Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Article

pubs.acs.org/JPCC

Presence of Gap States at Cu/TiO2 Anatase Surfaces: Consequences


for the Photocatalytic Activity
Nicola Seriani,*,† Carlos Pinilla,‡,§ and Yanier Crespo†,∥

The Abdus Salam International Centre for Theoretical Physics, Strada Costiera 11, 34151 Trieste, Italy

University College London, London WC1E 6BT, England
§
Departamento de Fisica, Universidad del Norte, km 5 Via Puerto Colombia, Barranquilla, Colombia

International Institute of Physics, Av. Odilon Gomes de Lima, 1722, Capim Macio, CEP 59078-400, Natal RN Brazil
*
S Supporting Information

ABSTRACT: Copper-modified titania is a system of interest for its potential for


photocatalytic applications in the production of solar fuels. Still, the role of copper in the
process is unclear. In this work, small copper clusters on the (101) and (100) surfaces of
anatase have been investigated by first-principles simulations based on density functional
theory, to shed light on their atomic and electronic structure, and to understand their effect
on the photocatalytic process. The main effects of copper on the electronic structure are to
provide states above the edge of the valence band of titania and to lead to the formation of
midgap states. There are two types of midgap states, respectively, associated with direct
Cu−Ti bonds and to Ti3+ polarons. The latter are the result of charge donation from
copper and lie in the vicinity of the surface. Moreover, the copper tetramer (Cu4) displays
empty states at the bottom of the conduction band that play a key role in accommodating
excess electrons. We discuss how these features should enhance the photoresponse of
TiO2, contribute to increase the lifetime of the photogenerated electron−hole pairs and
contribute to increase the activity of this material for CO2 reduction, a key step in the
photoproduction of hydrocarbons.

■ INTRODUCTION
Titania is the material of choice in photocatalytic applications
decrease in the band gap of the material,21,22 but no equivalent
information is available for supported copper nanoparticles. On
and has been the subject of intense scrutiny to understand the the other side, the electrocatalytic conversion of CO2 to
mechanisms of the photocatalytic processes.1−13 Notwithstand- hydrocarbons seems to be dependent on the particle shape and
ing its popularity, it still suffers from major drawbacks that surface morphology of copper catalysts.23 Such a dependence of
hinder its widespread use in commercial applications. These catalytic activities on particle shape is typically linked to the
disadvantages comprise a relatively large band gap, ∼3 eV, that surface chemistry of the catalyst. The oxidation state of copper
prevents photoabsorption in the visible range, and a low overall plays a role in the activity of the catalyst as well, with
efficiency of the photocatalytic process. For these reasons, nanocrystalline metallic copper being particularly active for the
many attempts to enhance its properties through proper electroreduction of CO.24 Also, Liu et al. investigated the effect
catalyst modifications have been undertaken. Among these, the of different treatments during catalyst preparation for a Cu/
addition of copper to titania has recently received considerable TiO2 photocatalyst for CO2 reduction, and found that a
attention, as this system is able to produce hydrocarbons from reducing treatment gives the best activity.25 Moreover, a high
carbon dioxide and water vapor under solar illumination.14−16 dispersion of the copper catalyst is important to achieve high
TiO2 nanotubes with copper particles led to the production of conversion rates.26 Regarding the mechanistic aspects of surface
carbon monoxide, hydrogen, and hydrocarbons.14−16 Hydrogen chemistry of the Cu/TiO2 system, is is known that CO2
production was also observed when copper particles were dissociates spontaneously on defective Cu(I)/TiO2 in the
embedded in nanostructured TiO2.17−20 dark.27 It was proposed that electrons, trapped at Ti3+−VO sites,
Still, fundamental questions remain unanswered about the can migrate to CO2 resulting in a dissociative electron
mechanisms of photoabsorption, charge separation and trans- attachment and in CO adsorbed to Cu+.27 This hypothesis
fer, molecule adsorption, and reaction. In particular, the role of underlines the importance of the reduced state of the catalyst
copper in each of these steps remains to be elucidated, and such and relies on the presence of trapped electrons in the vicinity of
understanding could lead to a more systematic improvement of
the catalyst’s properties. Regarding the effect of copper on Received: January 27, 2015
photoabsorption, the only experimental evidence regards Revised: March 10, 2015
copper doping in titania, which has been shown to lead to a Published: March 10, 2015

