ni, which is the catalytically active form. Ni0
particles grow on top of the NiO interface later [1]. Furthermore,
* Corresponding author. Tel.: +41 056 310 4464; fax: +41 056 310 2199. charge transfer processes accompanying the metal–support inter-
E-mail address: izabela.czekaj@psi.ch (I. Czekaj). actions are important additionally as they strongly influence the

0039-6028/$ - see front matter Ó 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.susc.2009.04.032
F. Loviat et al. / Surface Science 603 (2009) 2210–2217 2211

electronic structure of the Ni surface and therefore the chemical compounds. Furthermore, small particles require studies including
activity of Ni-based catalysts. Applying XPS, Sarapatka [5] studied support interactions, while in the case of large particles (with sizes
the interactions between nickel particles and the Al2O3/Al support of several nm) on the support just a (pure) metal model will be va-
as well as the charge transfer from Al2O3 to the deposited Ni. The lid for the description.
observed chemical shift of the nickel core levels are attributed to The goal of the theoretical part of this work is to define the
a transfer of electrons from anions of the oxide surface to the dis- geometry and favourable localisation of metal particles on the
persed Ni particles. However, there are still many open questions support (disregarding steps or kink sites), metal–support interac-
concerning the modification of the electronic structure of Ni-parti- tions and the role of the support in the modification of the geomet-
cles and the support (e.g. Al2O3). rical/electronic structure of metal particles and active sites. For the
Several theoretical investigations (mostly DFT calculations) of experiments, model catalysts have been chosen due to the fact that
the pure Ni system [14–16] focussed on the stability and diffusivity they (a) possess flat surfaces, which allow the application of AFM,
of surface species, possible mechanisms for the methane activation SE, etc., and (b) allow a sufficient electron tunnelling through the
and also try to explain graphite/graphene formation on the surface (10 nm thin) c-Al2O3 layer, making XPS and AES investigations
of nickel particles, which could lead to the formation of carbon possible. This ‘‘ideal” definition of the surface allows the compari-
whiskers. Numerous theoretical studies about pure and modified son of experimental data with theoretical modelling, leading to
c-Al2O3 systems were performed previously [17–19]. However to additional insights into the topology of metal particles on the
our knowledge, theoretical studies about the combined Ni/c- support. The interactions of different nickel clusters are presented
Al2O3 system are still missing in the literature. A promising study in details as a prerequisite to understanding the role of metal–
of combined systems – Mo-based catalysts, mainly Mo-methyli- support interactions in catalytic activity and stability of metal
dene over c-Al2O3 – was already achieved by Handzlik et al. [17] particles.
using a cluster model. Following the strategy outlined in this com- This paper consists of the following parts: information about
bined model, the effect of metal–support interactions, the role of the crystal structure of the material used in our theoretical calcu-
edges as well as their effect on catalytic reactivity can be studied lations is presented in Section 2. Experimental and computational
for the combined Ni/c-Al2O3 system. It is well known from the lit- details are described in Section 3. The results are presented in Sec-
erature [2] that hydrogenation processes take place on metallic Ni0 tion 4. In Section 4.1, THE results obtained from XPS experiments
sites, only. Therefore, any other Ni particles, like oxides, hydroxides are presented and the role of the deposition time with regard to
or carbides are not active for the methanation reaction. However, changes of the chemical state of surface is shown. The DFT model-
the co-existence of these inactive Ni compounds either as separate ling of Ni/Al2O3 interface is described in Section 4.2. Finally, the
clusters or even inside an otherwise catalytically active metallic conclusions are summarised in Section 5.
cluster are observed over c-Al2O3 support, as well [1]. The latter
is likely due to better structural compatibility of these kind of 2. Crystal structure of Al2O3 and nickel clusters
not-reactive nickel compounds (Ni3C, NiO or Ni(OH)2) with the
c-Al2O3 support (see Table 1) thus forming the interface of the The cubic phase [24,25] of defective alumina spinel, Al2O3, is de-
otherwise metallic nickel particles. scribed by the space group Fd-3m (No. 227) with lattice constants
In contrast to the Ni case, theoretical investigations of the me- a = b = c = 7.911 Å. The crystal unit cell of defective alumina spinel
tal–support interactions for the adsorption of Pd on c-Al2O3 have contains 56 atoms, where Al occupies two different types of posi-
been published already [21,22]. The conclusions are that metal tions – 8a, 16d, and oxygen 32e sites.
atoms prefer tetrahedral Al sites and create localised atom–surface Inside the bulk both AlO6 octahedra and AlO4 tetrahedra are
bonding. Metal–support interface formation is determined by the present with Al–O distances equal 1.94 Å and 1.78 Å, respectively.
acid properties of the cationic sites available at support. However, The bulk is built by sixfold and fourfold coordinated aluminium
these studies include only single metal adsorption on the support, and threefold and fourfold coordinated oxygen. Fig. 1 shows the
which correspond only to a very low metal coverage. structure of the c-Al2O3(100) surface. Two different aluminium
When atoms or small metal particles are deposited on a surface, sites with fourfold and fivefold coordinated aluminium, AlO4 and
they may create geometrical and electronic structures different AlO5, as well as two different threefold coordinated oxygen sites,
from their bulk. This is due to differences in the lattice constants O(3) and O0 (3), have been distinguished. It is important to note that
between metal and support. By starting from a small cluster-level, Al(4) is always located below the Al(5) centres and that the rows of
orbitals still have discrete levels. By increasing the number of AlO5 pyramids are separated from each others by rows of AlO4
atoms in the particles so that the particles reach several nm diam- tetrahedra.
eters, a band structure appears [23]. In this sense, a model system Two surface planes are predominantly detected in the diffrac-
must consider several surface scenarios, namely the presence of tion patterns of c-Al2O3, namely the (1 1 0) and (1 0 0) surfaces.
small clusters and large agglomerates, as well as different surface Following some previous studies of alumina by Handzlik et al.

