You are on page 1of 7

ARTICLE

pubs.acs.org/JPCC

Structural Defects in W-Doped TiO2 (101) Anatase Surface: Density


Functional Study
Antonio M. Marquez,* Jose J. Plata, Yanaris Ortega, and Javier Fdez. Sanz
Departamento de Química Física, Universidad de Sevilla, Facultad de Química, 41012 Sevilla, Spain
bS Supporting Information
ABSTRACT: In this Article, the structural and electronic
properties of the W-doped anatase (101) surface are investi-
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

gated by first-principles density functional theory calculations.


Downloaded via INDIAN INST OF TECH MANDI on November 23, 2023 at 09:51:47 (UTC).

Several surface and subsurface substitutional positions are


examined as well as the interaction of the W-dopant atoms
with structural defects: cation vacant sites and additional oxygen
atoms that are required to compensate the extra charge of the
W6+ cations. It is found that the preferred configurations are
those on which one W6+ cation and one Ti vacant site are first
cationic neighbors with simultaneous formation of a wolframyl
entity. The main mechanism of system stabilization is found to
be based on the formation of wolframyl species that result from
the close proximity, as first cationic neighbors of the W-dopant atom and the Ti vacant site. A second factor for system stabilization
seems to be the separation of W6+ cations to reduce the energetic cost of the structural distortions introduced by the doping process.
Results from molecular dynamics calculations indicate that W6+ cations have a 5 + 1 coordination with two W O distances at 1.8 to
2.0 and 2.5 to 2.6 Å. All of these structural results are used to understand the experimental information available for W Ti
nanostructured oxides. The modifications introduced in the electronic structure of the anatase (101) surface by the doping process
are discussed and rationalized. A comparative analysis of the density of states of doped and undoped slab models of this surface and a
Bader charge analysis will be used to understand the electronic redistribution that takes place around the impurity atoms.

1. INTRODUCTION examining its performance in the photocatalytic degradation of


Titanium dioxide (TiO2) is a material that has received methyl orange. However, in their procedure, the W-dopant was
considerable attention in the literature over the years because incorporated to only ∼4 mol %, and the limited effect of doping in
of its versatility and multitude of its applications in catalysis, the photocatalytic activity of the solid was related to the surface
photocatalysis, solar cells, waste remediation, and biocompatible presence of W6+ acid centers. A more detailed account of the
materials.1 9 In particular, nanocrystalline TiO2 has been ex- structural and electronic effects of W doping into the anatase
tensively studied because of its notable catalytic and photocata- structure of nanosized materials is the work of Fernandez-García
lytic activity. Various experimental approaches are used to scale et al.17,18 They have synthesized nanostructured Ti W mixed
the TiO2 particle size down to the nanometer range, resulting in metal oxides using a microemulsion preparation technique. The
an increased surface area that goes parallel with an enhancement structural and electronic properties of the materials resulting after
of the chemical and photochemical activity. Of the different calcination were characterized by X-ray diffraction, Raman, EX-
titania polymorphs, anatase appears to be the most important for AFS, XANES, and UV vis spectroscopies. From their data, these
these applications because the majority of nanostructured mate- authors concluded that no phase segregation appears up to a W
rials displays this specific structure,10,11 with the (101) face content of ∼20 atom %, with cation vacancies being the main
dominating >94% of the crystal surface.12 An obvious problem mechanism to achieve charge compensation in the nanostructured
hindering the photochemical applications of TiO2 is its poor mixed W Ti oxides. The Raman spectra data indicate the
visible light response due to its wide band gap (∼3.2 eV on the presence of wolframyl species (WdO) that, however, were not
anatase polymorph). Different approaches have been pursued to detected by EXAFS spectroscopy. Besides, two kinds of W O
narrow this band gap and increase visible light absorption. One of distances, distinctive of the anatase-type structure, are detected,
the most common methods is the introduction of impurities in indicating that some distortion of the anatase network is required
the TiO2 structure, either nonmetals or metals.13,14 to accommodate the W cations. The analysis of the XANES and
Nanosized W-doped TiO2 photocatalysts have been synthe-
sized and characterized by a few research teams.15 18 Tian et al.16 Received: April 7, 2011
have used a simple hydrothermal process for the synthesis and Revised: July 26, 2011
evaluated the photocatalytic activity of the prepared material by Published: July 26, 2011

r 2011 American Chemical Society 16970 dx.doi.org/10.1021/jp203223f | J. Phys. Chem. C 2011, 115, 16970–16976
The Journal of Physical Chemistry C ARTICLE