© 2015 American Chemical Society 6696 DOI: 10.1021/acs.jpcc.5b00846


J. Phys. Chem. C 2015, 119, 6696−6702
The Journal of Physical Chemistry C Article

the surface. These trapped electrons were not identified in the performed with the Quantum-Espresso code,55,56 employing
experiments in the presence of copper. For pure titania it is plane-wave expansion and Vanderbilt ultrasoft pseudopoten-
known that the reduced phase (TiO2−x) shows a Ti3+ state tials.57 Calculations were performed with the simplified version
located within the band gap.28−30 Regardless of whether the of DFT+U by Dudarev et al.,58 as implemented in the
Ti3+ state is related to an oxygen vacancy27 or a titanium Quantum-Espresso code.59 The values of U have been taken
interstitial defect,31 it is generally accepted that it has an from the literature: a value of 4.2 eV was used for
essential role in enhancing the photocatalytic performance of titanium42,43,53 and of 5.2 eV for copper.60 Convergence of
this material.31−34 However, whether copper has an effect on the total energy, for bulk anatase, has been achieved using
the presence of Ti3+ sites remains an open question. Moreover, cutoffs of 40 and 480 Ry for wave functions and charge density
it has been shown that metal nanoparticles enhance the charge respectively, and a grid of (4 × 4 × 2) k-points generated with
separation through the formation of a Schottky barrier, and the method of Monkhorst and Pack.61 The (101) surface has
photogenerated electrons are transferred to the metal nano- been modeled with a rectangular (1 × 2) surface cell of size
particles as their Fermi energy is typically lower than that of the (10.27 Å × 7.57 Å), the (100) surface with a (2 × 1) surface
oxide.27,35 Whether this is true also for copper remains to be cell of size (7.57 Å × 9.55 Å). This ensures that the minimal
seen. It is thus clear that a better characterization of the effects distance between clusters is at least of 5 Å, avoiding direct
of copper on the electronic structure of the system is necessary binding between neighboring clusters. This corresponds to less
and would contribute to shedding light on the overall than 0.5 monolayers of copper on titania. Similar fractional
mechanism of the photocatalytic process and the specific role coverages were obtained also in experiments.36 The k-point
of copper in it. grids were chosen to have a k-point density similar to that
Given the many variables that can influence the activity of the employed for the bulk. The TiO2 slabs have a thickness of ∼10
catayst (shape, oxidation state, mixing state with titania, ...) and Å, and the vacuum between the TiO2 slab and its copy is ∼15 Å
the many subprocesses involved in the photocatalytic reaction thick. Clusters have been deposited on one side of the slab. We
(photoabsorption, charge separation and transfer, molecule have checked that this does not introduce any important
adsorption, chemical reactions, product desorption), it is dipolar interaction with the copies: by increasing the vacuum by
important to be able to consider different aspects of the 20 bohr (corresponding to ∼10.6 Å), the surface energy
system one at a time, to disentangle the complex dependencies changes remain well below 1 meV/Å2. All atomic positions
of all these issues with one another. In this context, help can were relaxed and atomic relaxations were performed until forces
come from computer simulations from first principles. In the were smaller than 10−3 a.u.
present work, we have investigated small copper clusters (1−4
atoms) on the (101) and (100) surfaces of anatase by density
functional theory (DFT) and DFT+U methods, to shed light
■ RESULTS AND DISCUSSION
Copper clusters deposited on two surfaces of anatase TiO2 have
on the interaction between copper and titania and its effect on been considered: (101) and (100). Anatase is the most active
the electronic structure of the system. It is indeed possible to titania phase for photocatalytic applications,62 and these are the
obtain atomically dispersed copper on titania surfaces.36−38 We two surfaces with the lowest surface energies.63−66 For the sake
shed light on the interaction between copper and titania and its of brevity, in the main body of the article, we show the figures
effect on the electronic structure of the system. We find that relative to the (101) surface and we comment about the
copper can induce both a shift in the bands of the material and differences between the two surfaces. The full figures for the
the appearance of midgap states, depending on its coordination. (100) surface are reported in the Supporting Information (SI).
Moreover, charge donated by the copper clusters to the TiO2 Atomic structures of small copper clusters (with 1−4 atoms)
surface induces the formation of Ti 3+ species in its have been generated by taking several configurations used in
neighborhood. We argue that the presence of the dipole the literature for similar systems, that is, Pd on TiO2(101)67
formed by the cluster and the Ti3+ species has an important role and Ag on TiO2(100),68 and by fully relaxing them. This
in delaying the recombination of the photoinduced electron approach is justified by the fact that we are considering small
and hole, where the hole is mainly attracted to the Ti3+ site, clusters, of up to four atoms. In principle, extensive structure
while electrons get trapped in the Cu4 cluster. search employing global optimization techniques such as
In the next section, the computational methods are genetic algorithms can be used to find the lowest-energy
described, then the results are presented and discussed. Finally, structures of clusters and surfaces.69,70 Still, for such small
a summary closes the article.