Table 1
Comparison of structure parameters.
Compound M and Y definition M–M distance (Å) Differences in respect to Ni/c-Al2O3 (%) M-Y distance ((Å) Crystal structure
(a) NiYx compounds and support [24–26].
Ni Nimetal 2.49 –/10.4 – Cubic
NiO Ni, O 2.95 +18.5/+6.1 2.08 Cubic
Ni(OH)2 Ni, OH 2.71 +8.8/2.5 2.01, 2.39 Trigonal
NiAl2O4 Ni or Al, O 2.85 +14.5/+2.5 1.83, 1.96 Cubic
c-Al2O3 Al, O 2.78 +11.6/– 1.78, 1.93 Cubic
Compound M–M lateral distance (Å) M–M distance (Å) Differences in respect to c-Al2O3 (%) M–O/O–O distance (Å)
(b) Si, SiO2 and c-Al2O3 compounds with cubic structure [32,33]
Si 3.80/5.38 2.33/3.80 28.9/39.2 –
SiO2 7.15/5.05 3.09/5.92 5.8/5.3 1.55/2.53
c-Al2O3 7.91/2.78 3.28/6.25 – 1.78, 1.94/2.78
2212 F. Loviat et al. / Surface Science 603 (2009) 2210–2217

Fig. 1. Crystal structure of the c-Al2O3(1 0 0) surface: (a) top view and (b) side view. Al (grey spheres) and O (red spheres). Note: the radii of the spheres do not correspond to
the real radii. This is made for better visibility of active centres. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

[17], the (1 0 0) surface was chosen for our studies. Three different energy (BE) scale was adjusted by setting the main C1s peak to
distances are observed between the aluminium centres on the 284.5 eV. Quantification of the XP spectra was carried out using
(1 0 0) surface of alumina: (i) 3.28 Å between Al(4)–Al(5) centres, CasaXPS [27] and applying the transmission function of the elec-
(ii) 2.79 Å between Al(5)–Al(5) centres in the AlO5 row and (iii) tron energy analyzer and cross-sections calculated by Scofield.
5.59 Å between Al(4)–Al(4) centres. The surface exhibits specific The spectra were deconvoluted by applying Gaussian–Lorentzian
holes with a distance of 2.90 Å and 2.79 Å between the O(3) cen- line-shapes with a Shirley-type background.
tres. All these surface features are important, since sufficient com- Attached via a distribution chamber is a preparation chamber,
patibility of the (1 0 0) surface with the Ni particles is necessary in which is equipped with a home-made Metal Evaporator (ME).
order to obtain stable Ni-clusters on the c-Al2O3 support. The latter can evaporate thermally two metals (Ni and Pd), stored
Concerning nickel particles, the cubic phase [26] of metallic and heated independently in separate chambers. The evaporation
nickel is described by the space group Fm-3m (No. 225) with lat- rate is determined using a quartz crystal microbalance (QCM).
tice constants a = b = c = 3.5239 Å. The crystal unit cell of nickel For a detailed description of the set-up see [28]. The systems base
contains four atoms, where Ni occupies 4a positions. The distance pressure is below 1  1010 Torr and during metal evaporation the
between Ni–Ni is equal to 2.49 Å. Each Ni atoms has six neighbours pressure in the preparation chamber was always below
at the surface as well as three in the sub-surface. 1  109 Torr. XPS characterisation was performed before and after
any Ni evaporation and without exposure of the catalyst to air.
3. Technical details The c-Al2O3/Si model substrates were prepared in the Labora-
tory for Neutron Scattering at the Paul Scherrer Institut by M. Hor-
3.1. Experimental details isberger. Thick films of aluminium oxide (10 nm) were deposited
on a Si(1 0 0) wafers by argon sputtering of an aluminium target
The XPS measurements were carried out in a VG ESCALAB 220i in a Leybold Z600 DC magnetron sputtering apparatus. The partial
XL (VG Scientific) electron spectrometer. XPS measurements were pressures of argon and oxygen during the deposition were
made using non-monochromatic Mg Ka (1253.6 eV) radiation. 3.2  103 mbar. The aluminium oxide deposition rate of
While measuring, the chamber pressure was always lower than 0.25 nm/s was determined with a QCM. The structure, morphology
1  109 Torr. The electron energy analyzer was operated in the and chemical composition of c-Al2O3/Si substrates were investi-
constant pass energy mode with a pass energy of 50 eV for survey gated by the combined application of XPS, AES, SE, and AFM.
scans and 20 eV for detailed scans of selected spectral regions. The Merckling et al. [29,30] reported that c-Al2O3 grows epitaxially
conductivity of the samples was high enough to keep sample on Si(0 0 1) in spite of the large apparent large lattice mismatch
charging below a few eV. For charge compensation, the binding (>30%) between c-Al2O3 and Si. This is possible when c-Al2O3
F. Loviat et al. / Surface Science 603 (2009) 2210–2217 2213