UV vis results seems to conclude that electronically Ti and W are


in +4 and +6 oxidation states, respectively, although some variation
of the W d-population is detected. The second main feature
obtained from the UV vis data is a moderate (∼0.2 eV) reduction
in the band gap of these binary solids with respect to that of bare
anatase that is related to an enhanced photocatalytic activity upon
sunlight excitation in the gas-phase photo-oxidation of toluene.
These authors have also conducted some LDA-density functional
theory (DFT) calculations in bulk models of the anatase structure
considering that W dopants are located in Ti sites and that charge
compensation occurs exclusively by the presence of Ti cation
vacancies. These calculations offered an explanation for the band
gap reduction in terms of the presence of W 5d states at the low-
energy edge of the conduction band but failed to account for the
structure distortions experimentally observed and also did not
reproduce the presence of WdO entities detected in Raman
spectroscopy. Other theoretical studies have been focused on the Figure 1. One 2 supercell model of a TiO2 (101) anatase surface with
effect of titania W-doping on the catalytic (thermal) oxidation of three O Ti O layers. Atom labeling for the substitutional W positions:
CO,19 electrical conductivity properties,20,21 and effects of N/W- 5s, pentacoordinated surface site; 6s, hexacoordinated surface site; 6u,
codoping22,23 in the electronic structure of bulk TiO2. However, subsurface hexacoordinated (up) site; and 6d, hexacoordinated subsur-
none of these works has investigated, to any extent, the structural face (down) site. Atom colors: Ti, gray; O, red.
implications on the presence of Ti vacancies and WdO entities
evidenced by the experiment,18 and only one19 has taken into Iterative relaxation of the atomic positions was stopped when the
account the presence of surface wolframyl species but in the (110) forces on the atoms were <0.01 eV/Å.
surface of the rutile polymorph. The anatase bulk lattice parameters are taken from previous
In the present work, we use DFT calculations to address the work33 in which the same theoretical approximations of this work
issue of the interaction between the W-dopant and the structural were used. The resulting parameters are a = 3.789 Å and c = 9.486 Å,
charge-compensating point defect structures: Ti-vacancies and which are within 0.2% of the experimental values (a = 3.782 Å
wolframyl species, experimentally manifested. In particular, we and c = 9.502 Å).34 The anatase (101) surface was modeled with
focus on the anatase (101) surface because, as previously different surface supercells to simulate various dopant and defect
indicated, anatase is the main structural polymorph found in concentrations. The first model (model I, Figure 1) is a supercell
nanostructured TiO2 materials, and the (101) surface is the containing 72 atoms, of dimensions 1  2 in the [101] and [010]
thermodynamically most stable crystal face. Therefore, its struc- directions, respectively. In the [101] direction, the slab was three
ture is expected to play a fundamental role in any chemical or TiO2 trilayers thick (nine atomic layers), with a vacuum of 9 Å
photochemical process that may take place on it. Moreover, the separating the slab images. The second model (model II) is
activity of anatase in photocatalytic reactions is found to be much obtained by doubling the previous one along the [101] direction.
higher than that of the rutile polymorph (see, e.g., refs 10, 24, and Therefore, it is a 2  2 supercell three TiO2 trilayers thick.
25 and references therein). According to their size, structural calculations were performed in
After briefly presenting the necessary computational details model-I-derived structures using a Monkhorst Pack set of 2 
about the models and the methodology used, we will start by 2  1 k points, whereas a 1  2  1Monkhorst Pack set was used
examining the energetics and geometries of surface defects on for model-II-derived configurations. In all cases, the atoms in the
W-doped anatase TiO2. Experimental facts, like the observation bottom layer were kept fixed to their bulk positions during
by EXAFS spectroscopy of a single local arrangement around the geometry optimizations to simulate the presence of the bulk
W-dopant, the formation of surface WdO species, and the underneath. Density of states (DOS) calculations were performed
interaction among W-dopant atoms, Ti-vacant sites, and surface using a denser 4  4  2 Monkhorst Pack set of k points.
wolframyl species will be discussed and rationalized. In a later Substitutional W doping was modeled by replacing one or two
section, the implications on the electronic structure of these Ti per supercell with W atoms. This results in models having
materials and the electronic redistribution that take place around either ∼8 or ∼4% dopant atom concentration: 4% when only
the position of the impurity atom will be explored. one W/Ti substitution in model I or two substitutions in model
II are considered and 8% when two W/Ti substitutions are
considered in model I. Further inclusion in these models of the
2. MODEL AND COMPUTATIONAL DETAILS kind of defects experimentally observed: Ti vacancy (VTi) sites
Periodic 3D calculations were carried out using the VASP 4.6 and wolframyl (WdO) species result in slab models having
code26 28 with the projector-augmented wave method either Ti1 3xO2W2x or Ti1 xO2+xWx composition, respectively
(PAW).29,30 In these calculations, the energy was computed (Figure 2). On the latter, an Oad is added per W dopant atom
using the generalized gradient approximation (GGA) of DFT included in the model to allow for the formation of a surface
proposed by Perdew, Burke, and Ernzerhof (PBE),31 and the wolframyl specie. In the former, a VTi site is created every two W
electronic states were expanded using plane wave basis set with a dopant atoms to compensate for the extra charge added in the
cutoff of 500 eV. Forces on the ions were calculated through the system by the W6+ cations. Relative energies are computed for
Hellmann Feyman theorem, including the Harris Foulkes model structures corresponding to either Ti1 3xO2W2x or
correction to forces.32 This calculation of the force allows a Ti1 xO2+xWx composition with respect to the most stable
geometry optimization using the conjugated gradient scheme. configuration of the same composition. Full optimization of
16971 dx.doi.org/10.1021/jp203223f |J. Phys. Chem. C 2011, 115, 16970–16976
The Journal of Physical Chemistry C ARTICLE