clusters these techniques invariably yield simple planar or
pyramidal geometries.71,72 Of course, the complexity of the
COMPUTATIONAL METHODS geometries increases very quickly if one considers clusters
Spin-polarized density functional theory (DFT) in the larger than those investigated in the present work.
generalized-gradient approximation (GGA) of Perdew, Burke, Structural Properties. TiO2(101). The configurations of
and Ernzerhof (PBE)39 has been used throughout, also with the the energetically most stable clusters on defect-free (101) are
addition of the Hubbard term (DFT+U).40 It is well- shown in Figure 1. In Table 1 the adsorption energies of all the
known41−44 that simple DFT-based approaches fail to predict clusters are reported, calculated as
the presence of the Ti3+ state in the electronic structure, due to
the self-interaction error inherent to the functionals used in the ECu@slab − Eslab − N × ECu atom
Eads = −
local density and the generalized gradient approximations. To NCu atoms (1)
correct this error, post-DFT methods like DFT+U and hybrid
functionals have been implemented.41−54 In fact, it has been where ECu@slab is the total energy of the system with the cluster
shown that DFT+U is able to improve the description of the on the slab, Eslab is the total energy of the bare surface, and
electronic structure of 3d transition metal oxides, including that ECu atom is the energy of an isolated copper atom, and NCu atoms
of localized Ti3+ defect states.30,42−44,46,53 All simulations were is the number of copper atoms in the cluster.
6697 DOI: 10.1021/acs.jpcc.5b00846
J. Phys. Chem. C 2015, 119, 6696−6702
The Journal of Physical Chemistry C Article