grows with its oxide lattice being rotated by 45° with respect to the appropriate broken Al–O bonds. The thus obtained cluster
silicon lattice. This reduces the resultant lattice mismatch to 3.5%. Al15O40H35 has been selected for the investigation (Fig. 2). The
Further, their studies showed that Al2O3 films grow pseudomor- Al15O40H35 cluster consists of all characteristic centres observed
phically on Si(1 0 0) up to thickness of 2 nm, and that for higher on the Al2O3 (1 0 0) surface (for comparison see Fig. 1). For our
thicknesses, a cubic-to-hexagonal surface phase transition occurs. investigations we have chosen the c-Al2O3 surface with a (1 0 0)
Another possibility of overcoming the above-mentioned lattice termination based on previous investigations [17,18] as well as
mismatch is described by Jung et al. [31]. The authors investigated experimental results [28], which show in details the structure of
the hetero-epitaxial growth of c-Al2O3 films on Si substrates. a model catalyst with specific ‘‘valley-hill” rowed structure.
Atomically smooth single-crystalline epitaxial c-Al2O3 layers were The clusters representing nickel (Ni2, Ni7, and Ni9 cluster) were
successfully grown on Si(1 1 1) at growth temperatures between cut from the ideal bulk structure without saturation by hydrogen
650 °C and 750 °C. When growing Al2O3 films at temperatures be- or point charges to realistically simulate small surface particles
low 650 °C, initially a SiO2 layer is being formed on Si, on which (see Fig. 3a).
polycrystalline Al2O3 films are growing consecutively. This SiO2 Fig. 3b–d shows the results of a geometric optimization of the
interlayer thus forms a ‘‘matching” interface between the silicon chosen Ni clusters on c-Al2O3 cluster representing the Nix/
and the Al2O3 helping to overcome the lattice mismatch (see Table Al15O40H35 system (x = 2,7,9) used for the theoretical modelling.
1b). The electronic structure of all clusters was calculated by ab ini-
Introduced into vacuum, the samples were cleaned by an initial tio density functional theory (DFT) methods (program code StoBe
flash to 573 K and a subsequent sputtering (at 1 keV, 10 lA ion cur- [34]) using the non-local generalized gradient corrected function-
rent) in 105 mbar of Ar, annealed in 5  106 mbar oxygen, and fi- als according to Perdew, Burke, and Ernzerhof (RPBE) [35,36], in
nally cooled down maintaining the oxygen pressure until the order to account for electron exchange and correlation. All Kohn–
substrate temperature decreased below 373 K. After two cleaning Sham orbitals are represented by linear combinations of atomic
cycles, the surface was considered as free of any external contam- orbitals (LCAO’s) using extended basis sets of contracted Gaussians
ination. The images of the c-Al2O3/Si substrates obtained by AFM from atom optimizations [37,38]. Detailed analyses of the elec-
did not allow distinguishing any structure on the surface, indicat- tronic structure in the clusters are carried out using Mulliken pop-
ing therefore a very smooth morphology. SE and XPS depth profil- ulations [39] and Mayer bond order indices [40,41]. During
ing suggested an additional layer of SiO2 with a few nm thickness relaxation, the nickel atoms in the supported clusters had allowed
at the interface between Si and Al2O3. This SiO2 interface helps to move in 3D space.
overcoming the above-mentioned lattice mismatch.
The Ni/c-Al2O3/Si model catalysts were prepared by depositing 4. Results
various amount of Ni from the metal evaporator onto the c-Al2O3/
Si substrates at substrate temperatures of 623 K. The growth 4.1. XPS results of Ni deposition on Al2O3: the role of deposition time
mechanism of nickel over c-Al2O3 has been investigated by a com- with regard to changes in the chemical state of surface
bination of XPS and AFM analysis.
Fig. 4 gives the Ni 2p3/2/Al 2p ratio versus the deposition (ex-
3.2. Computational details pressed in seconds) of nickel onto the c-Al2O3 support. Dots repre-
sent the experimentally obtained data. The deposition was
In our studies the Ni and c-Al2O3 support surfaces are modelled performed with a deposition rate corresponding (theoretically) to
by clusters of different size and geometry, which reflect local sec- 0.75 ML/min. The red line symbolizes the theoretically expected
tions of the ideal surface. Clusters are constructed by a successive Ni 2p3/2/Al 2p ratio for a pure layer-by-layer growth. For this esti-
addition of neighbouring shells to a small core structure. This pro- mation, we presumed that this (theoretical) Ni monolayer would
cedure is repeated until convergence of the electronic properties of have the dense geometry of a Ni(1 0 0) monolayer on the c-Al2O3.
the clusters is obtained. Up to a deposition time t  45 s, our experimental data follow al-
For the c-Al2O3 cluster representing the support surface, formal most the expected layer-by-layer growth line. This suggests that
valence charge saturation (based on Al3+ and O2) is imposed and in the initial phase Ni does not form clusters immediately, but 1D
cluster neutrality with respect to the surface is achieved by satu- (or possibly 2D) agglomerates. In a second stage, the Ni 2p3/2/Al
rating the peripheral oxygen centres with hydrogen atoms placed 2s ratio is below the ‘‘layer-by-layer” line. This makes it likely that
at the standard OH distance, ROH = 0.97 Å, in the direction of the in this phase 3D Ni clusters are growing on the surface. We