Table 1. Relative Energies (ΔEr) and Shortest W O Dis-


tances in Ti1 xO2+xWx Models Organized According to the
Position of the W Dopant Atom
W site ΔEr/eV rW O/Å

5s 0.00 1.76 1.713, 1.772, 1.782


6s 0.92 1.52 1.790, 1.836, 1.850
6u 1.34 1.79 1.817, 1.860, 1.871
6d 1.54 1.78 1.796, 1.817, 1.823

Figure 2. Supercell models of W-doped TiO2 (101) anatase surface


showing the defects introduced to achieve charge compensation: (a)
model with two W atoms incorporating a VTi site (light blue sphere) and
(b) model with one W atom and an extra O included. The models show
the most relevant bond distances (in angstroms). Top: side view.
Bottom: top view. Atom colors: Ti, gray; O, red; W, ochre.

the cell parameters in some representative doped model struc-


tures results in relative changes in cell volume similar in both sign
and magnitude to the experimental data (∼0.8 to 0.9%). Con-
sequently, the changes in the computed relative energies are Figure 3. Relative energies of Ti1 xO2+xWx configurations as a func-
quite small, 0.03 eV in the biggest case. So, we have kept constant tion of the W Oad distance labeled according to the position of the W
the cell parameters in these calculations given the small influence impurity.
of the remaining stress in the relative energies.
Molecular dynamics (MD) calculations were performed on presence of surface species. In particular, in W-doped TiO2, it
the Born Oppenheimer surface in the canonical ensemble at a is quite clear that bulk charge neutrality is achieved by formation
given temperature (298 K) considering different dopant contents of cation vacancies. However, surface WdO entities are also
and defects. The temperature was controlled through the Nose well-characterized by the presence of a line at ∼970 cm 1 in the
thermostat35,36 implemented in VASP. The numerical integra- Raman spectra of Ti W binary metal oxides.18 Thus, it is
tion was done using a time step set to 3 fs. The system was insightful to start by examining the interaction between the
equilibrated for 1 ps, followed by a 3 ps production run. dopant atoms and additional Oad atoms, available to produce
Histograms for both systems’ temperature and total energy surface wolframyl species. Table 1 summarizes the energetics of
during the production run were extracted. (See the Supporting Ti1 xO2+xWx configurations examined, organized according to
Information.) The width of these histograms confirms that the the position of the W impurity. The individual relative energies
system was adequately equilibrated because the temperature of are plotted in Figure 3 against the W Oad distance.
the system is found to be 298 K ( 6% and the energy of the The computed relative energies suggest that without the
system is found to be constant within (0.5 eV. Both values are presence of VTi defects W atoms clearly prefer a superficial
considered to be reasonable given the system’s size. These MD pentacoordinated position, binding the Oad at a short 1.71 Å.
calculations were performed at the Γ point of the Brillouin zone This distance is typical of a WdO double bond and is in close
as a larger 4  4  3 slab was considered, which results in a model agreement with the experimentally W O distance of 1.75 Å
composed of 288 atoms (96 Ti and 192 O). For these calcula- deduced from Raman spectra information. Substitution of a
tions, concentrations of 4 and 8% of W dopant atoms and a surface Ti6s by a W atom is moderately less stable (0.92 eV),
mixture of both defects (adsorbed oxygen and titanium whereas all subsurface W/Ti substitutions examined show to be
vacancies) are taken into account. From these MD runs, the quite unstable (1.34 to 1.78 eV per W/Ti substitution). It should
radial distribution function (rdf) of the W O pair and the be noted that we have attempted to check the possibility of
vibrational density of states of these systems were obtained and introducing the charge compensation O anion as an interstitial
analyzed. The vibrational spectrum of the system was obtained impurity. However, when the O atom was initially located near to
by considering the Fourier transform of the velocity autocorrela- the W dopant, the resulting final configuration displayed the
tion function (VACF) from the MD data.37 All atoms in the formation on a surface WdO entity with no interstitial O anion.
above-mentioned slab model are included in the VACF. If the interstitial O atom was initially located far from the W
impurity, then the resulting geometry was less stable, ∼2.5 eV
3. RESULTS AND DISCUSSION with respect to the reference configuration. The trends found can
be related to the distortions introduced in the TiO2 anatase
3.1. Structure. Experimentally quite intrincated interactions network by the introduction of the dopant. The W O distances
are found to exist in transition-metal doped TiO2 between displayed on the same Table 1 show that the introduction of the
different types of defects: cation/anion vacancies and the W-dopant implies a shortening of the cation anion distances
16972 dx.doi.org/10.1021/jp203223f |J. Phys. Chem. C 2011, 115, 16970–16976
The Journal of Physical Chemistry C ARTICLE