slightly elongated with respect to the smaller clusters (2.37−


2.38 Å). We calculated also planar geometries for the tetramer,
but they lie ∼0.5 eV higher. It is to be noticed that, in vacuum,
the planar geometry is the lowest-energy configuration;74 the
pyramid is thus stabilized by the surface. As in the case of the
Cu2 cluster, in the Cu4 there is a direct Cu−Ti bond, although
with a longer distance of 2.92 Å. As we will show later, this
bond will introduce changes in the electronic structure that
could play an important role in the photocatalytic activity of the
Cu/TiO2 compound.
TiO2(100). TiO2(100) has a higher surface energy as a
consequence of the higher number of dangling bonds present at
the surface. In the stable configuration of the Cu monomer, the
distance of copper to two oxygen atoms is 1.89 Å, and its
distance to two Ti atoms is 2.82 Å (see Figure 1a of the SI).
The Cu atom lies almost exactly along the internuclear O−O
axis, whereas silver lies at the center of a triangle formed by
three O atoms (see Figure 2a in ref 68). The difference is
Figure 1. Atomic structure of the stable clusters with 1−4 Cu atoms
on the defect-free surface of anatase TiO2(101). Blue balls: titanium; probably due to the different size of the Cu and Ag ions: the
red balls: oxygen; gray balls: copper. smaller Cu would be too far from the oxygen if put at the
center of the triangle, and therefore this configuration lies 0.2
Table 1. Adsorption Energies (Eads), Total Charge Donated eV higher in energy. The interaction with the surface is so
to the Surface (Qdon(|e|)) and Change in the Polarization of strong that the dimer is not stable, and the most convenient
the 6c-Ti+3 Atom (Δμ(μB)) Produced by Copper Clusters on configuration for two Cu atoms is one where each Cu atom lies
Anatase Surfaces (101) and (100), as Calculated with DFT in a monomer configuration, sharing one oxygen with the other
+U Cu atom. In fact, as shown in Table 1, on (100) the monomer
is more stable than all other clusters considered. We take this as
Eads (eV/Cu) Qdon(|e|) Δμ(μB) a sign of the higher reactivity of this surface, as shown also by
Cu (101)/(100) (101)/(100) (101)/(100) the higher surface energy. The trimer prefers a triangular
1 2.30/2.83 0.66/0.69 0.98/0.97 configuration with one copper atom pointing toward the
2 2.08/2.55 0.26/1.34 0.01/0.96 vacuum and being coordinated only to the other two Cu atoms.
3 2.48/2.57 0.73/0.73 0.97/0.95 The triangular configuration is the same as that found on (101).
4 2.34/2.55 0.76/1.27 0.95/0.98 Also, the tetramer prefers to be in the tetrahedral geometry
(pyramid) as on (101), probably favored by the ideal
A single copper adatom (Cu1) prefers to lie at a bridge configuration of the surrounding oxygen atoms, such that
position between two surface oxygen atoms. In this each of the three Cu atoms of the basis of the pyramid are
configuration, the distance of copper from the two neighboring bound to the other three Cu atoms and to one oxygen atom
oxygen atoms is 1.88 Å and its distance to two Ti atoms is 3.04 (1.82−1.84 Å; see Figure 1d of the SI). A flat tetramer parallel
Å. A similar bridge position was found to be the most stable for to the surface lies 0.1 eV higher in energy, while flat tetramers
the copper adatom also on the (110) surface of rutile.73 This perpendicular to the surface, such as those shown in Figure 4 of
structure was also the most stable for palladium on anatase ref 68 lie at least 0.7 eV higher. In the stable trimer and
TiO2(101).67 tetramer there is no direct Cu−Ti bond and the trimer is
On the contrary, the case of the Cu dimer (Cu2) is different slightly more stable than the tetramer on both surfaces.
from palladium. We suspect that this is due to the different Electronic Properties. TiO2(101). The atomic configura-
dimer length and stiffness for Pd and Cu: for palladium, lengths tion has important consequences for the electronic structure of
of 2.63−2.86 Å were calculated, while copper displays lengths the system, showing differences between the results obtained
in a much narrower range (2.28−2.31 Å). The most stable for the smaller clusters Cu1−3 and the Cu4 cluster, as shown in
structure, shown in Figure 1b, is the one that can better Figure 2. The extension of the edge of the valence band (VB)
accommodate this Cu−Cu length on the surface. The other into the gap is mainly due to low-coordinated Cu atoms present
structures reported in Figure 6 of ref 67 lie some tenths of eV in the Cu1−3 clusters (Figure 2a−c). Here the 3d and 4 s states
higher in energy. In the stable dimer structure, the second atom of the copper atoms with coordination number smaller than 3
is bound only to the first Cu and to a five-coordinated Ti (5c- are those that contribute most to the top of the VB, extending it
Ti) atom at the surface, with a Cu−Ti distance of 2.59 Å. In by ∼0.7 eV into the band gap of titania. The higher the
fact, the dimer is the least stable among the cluster sizes coordination number is, the lower in energy the 3d and 4s
considered (Cu1−4), probably because of the unfavorable states; for example, there is a shift of ∼0.6 eV between the peak
geometry of this second Cu atom. of the 3d states of the three-coordinated Cu atom and the two-
The stiffness of the Cu−Cu bond is found also in the trimer coordinated Cu atom in the dimer (see Figure 2 of the SI).
(Cu3), where the bond lengths are 2.31−2.34 Å. The stable This trend is confirmed in the tetramer where just a small
atomic configuration for the trimer has two Cu atoms building contribution, from the 3d states, to the edge of the VB is
a bridge between two surface oxygens, and the third is pointing observed (see Figure 3 of the SI).
toward the vacuum (Figure 1c). The tetramer (Cu4) prefers the A second important feature in the electronic structure is the
compact coordination of a tetrahedron (pyramid), where each appearance of midgap states. There can be two types of such
Cu atom is bound to the other three. Here the distances are states. The first state consists of Cu-4 s and Ti-3d orbitals, is
6698 DOI: 10.1021/acs.jpcc.5b00846
J. Phys. Chem. C 2015, 119, 6696−6702
The Journal of Physical Chemistry C Article

Figure 2. Density of states of the stable clusters with 1−4 Cu atoms on the surface of anatase TiO2(101), calculated with DFT+U. Upper part of
each diagram: majority spin. Solid line: total DOS; red dashed line: Cu-3d states; green solid line with triangles: 4s states of the Cu atom with a Cu−
Ti bond; brown solid line with circles: Ti-3d states of a 6-fold coordinated Ti atom in the vicinity if the Cu cluster (see text); dark brown solid line
with diamonds: Ti-3d states of a 5-fold coordinated Ti atom with a direct Cu−Ti bond. The projections on the single states have been multiplied by
20 to put them on the same scale as the total DOS. The zero of energy has been set to the enegy of the lowest unoccupied orbital. We have chosen
to set the zero of energy to the energy of the lowest unoccupied orbital because this choice makes the electronic structure of different systems easier
to compare.