Fig. 2. Top view (a) and side view (b) of the Al15O40H35 model cluster describing the c-Al2O3(100) surface.
2214 F. Loviat et al. / Surface Science 603 (2009) 2210–2217

Fig. 3. Models of Ni clusters: (a) cut from Ni(100) surface; as well as deposited + relaxed on the c-Al2O3(100) surface represented by Al15O40H35 cluster: (b) Ni2 cluster, (c) Ni7
cluster, and (d) Ni9 cluster.

Fig. 4. Changes of Ni 2p3/2/Al 2p ratio during deposition of nickel onto the Al2O3
support. The red line corresponds to the theoretical prediction of the Ni 2p3/2/Al 2p
ratio for a layer-by-layer growth. The theoretical depositions times for 0.5 ML and
1.0 ML of Ni (layer-by-layer growth) are indicated, too. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of
this article.)

conclude that Ni deposition on c-Al2O3 support follows a kind of


‘‘modified” Stranski–Krastanov growth mode under the applied
experimental conditions, which is in accordance to the findings of
Jacobs et al. [4].
A support for the suggested change in the growth mode can be
found by XPS. In Fig. 5, the spectra of the Ni 2p region for several Ni
deposition times are displayed, and in Fig. 6 the summarised
Fig. 5. XP spectra of the Ni 2p region. The different traces correspond to various
changes of Ni 2p3/2 binding energies during deposition are given. deposition times: (a) 10 s; (b) 50 s; (c) 100 s; (d) 150 s; (e) 210 s; and (f) 380 s. The
The change of the binding energy can be separated into three parts, two bottom traces show spectra of bulk Ni and NiO, and are included as reference.
as indicated by the two vertical lines. In the initial part, the first
found BE value (854.0 eV) is close to published Ni oxide (NiO) BE
data [12,13,42,43]. This can be interpreted as single Ni atoms being Interestingly, the time of the transition from the first to the sec-
deposited and chemically modified by being bonded strongly to ond part (45 s) is comparable with the time, where the deviation
the oxygen of the support. The further initial part up to about 45 of the experimentally derived Ni 2p3/2/Al 2p ratio from the ‘‘layer-
s is characterised by a fast decrease of the BE down to circa by-layer” ratio starts, as shown in Fig. 4 (see the vertical line). This
853.3 eV. In the second part, between 50 s and about 150 s deposi- makes it likely that in the second stage additional Ni atoms are
tion times, the slope of the BE change is decreased and the BE value added to the already existing ones forming 3D Ni agglomerates.
is lowered to 853.0 eV. The third part, i.e. deposition times more Finally, the third part indicates that continual Ni deposition and
than 170 s, is characterised by a slow approach to 852.75 eV (ob- cluster growth processes leads to extended three-dimensional Ni
tained for depositions times >350 s (not shown)), the value of clusters, exhibiting a band structure typical for bulk Ni. This behav-
metallic Ni as indicated by the horizontal dash-dotted line. iour is in agreement with a model described by Wertheim [44]. A
F. Loviat et al. / Surface Science 603 (2009) 2210–2217 2215