factor seems to be the separation of the W cations. On average,


structures with W W distances lower than 4.0 Å seems to be less
stable by ∼0.7 eV than structures in which the two W W
dopants are 4.5 to 5.5 Å apart. The reason behind this behavior is
that in the first group both W6+ cations are first cationic
neighbors, so the distortions introduced in the TiO2 network
are concentrated into a small volume. In the second group of
structures, on the contrary, whereas the two W6+ cations are still
first neighbors to the VTi site, they are second cationic neighbors,
so there is more room to accommodate the distortions intro-
duced by the doping process. Further separation of the two W
dopants will result in one of the two W atoms being separated
from the VTi site. However, data from Ti1 3xO2W2x models with
a 4 atom % W concentration (not shown) indicate that this
Figure 4. Relative energies of Ti1 3xO2W2x configurations (8% dopant situation is less favorable by ∼1.0 eV with respect to configura-
concentration) as a function of the W W distance and labeled accord- tions where the two W dopants are first cationic neighbors to the
ing to the positions (surface or subsurface) of the three defects Ti vacant site. These trends can be related to experimental
introduced (two W and one VTi) into the TiO2 structure. Key: m,
number of defects in surface layer; n, number of defects in subsurface
findings that indicate a highly homogeneous W Ti mixed oxide
layer. Triangles: no formation of WdO. Circles: a WdO entity is distribution where the coordination number of the W Ti shell is
observed. roughly three, supporting those models where charge compensa-
tion is fundamentally achieved by the presence of cationic Ti
with respect to those characteristic of the bare anatase structure vacancies.
(1.93 to 1.98 Å). This explains that in the absence of VTi defects To obtain further understanding of the short-range structure of
that may ease the deformations introduced by the W-dopant, W Ti mixed oxides, we have performed MD simulations on
surface substitutional sites are preferred over subsurface ones larger 4  4  3 TiO2 slab models with anatase structure having 4
because they require lower anatase structure distortion. and 8% Ti atoms substituted by W atoms. The W atoms occupy
EXAFS spectroscopy data suggest the exclusive presence of a both surface and subsurface positions, and charge compensation
single local arrangement, where W-dopants have one VTi site in is achieved in these models by a mixture of both VTi and Oad
their first coordination shell.18 To obtain a detailed understand- defects. From these MD calculations, we have obtained the rdf for
ing of the interaction among W-dopants, VTi sites, and WdO the W O pair that is displayed in Figure 5. The first feature in the
entities that may simultaneously coexist, we have performed rdf is a band centered at ∼1.9 Å asymmetric toward shorter W O
DFT calculations on Ti1 3xO2W2x models that introduce a VTi distances. The asymmetry of this band can be interpreted in terms
every two W/Ti substitutions to achieve the necessary charge of two differentiated distances in the first coordination shell of
compensation. We have not systematically explored all possible W6+ cations: a shorter one at ∼1.7 Å and a longer one at 2.0 Å
distributions of W-dopants and VTi sites in our models because (labeled A and B, respectively, in the same rdf diagram). These
the number of different configurations to examine was too large. features are associated, respectively, with double-bonded WdO
Instead, taking advantage of the energetic preference, previously species and with single-bonded W O. (See the inset in Figure 5.)
discussed, for the formation of WdO entities and the available A second and less intense feature appears in the rdf at W O
experimental information, we have concentrated our configura- distances of ∼2.4 to 2.6 Å (labeled C in the rdf diagram)
tional space search into distributions where at least one W-do- associated with surface WdO entities that significantly displace
pant and one VTi site are first cationic shell neighbors. a subsurface O anion. Integration of the rdf (Figure 5) indicates
Figure 4 shows the relative energies of Ti1 3xO2W2x config- that W atoms have a 5 + 1 coordination with the 5 first O atoms
urations at the 8 atom % W-dopant concentration tested. These located at distances below 2.0 Å and the last one displaying a 2.2 to
data and examination of the optimized structures obtained clearly 2.6 Å W O distance. These results can help us to understand the
indicate that the preferred configurations are those on which one experimental EXAFS data18 that clearly show, for samples with W
of the W6+ cations and the VTi site are first cationic neighbors content below 15 atom %, that the first coordination shell of the
with simultaneous formation of a WdO entity. This situation is W cations displays two well-differentiated distances at 1.81 to
most easily achieved when the three defects are at the surface. 1.83 Å and 2.54 to 2.61 Å with a predominance of 5 + 1 coordina-
The presence of a VTi site near a bicoordinated O anion leaves tion. The shorter W O distances are associated with features A
this atom in an undercoordinated situation, and stabilization is and B in our rdf data. The longer W O distances should be
achieved by formation of a surface wolframyl group. However, interpreted as contributions arising from surface WdO entities
formation of WdO entities is also possible even when the three that significantly displace the subsurface anions. This last feature
defects (the two W-dopants and the VTi) are at subsurface was not reproduced in previous LDA calculations in bulk models
positions. The mechanism of formation of the wolframyl is with anatase structure that considered that the only relevant
similar to that previously discussed: the formation of the VTi charge compensation mechanism was the formation of VTi sites.18
reduces the coordination of one tricoordinated O anion and The inclusion in our models of surface (and subsurface) WdO
facilitates its capture by the nearby W6+ cation, although the entities thus proves to be quite relevant. In any case, it is clear that
WdO species remains in close coordination to a nearby Ti these materials have a complex defect structure resulting from the
cation. Whereas the data presented in Figure 3 indicate that the interaction of W-dopants, associated VTi sites, and surface addi-
main driving force to system stabilization is the formation of tional O anions that certainly must be present in the charge
wolframyl species that result from the proximity as first cationic compensation scheme. The analysis of the MD data is completed
neighbors of the W-dopant and the VTi, a secondary stabilization by the calculation of the vibrational spectra of those systems.
16973 dx.doi.org/10.1021/jp203223f |J. Phys. Chem. C 2011, 115, 16970–16976
The Journal of Physical Chemistry C ARTICLE