due to a direct Cu−Ti bond (Figure 2b,d), and is present both donated charge, Qdon, calculated using the Bader method,75,76 is
in DFT and DFT+U. A second state is present only when the between +0.66|e| and +1.34|e| for systems with polarons (and
system is relaxed with DFT+U, because it is due to polaron just 0.26|e| for the dimer on the (101) surface), as shown in
formation at a six-coordinated Ti (6c-Ti) atom lying right Table 1, resulting in a net positive charge present on the copper
below the copper cluster. The important role of the Cu−Ti cluster. Of the donated charge, only 0.2−0.3|e| is transferred to
bond in determining the presence of states in the gap is the Ti atom where the polaron is sitting, while the rest is
confirmed by the analysis of the trimer on (100), where there is delocalized on the oxygen atoms. This Ti atom is usually
no direct Cu−Ti bonding. As a consequence, the midgap state indicated as a Ti3+ ion, but it is known that effective atomic
due to this bond is absent (see Figure 2c). Midgap states due to charges hardly ever reproduce the formal charges in transition
the presence of metal−Ti bonds have also been seen in the case metal compounds (see ref 77 and references therein). This is
of Ag clusters.68 Below we will discuss how the roles of these usually attributed to the idealized nature of formal oxidation
two midgap states for the dynamics of excess charges should be states and to the fact that electronic charge is anyway
different. delocalized over many atoms in the real material. Moreover,
TiO2(100). The projected DOS of clusters on (100) shows effective atomic charges may be built according to different
the same features as on (101) (see Figure 4 of the SI). The principles, either from the total electronic charge distribution,
main differences are related to the different atomic config- as in the case of Bader charges,75,76,78,79 or from the wave
urations with respect to (101). The shortest distance between functions, as in the case of Löwdin charges.80,81 In the latter
the Cu adatom and Ti atoms is 2.82 Å, much larger than the case, the actual values of the effective charges may depend on
Cu−Ti bond found in the dimer on (101), and there are O the choice of basis set and atomic wave functions used for the
atoms in between. As a consequence, the only midgap states are projection. This makes an assignment of charge to atoms not
observed for polarons on subsurface titanium ions. Since the univocal, and therefore the resulting values should be used with
stable configuration with two copper atoms consists of two care. To make the situation more compicated to analyze, the
separately adsorbed Cu atoms, each induces its own subsurface electrons donated to these metal oxides go also to oxygen
polaron. Also for the tetramer, two polarons are found below atoms around the cation.82 Still, it is accepted in the literature
the copper cluster. In these two cases the charge donated to the that a polaron in a transition metal oxide can be individuated by
surface (Qdon) is larger (see Table 1). an investigation of atomic effective charges, of the magnetic
Polaron State. Polaron formation has been observed for all moments and of the geometric distortion of the atomic
systems in both the (101) and the (100) surfaces, with the configuration. In particular, it has also been shown that in these
exception of the dimer on the (101) surface. In all cases we cases the spin localization, given by an increase in the atomic
have characterized the polaron by the magnetic moment of the magnetic moment, is a better indicator of the presence of the
Ti atom, its charge and by the geometric distortion. As Ti3+ ions than the effective charges.83−86 The change in
mentioned above, the polaron forms at a 6c-Ti atom lying right magnetic moment for the Ti atom in question is between 0.95
below the copper cluster. Depending on the cluster size, the and 0.98 μB, as calculated through Löwdin analysis80 (see Table
6699 DOI: 10.1021/acs.jpcc.5b00846
J. Phys. Chem. C 2015, 119, 6696−6702
The Journal of Physical Chemistry C Article