clusters Al15O40H35 and Nix/Al15O40H35 (x = 2,7,9). The data are bas-


ing on optimization calculations described in Section 3.2. Geomet-
ric structures of all ideal clusters are presented in Figs. 2 and 3a.
The charge of aluminium centres on the (1 0 0) surface
increases in the order q[Al(5)] < q[Al(6)] < q[Al(4)]. The Al(5) centre
P
is stronger bound to the surface ( b.o. = 3.1) than Al(4)
P
( b.o. = 2.3), but is not fully coordinated. Therefore, the Al(5) cen-
tre can be easily oxidized or hydratized.
The two types of surface three-coordinated oxygen centres, O(3)
P
and O0 (3), are similarly bound to the surface ( b.o.  2.0), but
0
show a different ionic character. The O (3) centres, connected with
five and four coordinated aluminium centres, have higher charge
and are supposed to be more reactive than O(3).
The Al15O40H35 cluster was found to be a good representation
for the electronic states of both oxygen O0 (3) and O(3) as well as
three types of aluminium centres, Al(4), Al(5) and Al(6), present
on the Al2O3(100) surface. Taking into account computational time
(including optimization and metal deposition at the surface) as
Fig. 6. Changes of XPS binding energies of the Ni 2p3/2 peak during deposition of
nickel on c-Al2O3. (a) single atoms, (b) 1D (or 2D) growth, and (c) 3D growth. well as accuracy of the electronic structure, the Al15O40H35 cluster
seems to be a good enough compromise of cluster convergence cri-
teria and therefore was chosen for all further considerations in me-
comparable behaviour and BE shifts as function of the aggregate
tal deposition studies.
size have been measured by XPS for different metals deposited
The stabilization energy of nickel particles deposition on
on alumina and other substrates [45,46].
Al15O40H35 cluster is calculated as:
This behaviour can be explained with the drastic changes of the
electronic structure of the deposits as the cluster size increases. Edep ðM n =Al2 O3 Þ ¼ Etot ðMn =Al15 O40 H35 Þ  Etot ðAl15 O40 H35 Þ
The size of the cluster determine the final state in the photoemis-
 nEtot ðMÞ;
sion process, because after the photoemission the positive charge
left on the aggregate can be less screened respectively delocalised where Etot(Mn/Al15O40H35) is the total energy of the metal deposited
when compared with an extended metallic system. This results in a at the Al2O3 surface. Etot(Al15O40H35) and Etot(M) are the total ener-
shift of the XP spectrum and corresponds to the Coulomb energy of gies of pure Al15O40H35 and the M atom, respectively. In all cases,
the localised charge. It was found that the shift is proportional to the first layer of nickel deposits on the Al2O3(100) in positions clo-
the reciprocal particle diameter [44,47]. Consequently, the binding ser to O(3) centres with different stabilization energy per Ni atom
energies of small metal aggregates should be found at higher (for Ni2: 0.82 eV, for Ni7: 0.90 eV and for Ni9: 0.71 eV). The
values as is indeed observed in many cases. determined stabilization energies of the different Ni clusters are
in the range obtained for the adsorption of different metals, mainly
4.2. DFT modelling of Ni/Al2O3 interface Pd, on c-Al2O3 [19,20]. The deposited nickel influences the elec-
tronic structure of c-Al2O3 and is strongly bound to octahedral
Table 2 presents information about atomic charges (Mulliken O(3), as can be seen from the bond order analysis shown in Table
analysis), distances (Å) and bond orders (Mayer bond analysis) of 2. Interesting results are obtained for the aluminium ions Al(4),

Table 2
Results obtained by DFT calculations summarising atomic charges, distances (Å) and bond orders (Mayer bond analysis) of (a) the Al15O40H35 cluster, (b) the Ni2 cluster at
Al15O40H35 cluster, (c) the Ni7 cluster at Al15O40H35 cluster, and (d) the Ni9 cluster at Al15O40H35 cluster.