Figure 5. Radial distribution function for the W O pair (left) and integrated coordination number for the W atoms (right). Inset showing the local
structure around the W atom with its six (5 + 1) first neighbors labeled.

description of the electronic structure of defected semiconductor


oxides.40,41 However, their application is rather expensive when
associated with the use of a plane wave basis set, and the amount
of HF exchange included in the potential is an external input that
largely affects the final description.42,43 Alternatively, the so-
called DFT+U approaches correct some of the inadequacies of
pure GGA functionals by making use of a Hubbard-like term to
account for the strong on-site Coulomb interactions. However,
this approach suffers from the dependence of the results on the
value of U, de facto an empirical parameter that has to be
adjusted. Whereas there is some consensus in the literature on
a value of U for the Ti 3d states of ∼4.0 to 5.0 eV to be the most
adequate,44,45 the situation is not the same when one looks for
Figure 6. Computed vibrational spectra (150 1200 cm 1 region) for the value of U to be applied to the W 5d states. In the previous
the W-doped (101) anatase surface models (continuous line) and work of Long and English23 on bulk W-doped TiO2, different U
simulated experimental Raman WdO band (dotted line, data taken values were applied to the W dopant atoms. Values of U e 2.0 eV
from ref 18). as well as a standard GGA approach were found to result in an
incorrect metallic ground state. Thus, a value of U = 3.0 eV was
In agreement with the experimental infrared (IR) spectrum of finally selected as in this case, W-doped TiO2 exhibited the
anatase,38 our simulated vibrational spectra for the (101) correct semiconductor behavior. It is important to stress that
W-doped anatase surface (Figure 6) display a series of strong these calculations did not take into account the presence of Ti-
absorptions in the 100 850 cm 1 range. Neither anatase nor vacancies or wolframyl species or any other charge compensation
rutile TiO2 shows any IR or Raman absorption above ∼850 cm 1. mechanism. As result, in their calculations, there is an excess of
However, in our computed vibrational spectrum, a new feature electrons that may be responsible for the incorrect metallic
develops at ∼1000 cm 1 associated with the WdO stretch mode. electronic structure found.
This data is in agreement with the experimental line at In any case, GGA-DFT methods are useful when their results
∼970 cm 1 observed in the Raman spectrum of nanostructured are critically used to understand trends and obtain qualitative
Ti W mixed metal oxides18 (Figure 6). In light of these results features on the electronic structure of semiconductor oxides. To
and of the experimental data mentioned above, we can conclude this end, we have examined the DOS of different W-doped
that the surface tungsten species detected experimentally is anatase surface models. In Figure 7, we report the computed total
definitely a surface wolframyl group. DOS for the 4% W-doped Ti1 3xO2W2x model and, to establish
3.1. Electronic Structure. It is well known that GGA-DFT comparisons, we also illustrated the total DOS of the undoped
methods usually underestimate the band gap values due to anatase surface obtained from a similar slab model. The atom
inherent deficiencies in the functional, mainly the insufficient projected density of states (PDOS) for the 4% W-doped
cancellation of the self-interaction energy.39 The use of hybrid Ti1 3xO2W2x model is reported in Figure 7. The DOS of the
exchange-correlation functional allows us to overcome these W-doped and undoped surface models has been aligned to have
deficiencies, and some of them are known to provide a good the same Fermi level reference. Similar plots have been obtained