1). Bader and Löwdin analysis give similar results when calculations support the view that copper can induce the
calculating the magnetic moment of an ion.86 The Ti−O formation of further Ti3+ sites, beside those present near
distances are also increased and amounts to 2.16 and 2.14 Å for oxygen vacancies. If it is true that Ti3+ sites play a decisive role
the azimuthal O atoms, and to 2.07, 2.18, 1.98, and 1.97 Å for in fostering the photocatalytic activity, then one of the main
the basal O atoms (here we show data just for the Cu4 cluster positive effects of copper could be that of increasing their
on (101)). For comparison, the other 6-fold coordinated Ti number in the vicinity of the surface.
atom under the Cu cluster has distances of 2.00 and 2.05 We note however that there is a major difference in the
(azimuthal) and 1.89, 1.92, 2.03, and 2.05 Å (basal). The electronic structure between Cu clusters with less than four
polaronic state is visualized in real space in Figure 3a. This state atoms and the Cu4 cluster. The tetramer displays Cu empty
states right at the bottom of the conduction band, while these
states are absent for the other clusters. These states are
important because they are related to the dynamics of
photogenerated electrons. Indeed, metal nanoparticles on
TiO2 can typically trap photogenerated electrons and make
them available for reduction of adsorbed molecules.27,35 This is
not the case for clusters containing less than four atoms, as the
available free states of Cu are high in energy in the CB (Figure
2a−c). On the contrary, the Cu4 cluster shows a localized
empty state at the bottom of the CB, which is the minority spin
channel of the state shown in Figure 3b. This is a state where
Figure 3. Isodensity surfaces for single midgap states for the system photogenerated electrons can be trapped, in agreement with
with the tetramer on (101), as relaxed with DFT+U: (a) polaronic what is usually expected for metal clusters at the surface of
state consisting mainly of t2g states of subsurface six-coordinated titania, promoting charge separation and transfer to the
titanium atoms (see text for a detailed description); (b) state reactants.27,93 This situation is analogous to that encountered
consisting mainly of s and p states of two of the copper atoms. These at macroscopic metal−semiconductor interfaces, where a
states are those of the majority spin and are occupied.
Schottky barrier forms,35 retarding the recombination of a
e−−h+ pair and therefore increasing its lifetime.94,95
is pushed down into the gap and is clearly visible in the DOS To investigate the fate of electrons and holes in this system,
(Figure 2d). Its position is over 1 eV below the edge of the CB, we have considered the electronic structure of charged cells
close to the position of the F-centers created by oxygen with the copper tetramer on (101), a cell with one additional
vacancies at the (110) surface of rutile, which experimentally lie electron and a cell with a hole (Figure 4). The excess electron
0.7−0.9 eV below the edge.53,87−89 We have checked the occupies the minority-spin state at the Cu−Ti bond (Figure
possibility of polaron formation on other Ti ions, and found 4a); this state is spatially similar to its majority-spin counterpart
that the polaron can exist as a metastable state in other depicted in Figure 3b and was the lowest lying unoccupied
positions, with energy differences smaller than 0.2 eV in most
cases. We compare this with the situation of polarons induced
by VO in rutile (110) surface, where activations barriers were
∼0.1 eV,49 allowing for electron hopping at room temper-
ature.53 Therefore, we expect that at room temperature a
polaronic state will form in the 6c-Ti atom below the Cu
cluster, as shown in Figure 3a, and will remain in the near-
surface region. We think that this property is also important for
the enhancement of the photocatalytic performance of the Cu/
TiO2 system as argued below. As a final note, the existence of
such Ti3+ sites in the presence of copper clusters could be
tested by electron paramagnetic resonance.90
Consequences for the Photocatalytic Activity of Cu/
TiO2. We have shown that the effect of the Cu clusters on the
electronic structure of titania is 2-fold. First, Cu atoms with a
coordination number lower than three contribute to the
extension of the edge of the valence band (VB) into the gap.
Second, Cu clusters induce the appearance of midgap states of
two kinds: one state is due to the presence of a Cu−Ti bond,
while the other is a Ti3+ polaron state generated by the
localization of the charge coming from the copper on a 6c-Ti
atom right below the cluster. In pure titania, it has been
suggested that storage of electrons at Ti3+ sites determines the
photocatalytic activity,4,27,91 maybe by increasing the photo- Figure 4. Density of states for the charged cluster with four Cu atoms
absorption.92 Indeed, it is known that polarons easily migrate to system on the surface of anatase TiO2(101). Upper part of each
the surface in titania, and that electrons from Ti3+ sites are diagram: majority spin. (a) System with an excess electron; (b) System
involved in dissociative electron attachment of CO2 even in the with an excess hole. Here we have used the same lines and color codes
dark.27 In pure titania, Ti3+ ions have been observed in of Figure 2. The zero of energy has been set to the energy of the
correspondence of oxygen vacancies (VO-Ti3+ sites).96 Our lowest unoccupied orbital.

6700 DOI: 10.1021/acs.jpcc.5b00846


J. Phys. Chem. C 2015, 119, 6696−6702
The Journal of Physical Chemistry C Article

orbital in the neutral system. The polaron state remains


unchanged. On the contrary, in the system with a hole, the Ti3+
■ AUTHOR INFORMATION
Corresponding Author
polaron state disappears from the DOS, while the other features *E-mail: nseriani@ictp.it.
remain unchanged (Figure 4b). Thus, these results support the
Notes
idea that a copper cluster at the surface of titania is very
The authors declare no competing financial interest.