(a) (b) (c) (d)


Cluster Al15O40H35 Ni2/Al15O40H35 Ni7/Al15O40H35 Ni9/Al15O40H35
Centre Charge [eV]
Al(4) +1.58 +1.87 +2.20 +2.16
Al(5) +0.99 +0.91 +0.80 +1.01
Al(6) +1.11 +1.13 +1.13 +1.15
O(3) 0.68 0.82 0.89 0.92
O0 (3) 0.75 0.78 0.84 0.89
Ni1st – +0.28 +0.38 +0.35
Bond Distance [Å] Bond Order
O(3)–Al(5) 1.94 0.73 0.61 0.66 0.66
O(3)–Al(5) 1.94 0.66 0.58 0.45 0.51
O(3)–Al(6) 1.94 0.62 0.52 0.54 0.53
O(3)–Ni (see diff. Nix) – 0.21 (2.16 Å) 0.17 (2.04 Å) 0.15 (2.09 Å)
P
b.o. 2.01 1.92 1.82 1.85
O’(3)–Al(5) 1.94 0.66 0.61 0.58 0.57
O’(3)–Al(5) 1.94 0.66 0.61 0.58 0.57
O’(3)–Al(4) 1.78 0.63 0.43 0.48 0.38
O’(3)–Ni (see diff. Nix) – 0.08 (2.57 Å) 0.01 (2.96 Å) 0.004 (3.10 Å)
P
b.o. 1.95 1.73 1.65 1.50
Ni1st–Ni1st (see diff. Nix) – 0.34 (3.53 Å) 0.08 (4.52 Å) 0.09 (4.77 Å)
EHOMO/LUMO (eV) 4.51/4.46 3.80/3.70 4.32/4.14 4.12/4.09
ENi_deposition/cluster (eV) – 1.64 6.27 6.41
ENi_deposition/Niatom (eV) 0.82 0.90 0.71
2216 F. Loviat et al. / Surface Science 603 (2009) 2210–2217

Fig. 7. Topography of first (interface) Ni atoms (green spheres) for the relaxed Ni9 cluster from an Al15O40H35 cluster: (a) all Ni interface atoms included (hNi  0.4 ML); (b)
only Ni atoms in hole positions included (hNi  0.2 ML). Black dots symbolize the position of Ni atoms in an ideal Ni(1 0 0) monolayer. (For interpretation of the references to
colour in this figure legend, the reader is referred to the web version of this article.)