16974 dx.doi.org/10.1021/jp203223f |J. Phys. Chem. C 2011, 115, 16970–16976


The Journal of Physical Chemistry C ARTICLE

covalency also evidenced by the overlap of W 5d and O 2p levels


at the lower edge of the valence band. The large Bader charge
found for the W cation with respect to that of the Ti cations
implies that the high electric field generated by the W cation will
drain electron density from the nearest Ti O bonds (the O
atoms have a 1.42e Bader charge, whereas in the undoped
model, the average charge is 1.36e ), weakening these Ti O
bonds and lowering the position of the associated Ti 3d levels in
the conduction band.

4. CONCLUSIONS
In the present work, a systematic study of the interaction
between W-dopant atoms and structural point defects experi-
mentally evidenced in nanostructured TiO2 materials has been
carried out by means of plane-wave pseudopotential calculations.
We focus this study on the (101) anatase surface because this is
the main structural TiO2 polymorph found in these kind of
materials, and the (101) surface is the thermodynamically most
stable crystal face. Our calculations show that without the
presence of VTi defects, W atoms clearly prefer a superficial
pentacoordinated position, binding the extra O atom (added to
the model to compensate the charge of the W6+ cation) at a short
1.71 Å distance, typical of a WdO double bond, whereas all
subsurface W/Ti substitutions examined show to be less stable.
These findings are related to the distortions introduced by the
Figure 7. Total (top) and atom-projected (bottom) density of states of doping process into the anatase structure: in the absence of VTi
4% doped Ti1 3xO2W2x model and undoped TiO2 anatase (101) defects that may ease the deformations introduced by the
surface models. The origin is at the Fermi level of both systems. The
inset shows the lower edge of the conduction band.
W-dopant, surface substitutional sites are clearly preferred be-
cause they require lower structural distortions. DFT calculations
on models that introduce a VTi site every two W/Ti substitutions
regardless of the W substitutional position and charge compen- clearly indicate that the preferred configurations are those on
sation scheme considered. Therefore, the electronic structure of which one of the W6+ cations and the VTi site are first neighbors,
the W-doped anatase TiO2 surface is found to be essentially with simultaneous formation of a WdO entity. Whereas this
insensitive to the position and concentration of W dopant atoms. situation is most easily achieved when the three defects are at the
By comparing these DOS, we can see two evident changes that surface, it is also possible when the three defects are at subsurface
take place upon W doping the TiO2 anatase surface. The first one positions. The mechanism of formation is explained in terms of
is the occurrence, at the lower level energy edge of the valence the reduction of the coordination of one O atom by the presence
band of a series of states with mainly O 2p character but with a of the VTi and its posterior capture by the W-dopant, with
nonnegligible contribution of the W 5d energy levels. (See formation of the wolframyl group. Whereas it is found that the
PDOS in Figure 7.) These features are associated with the O main driving force to system stabilization is the formation of
atoms that are first neighbors of the W dopant (except the O wolframyl species, a secondary factor seems to be the separation
from the WdO, which appears in the middle of the valence of the W6+ cations: if both W6+ cations are first cationic
band), and the overlap of the O 2p and W 5d energy levels found neighbors, then the distortions introduced into the TiO2 net-
is indicative of some degree of covalency on the W O bond. The work are concentrated into a small volume; if both W6+ cations
second modification of the DOS that takes place upon W doping are second cationic neighbors, then there is more room to
the anatase surface is found at the lower edge of the conduction accommodate these distortions. The rdf for the W O pair
band. (See the inset in Figure 7.) This feature (that is similar in all obtained from MD calculations in models where charge com-
W-doped surface models) is responsible for the ∼0.1 to 0.2 eV pensation is achieved by a mixture of both VTi and Oad defects
lowering of the band gap found in W Ti mixed oxides and is shows a predominance of 5 + 1 coordination for W6+ cations,
mainly associated with a decrease in energy of the lowest Ti 3d with two W O distances at 1.8 to 2.0 and 2.5 to 2.6 Å, allowing
levels on the conduction band, as can be deduced from the atom for a rationalization of experimental EXAFS data. Our MD data
projected DOS. (See Figure 5b.) Detailed inspection of the calculations allow us to assign these longer W O distances to
atom-projected DOS shows that the majority of this contribution WdO entities that significantly displace a nearby O anion. In its
comes from the Ti atoms that are closer to the W impurity. turn, the formation of these WdO groups is the consequence of
These observations can be rationalized by examining the the presence of a VTi site that reduces the coordination of one
results of a Bader charge analysis. Of course the results of this tricoordinated O anion and facilitates its capture by the nearby
charge distribution analysis cannot be taken as absolute electron W6+ cation. The computed vibrational spectrum, obtained from
population values but are useful for comparative purposes. The our MD simulation, displays a band at ∼1000 cm 1 associated
formally W6+ cation has an average charge of +4.46e , whereas with the WdO stretch mode, in agreement with the experi-
the Ti cations are found to have a Bader charge of +2.63e . The mental Raman line at ∼970 cm 1. The combination of our
charge found the W-dopant is consistent with a high degree of structural data, the computed vibrational spectrum, and the
16975 dx.doi.org/10.1021/jp203223f |J. Phys. Chem. C 2011, 115, 16970–16976
The Journal of Physical Chemistry C ARTICLE