important for charge separation, with holes migrating toward
Ti3+ polaron sites and electrons moving toward the copper
ACKNOWLEDGMENTS
cluster. Additionally, the presence of Ti3+ near the Cu−Ti
interface could contribute to the creation of active sites for the Computational resources were provided by CINECA under
CO2 reduction near this interface, even in the dark. Indeed, in IscrC-Copsol and by The Abdus Salam ICTP. Y.C.’s position
pure reduced titania it was observed that Ti3+ sites lead to was partly sponsored by the “ERC Advanced Grant 320796 −
MODPHYSFRICT”.


dissociative electron attachment to CO2, with production of a
CO adsorbed to Cu+.27 This hypothesis will be addressed in
future studies. REFERENCES


(1) Fujishima, A.; Honda, K. Nature 1972, 238, 5358.
(2) Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Science
CONCLUSIONS 2001, 293, 269.
In this work, the atomic and electronic structure of small (3) Serpone, N.; Emeline, A. V. J. Phys. Chem. Lett. 2012, 3, 673.
copper clusters (1−4 atoms) on the (101) and (100) surfaces (4) Kamat, P. V. J. Phys. Chem. C 2012, 116, 11849.
of anatase have been investigated by DFT and DFT+U. The (5) Kamat, P. V. J. Phys. Chem. Lett. 2012, 3, 663.
(6) Maeda, K.; Domen, K. J. Phys. Chem. Lett. 2010, 1, 2655.
main conclusions are as follows:
(7) Teoh, W. Y.; Scott, J. A.; Amal, R. J. Phys. Chem. Lett. 2012, 3,
(a) On the more stable (101), copper tends to form clusters; 629.
on (100) the Cu-TiO2 interaction is stronger and, for (8) Girish Kumar, S.; Devi, L. G. J. Phys. Chem. A 2011, 115, 13211.
example, favors single adatoms over dimers. (9) Chen, X.-B.; Liu, L.; Yu, P. Y.; Mao, S. S. Science 2011, 331, 746.
(10) Tao, J.-G.; Luttrell, T.; Batzill, M. Nat. Chem. 2011, 3, 296.
(b) The effect of copper on the electronic structure is 2-fold:
(11) Linic, S.; Christopher, P.; Ingram, D. B. Nat. Mater. 2011, 10,
first, Cu atoms with small coordination numbers, that is, 911.
<3, provide states that extend the edge of the valence (12) Lee, J. H.; Hevia, D. F.; Selloni, A. Phys. Rev. Lett. 2013, 110,
band into the gap by as much as ∼0.7 eV, mainly through 016101.
their 3d states. Second, two kinds of midgap states (13) Seriani, N.; Pinilla, C.; Cereda, S.; De Vita, A.; Scandolo, S. J.
appear, due to (i) a Cu−Ti bond and (ii) the formation Phys. Chem. C 2012, 116, 11062.
of a polaronic Ti3+ state in the vicinity of the copper (14) Varghese, O. K.; Paulose, M.; LaTempa, T. J.; Grimes, C. A.
Nano Lett. 2009, 9, 731.
cluster by charge donated by the cluster. We expect that (15) Roy, S. C.; Varghese, O. K.; Paulose, M.; Grimes, C. A. ACS
these features will enhance the photoresponse of the Nano 2010, 4, 1259.
TiO2 in the visible light spectrum. (16) Shankar, K.; Basham, J. I.; Allam, N. K.; Varghese, O. K.; Mor,
(c) The formation of copper-induced Ti3+ states might be G. K.; Feng, X.; Paulose, M.; Seabold, J. A.; Choi, K.-S.; Grimes, C. A.
the main contribution of copper to the enhancement of J. Phys. Chem. C 2009, 113, 6327.
the photocatalytic activity of titania. (17) Montini, T.; Gombac, V.; Sordelli, L.; Delgado, J. J.; Chen, X.-
W.; Adami, G.; Fornasiero, P. ChemCatChem 2011, 3, 574.
(d) The electronic structure of the system with a Cu4 cluster (18) Cargnello, M.; Gasparotto, A.; Gombac, V.; Montini, T.;
is different from those with smaller clusters. The Cu4 Barreca, D.; Fornasiero, P. Eur. J. Inorg. Chem. 2011, 28, 4309.
cluster has an empty state at the bottom of the (19) Barreca, D.; Carraro, G.; Gombac, V.; Gasparotto, A.; Maccato,
conduction band (CB) and this state becomes full C.; Fornasiero, P.; Tondello, E. Adv. Funct. Mater. 2011, 21, 2611.
when an additional electron is added to the system. This (20) Gombac, V.; Sordelli, L.; Montini, T.; Delgado, J. J.; Adamski,
behavior is in agreement with the expectation that A.; Adami, G.; Cargnello, M.; Bernal, S.; Fornasiero, P. J. Phys. Chem. A
2010, 114, 3916.
photoexcited electrons are transferred from titania to
(21) Yoong, L. S.; Chong, F. K.; Dutta, B. K. Energy 2009, 34, 1652.
metal clusters at the surface. On the contrary, in clusters (22) Lopez, R.; Gomez, R.; Llanos, M. E. Catal. Today 2009, 148,
with less than four atoms, the empty states are well above 103.
the bottom of the CB, and therefore, it seems probable (23) Tang, W.; Peterson, A. A.; Varela, A. S.; Jovanov, Z. P.; Bech, L.;
that photoexcited electrons will not be transferred to the Durand, W. J.; Dahl, S.; Nørskov, J. K.; Chorkendorff, I. Phys. Chem.
cluster. Chem. Phys. 2012, 14, 76.
(e) A hole added to the system prefers to sit on the Ti3+ (24) Li, C. W.; Ciston, J.; Kanan, M. W. Nature 2014, 508, 504.
(25) Liu, L.; Gao, F.; Zhao, H.; Li, Y. Appl. Catal. B: Environ. 2013,
polaron site near the titania−copper interface, leading to 134−135, 349.
the disappearance of the polaron itself. Together with the (26) Liu, D.; Fernandez, Y.; Ola, O.; Mackintosh, S.; Maroto-Valer,
previous point, this suggests that copper actively M.; Parlett, C. M. A.; Lee, A. F.; Wu, J. C. S. Catal. Commun. 2012, 25,
contributes to separate photogenerated electrons and 78.
holes. (27) Liu, L.; Li, Y. Aerosol Air Qual. Res. 2013, 14, 453.