for which the positive charge is found to be increased after a higher of a very low number of metal atoms on c-Al2O3, only the local sur-
Ni deposition. Together with the changes of the charges of the oxy- face structure is influenced, in particular the neighbouring centres
gen centres O(3) and O0 (3) as well as of the neighbouring nickel of Ni, such as: Al(4), Al(5), O(3) and O0 (3). From the results pre-
atoms, Ni1st, an increasing polarity at surface has to be stated. This sented in Table 2 for three different nickel clusters, we can derive
is in agreement with a decrease of the oxygen bond orders, which that the O(3) centres are mostly influenced creating strong bonds
describe a covalent character of the system, also. As a consequence, with the first (interface layer) nickel atoms. This finding is in good
due to the deposition of metallic particles the surface of Ni/Al2O3 accordance with our experimental results (see Fig. 6), where we
system becomes more ionic. observe the existence of electronically strongly altered Ni being ad-
The largest investigated nickel cluster (Ni9) creates many inter- sorbed on c-Al2O3 in the first part of the deposition. As mentioned
esting structures at the support surface. Fig. 7 shows the interface above, we presume that initially single Ni atoms are deposited
atoms of the Ni9 cluster after relaxation (the additional atoms have (first 25 s of deposition), and that further Ni deposition (up to
been removed; the entire cluster can be seen in Fig. 3d). The figure 120 s) leads to the formation of either 1D Ni chains or 2D Ni
was created by multiplying the calculated Ni/Al2O3 cluster in x and agglomerates.
y directions. The interface nickel atoms are important due to the A further important question is the suggested existence of nick-
fact that the direct Ni–Al2O3 interaction can be directly responsible el aluminate in the interface and its influence on cluster stability at
for the stability of growing metal clusters on alumina. The black alumina. Studies going in that direction are presently running in
dots, added for comparison with the relaxed structure, symbolize our laboratory and will be presented elsewhere [48].
the mismatch of an ‘‘ideal” Ni(1 0 0) monolayer with the c-Al2O3
surface. Green spheres give the positions of Ni atoms with coverage
h  0.75 ML (Fig. 7a) and h  0.25 ML (Fig. 7b). As visible in Fig. 7a, 5. Conclusions
strong vertical and lateral re-arrangements of the interface Ni
atoms in the Ni9 clusters with respect to the position on the Nickel is stabilized on the c-Al2O3 surface influencing the elec-
Ni(1 0 0) surface are indicated by DFT. Part of the nickel atoms tronic properties of the newly formed surface. Our DFT data
seem to prefer the ‘‘valley regions” between AlO5 rows and close suggest that at low coverages (60.2 ML) Ni prefers being
to O(3) centres. This is visible more clearly in Fig. 7b. The nickel localised in AlO4 tetrahedra between rows of AlO5. The DFT
centres (‘‘Ni1st”) in that row increase their respective distance from results correspond well with the experimental data of the initial
3.5 to 4.8Å, and approach almost the O(3)–O(3) distance along stage of Ni deposition, where the formation of a partial Ni mono-
the AlO5 rows (= x-axis), which is about 5.6 Å. The neighbouring layer is suggested. Additional Ni deposition leads first to
nickel atoms (‘‘Ni2nd”) are three displaced in three dimensions: ver- non-metallic three-dimensional agglomerates, which by continual
tically (along the x-axis) and additionally by a slight relocation deposition are finally transferred to metallic Ni clusters on
along the y-axis. The lateral movement of the ‘‘Ni2nd” atoms is the surface. For Ni7 and Ni9 clusters, the initially deposited Ni
more prominent, most likely due to the repulsion of nickel and atoms, which represent the interface nickel atoms (‘‘Ni1st”), are
the aluminium Al(5) centres, and results in a lateral distance be- bound strongly to the oxygen of the support and are located in
tween ‘‘Ni2nd”–‘‘Ni1st” layers of about 0.54 Å. positions closer to O(3) centres as well as between rows of
The Al(6) centres, being located in subsurface region, are not AlO5 with adsorption energy, which varies with the size of the
influenced by the Ni deposition. This suggests that, after deposition cluster.
F. Loviat et al. / Surface Science 603 (2009) 2210–2217 2217