available experimental information definitely proves that the (12) Lazzeri, M.; Vittadini, A.; Selloni, A. Phys. Rev. B 2001,
tungsten surface species experimentally detected is a WdO 63, 155409.
group. Comparisons of the DOS for doped and undoped (13) Janisch, R.; Gopal, P.; Spaldin, N. A. J. Phys.: Condens. Mater.
(101) anatase surface models allow us to understand the changes 2005, 17, R657–R689.
that take place in the electronic structure of these materials upon (14) Di Valentin, C.; Diebold, U.; Selloni, A. Chem. Phys. 2007,
339, vii–viii.
W-doping. First, at the lower edge of the valence band, a series of (15) Kim, D. S.; Yang, J.-H.; Balaji, S.; Cho, H.-J.; Kim, M.-K.; Kang,
levels associated with the O atoms that are first neighbors of the D.-U.; Djaoued, Y.; Kwon, Y.-U. CrystEngComm 2009, 11, 1621.
W dopants but with a nonnegligible contribution of the W 5d (16) Tian, H.; Ma, J.; Li, K.; Li, J. Mater. Chem. Phys. 2008, 112, 47.
energy levels is found. The overlap of the O 2p and W 5d levels (17) Kubacka, A.; Colon, G.; Fernandez-García, M. Catal. Today
found in these features is indicative of an important covalent 2009, 143, 286.
contribution to the bonding, confirmed by the results of a Bader (18) Fernandez-García, M.; Martínez-Arias, A.; Fuerte, A.; Conesa,
charge analysis. The second modification of the DOS is found on J. C. J. Phys. Chem. B 2005, 109, 6075.
the lower edge of the conduction band associated with a decrease (19) Kim, H. Y.; Lee, H. M.; Pala, R. G. S.; Shapovalov, V.; Metiu, H.
in energy of the lowest Ti 3d levels with a majority contribution J. Phys. Chem. C 2008, 112, 12398.
coming from the Ti atoms that are closer to the W impurity. This (20) Aryanpour, M.; Hoffmann, R.; DiSalvo, F. J. Chem. Mater. 2009,
21, 1627.
feature is responsible of the ∼0.1 to 0.2 eV lowering of the band (21) Kamisaka, H.; Suenaga, T.; Nakamura, H.; Yamashita, K. J. Phys.
gap found in W Ti mixed oxides and is found to be independent Chem. C 2010, 114, 12777.
of the dopant atom concentration, in agreement with the (22) Long, R.; English, N. J. Appl. Phys. Lett. 2009, 94, 132102.
experimental information. (23) Long, R.; English, N. J. Chem. Mater. 2010, 22, 1616.
(24) Jiang, C. Z.; Dou, Y. L. Adv. Mater. Res. 2011, 239 242, 2671.
’ ASSOCIATED CONTENT (25) Xu, M.; Gao., Y.; Martínez Moreno, E.; Kunst, E.; Muhler, M.;
Wang, Y.; Idriss, H.; W€oll, C. Phys. Rev. Lett. 2011, 106, 138302.
(26) Kresse, G; Hafner, J. Phys. Rev. B 1993, 47, 558.
bS Supporting Information. Histograms for both systems’
(27) Kresse, G; Furthmuller, J. Phys. Rev. B 1996, 54, 11169.
temperature and total energy during the MD productions runs.
(28) Kresse, G; Furthmuller, J. Comput. Mater. Sci. 1996, 6, 15.
This material is available free of charge via the Internet at http:// (29) Kresse, G.; Joubert, D. Phys. Rev. B 1999, 59, 1758.
pubs.acs.org. (30) Bl€och, P. E. Phys. Rev. B 1994, 50, 17953.
(31) Perdew, J.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996,