Acknowledgements [21] A.M. Marquez, J.F. Sanz, Appl. Surf. Sci. 238 (2004) 82.
[22] M.C. Valero, P. Raybaud, P. Sautet, J. Phys. Chem. B 110 (2006) 1759.
[23] H.J. Freund, M. Bäumer, H. Kuhlenbeck, Adv. Catal. 45 (2000) 333.
The calculations were done partially using the unix farm as well [24] R.-S. Zhou, R.L. Snyder, Acta Crystallogr. B 47 (1991) 617.
as the linux farm at the Paul Scherrer Institut. We would like to ex- [25] E.J.W. Verwey, Z. Kristallograp. 91 (1935) 317.
[26] R.W.G. Wyckoff, Crystal Structures, vol. I, Interscience Publishers John Wiley &
press our thanks to M. Horisberger for preparation of the c-alu-
Sons Inc., New York, London, Sydney, 1965.
mina model support as well as to R. Kötz for Spectroscopic [27] N. Fairley, CasaXPS Software, Version 2.3.12, 2006.
Ellipsometry. [28] F. Loviat, Photoassisted activation of methane with a xenon excimer lamp, PhD
Thesisß ETH 2008.
[29] C. Merckling, M. El-Kazz, V. Favre-Nicolin, M. Gendry, Y. Robach, G. Grene, G.
References Hollinger, Thin Solid Films 515 (2007) 6479.
[30] C. Merckling, M. El-Kazzi, G. Delhaye, M. Gendry, G. Saint-Girons, G. Hollinger,
[1] I. Czekaj, F. Loviat, F. Raimondi, J. Wambach, S. Biollaz, A. Wokaun, Appl. Catal. L. Largeau, G. Patriarche, J Appl. Phys. 102 (2007) 024101.
A: Gen. 329 (2007) 68. [31] Y.-C. Jung, H. Miura, M. Ishida, J. Crystal Growth 201/202 (1999) 648.
[2] G. Ertl, H. Knözinger, J. Weitkamp, Handbook of Heterogeneous Catalysis, [32] D.N. Batchelder, R.O. Simmons, J. Chem. Phys. 41 (1964) 2324.
Wiley Company, Weinheim, 1997. [33] W.W. Schmahl, I.P. Swainson, M.T. Dove, A. Graeme-Barber, Z. Kristallogr. 201
[3] D. Nazimek, A. Machocki, T. Borowiecki, Adsorption Sci. Technol. 16 (1998) (1992) 125.
747. [34] The program package StoBe is a modified version of the DFT-LCGTO program
[4] J.P. Jacobs, L.P. Lindfors, J.G.H. Reintjes, O. Jylhä, H.H. Brongersma, Catal. Lett. package DeMon, originally developed by A. St.-Amant and D. Salahub
25 (1994) 315. (University of Montreal), with extensions by L.G.M. Pettersson and K.
[5] T.J. Sarapatka, Chem. Phys. Lett. 212 (1993) 37. Hermann.
[6] C.S. Shern, J.S. Tsay, T. Fu, Appl. Surf. Sci. 92 (1996) 74. [35] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865.
[7] R.E. Tanner, I. Goldfarb, M.R. Castell, G.A.D. Briggs, Surf. Sci. 486 (2001) 167. [36] B. Hammer, L.B. Hansen, J.K. Nørskov, Phys. Rev. B 59 (1999) 7413.
[8] J.D. Carey, L.L. Ong, S.R.P. Silva, Nanotechnology 14 (2003) 1223. [37] J.K. Labanowski, J.W. Anzelm (Eds.), Density Functional Methods in Chemistry,
[9] M.A. Karolewski, Surf. Sci. 517 (2002) 138. Springer-Verlag, New York, 1991.
[10] M.C. Seemann, T.J. Schildhauer, S.M.A. Biollaz, S. Stucki, A. Wokaun, Appl. Catal. [38] N. Godbout, D.R. Salahub, J. Andzelm, E. Wimmer, Can. J. Phys. 70 (1992) 560.
A: Gen. 313 (2006) 14. [39] R.S. Mulliken, J. Chem. Phys. 23 (1955) 1833. 1841, 2388, 2343.
[11] R. Lamber, G. Schulz-Ekloff, Surf. Sci. 258 (1991) 107. [40] I. Mayer, Chem. Phys. Lett. 97 (1983) 270.
[12] K. Shih, J.O. Leckie, J. Euro. Ceramic Soc. 27 (2007) 91. [41] I. Mayer, J. Mol. Struct. (Theochem) 34 (1987) 81.
[13] R. Lamber, G. Schulz-Ekloff, Surf. Sci. 258 (1991) A595. [42] J.F. Moulder, W.F. Stickle, P.E. Sobol, K.D. Bomben, J. Chastain (Eds.), Handbook
[14] J. Sehested, J.A.P. Gelten, I.N. Remediakis, H. Bengaard, J.K. Nørskov, J. Catal. of X-ray Photoelectron Spectroscopy, Perkin–Elmer Corporation, Physical
223 (2004) 432. Electronics Division, Eden Prairie, MN, USA, 1992.
[15] H.S. Bengaard, J. Nørskov, J. Sehested, B.S. Clausen, L.P. Nielsen, A.M. [43] C.D. Wagner, A.V. Naumkin, A. Kraut-Vass, J.W. Allison, Cedric J. Powell, J.R.
Molenbroek, J.R. Rostrup-Nielsen, J. Catal. 209 (2002) 365. Rumble Jr., NIST X-ray Photoelectron Spectroscopy Database, NIST Standard
[16] F. Abild-Pedersen, O. Lytken, J. Engbaek, G. Nielsen, I. Chorkendorff, J.K. Reference Database 20, Version 3.4 (Web Version), National Institute of
Nørskov, Surf. Sci. 590 (2005) 127. Standards and Technology, Gaithersburg, USA, 2006.
[17] J. Handzlik, J. Ogonowski, R. Tokarz-Sobieraj, Catal. Today 101 (2005) 163. [44] G.K. Wertheim, Z. Phys. B 66 (1987) 53.
[18] M. Digne, P. Sautet, P. Raybaud, P. Euzen, H. Toulhoat, J. Catal. 226 (2004) 54. [45] M. Bäumer, H.-J. Freund, Prog. Surf. Sci. 61 (1999) 127. and literature therein.
[19] A. Ionescu, A. Allouche, J.P. Aycard, M. Rajzmann, F. Hutschka, J. Phys. Chem. B [46] D.-Q. Yang, E. Sacher, Appl. Surf. Sci. 195 (2002) 187.
106 (2002) 9359. [47] G.K. Wertheim, Z. Phys. D 12 (1989) 319.
[20] C. Arrouvel, M. Breysse, H. Toulhoat, P. Raybaud, J. Catal. 232 (2005) 161. [48] I. Czekaj, F. Loviat, J. Wambach, A. Wokaun, Chimia 63 (2009) 193.

You might also like