’ AUTHOR INFORMATION 77, 3865.
(32) Harris, J. Phys. Rev. B 1985, 31, 1770.
Corresponding Author (33) Ortega, Y.; Cruz, N.; Menendez-Proupin, E.; Graciani, J.; Fdez.-
*E-mail: marquez@us.es. Sanz, J. Phys. Chem. Chem. Phys. 2011, 13, 11340.
(34) Burdett, J. K.; Hughbanks, T.; Miller, G. J., Jr.; Richardson,
J. W.; Smith, J. V. J. Am. Chem. Soc. 1987, 109, 3639.
’ ACKNOWLEDGMENT (35) Nose, S. Mol. Phys. 1984, 52, 255.
This work was funded by the Ministerio de Ciencia e (36) Nose, S. J. Chem. Phys. 1984, 81, 511.
(37) Postnikov, A. In Computational Materials Science; Catlow, R.,
Innovacion (Spain, grants MAT2008-04918, CSD2008-0023) Kotomin, E., Eds.; NATO Science Series, IOS Press: Amsterdam, 2003;
and the Junta de Andalucía (P08-FQM-3661). Computational p 162.
resources were provided by the Centro Nacional de Super- (38) Coblentz Society, Inc. Evaluated Infrared Reference Spectra. In
computacion-Barcelona Supercomputing Center (Spain) and NIST Chemistry WebBook; Linstrom, P.J., Mallard, W.G., Eds.; NIST
Centro Informatico Científico de Andalucía (CICA). We thank Standard Reference Database Number 69; National Institute of Stan-
Mr. Roger Nadler for providing us with the program to obtain the dards and Technology: Gaithersburg, MD. http://webbook.nist.gov
vibrational spectra from the MD data. (accessed May 21, 2011).
(39) Koch, W.; Holthausen, M. C. A Chemist Guide to Density
Functional Theory; Wiley: Weinheim, Germany, 2002.
’ REFERENCES (40) Perdew, J. P.; Ernzerhof, M.; Burke, K. J. Chem. Phys. 1996,
(1) Henrich, V.; Cox, P. The Surface Science of Metal Oxides; 105, 9982.
Cambridge University Press: Cambridge, U.K., 1994. (41) Heyd, J.; Scuseria, G. E. J. Chem. Phys. 2006, 125, 224106.
(2) Linsebigler, A. L.; Lu, G.; Yates, J. T. Chem. Rev. 1995, 95, 735. (42) Graciani, J.; Marquez, A. M.; Plata, J. J.; Ortega, Y.; Cruz, N. C.;
(3) Henry, C. R. Surf. Sci. Rep 1998, 31, 231. Meyer, A.; Zicovich-Wilson, C. M.; Fdez.-Sanz, J. J. Chem. Theory
(4) Carp, O.; Huisman, C. L.; Reller, A. Prog. Solid State Chem. 2004, Comput. 2011, 7, 56.
32, 33. (43) Wang., F.; di Valenti, C.; Pacchioni, G. J. Phys. Chem. C 2011,
(5) Diebold, U. Surf. Sci. Rep 2003, 48, 53. 115, 8345.
(6) Thompson, T. L.; Yates, J. T. Chem. Rev. 2006, 106, 4428. (44) Calzado, C. J.; Hernandez, N. C.; Sanz, J. F. Phys. Rev. B 2008,
(7) Pang, C. L.; Lindsay, R.; Thornton, G. Chem. Soc. Rev. 2008, 77, 045118.
37, 2328. (45) Finazzi, E.; di Valentin, C.; Pacchioni, G.; Selloni, A. J. Chem.
(8) Diebold, U.; Li, S.-C.; Schmid, M. Annu. Rev. Phys. Chem. 2010, Phys. 2008, 129, 154113.
61, 129.
(9) Diebold, U. Nat. Chem. 2011, 3, 271.
(10) Fujishima, A.; Zhang, X.; Tryk, D. A. Surf. Sci. Rep. 2008,
63, 515.
(11) Fernandez-García, M.; Rodriguez, J. A. Metal Oxide Nano-
particles. In Nanomaterials: Inorganic and Bioinorganic Perspectives;
Lukehart, C. M., Scott, R. A., Eds.; Wiley-Blackwell: West Sussex,
U.K., 2008.

16976 dx.doi.org/10.1021/jp203223f |J. Phys. Chem. C 2011, 115, 16970–16976

You might also like