(28) Henderson, M. A.; Epling, W. S.; Peden, C. H. F.; Perkins, C. L.
J. Phys. Chem. B 2003, 107, 534.
ASSOCIATED CONTENT (29) Kurtz, R. L.; Stock-Bauer, R.; Msdey, T. E.; Roman, E.; De
Segovia, J. L. Surf. Sci. 1989, 218, 178.
*
S Supporting Information
(30) Setvin, M.; Franchini, C.; Hao, X.; Schmid, M.; Janotti, A.;
Supporting figures. This material is available free of charge via Kaltak, M.; van de Walle, C. G.; Kresse, G.; Diebold, U. Phys. Rev. Lett.
the Internet at http://pubs.acs.org. 2014, 113, 086402.

6701 DOI: 10.1021/acs.jpcc.5b00846


J. Phys. Chem. C 2015, 119, 6696−6702
The Journal of Physical Chemistry C Article

(31) Wendt, S.; Sprunger, P. T.; Lira, E.; Madsen, G. K. H.; Li, Z.; (66) Nunzi, F.; Mosconi, E.; Storchi, L.; Ronca, E.; Selloni, A.;
Hansen, J. Ø; Matthiesen, J.; Blekinge-Rasmussen, A.; Lægsgaard, E.; Grätzel, M.; de Angelis, F. Eng. Environ. Sci. 2013, 6, 1221.
Hammer, B.; Besenbacher, F. Science 2008, 320, 1755. (67) Zhang, J.; Zhang, M.; Han, Y.; Li, W.; Meng, X.; Zong, B. J. Phys.
(32) Bennett, R. A. PhysChemComm 2000, 3, 9. Chem. C 2008, 112, 19506.
(33) Pang, C. L.; Lindsay, R.; Thornton, G. Chem. Soc. Rev. 2008, 37, (68) Mazheika, A. S.; Bredow, T.; Matulis, V. E.; Ivashkevich, O. A. J.
2328. Phys. Chem. C 2011, 115, 17368.
(34) Zuo, F.; Wang, L.; Wu, T.; Zhang, Z.; Borchardt, D.; Feng, P. J. (69) Ferrando, R.; Rossi, G.; Levi, A. C.; Kuntova, Z.; Nita, F.; Jelea,
Am. Chem. Soc. 2010, 132, 11856. A.; Mottet, C.; Barcaro, G.; Fortunelli, A.; Goniakowski, J. J. Chem.
(35) Tung, R. T. Appl. Phys. Rev. 2014, 1, 011304. Phys. 2009, 130, 174702.
(36) Park, J. B.; Graciani, J.; Evans, J.; Stacchiola, D.; Senanayke, S. (70) Seriani, N.; Mittendorfer, F.; Kresse, G. J. Chem. Phys. 2010,
D.; Barrio, L.; Liu, P.; Sanz, J. F.; Hrbek, J.; Rodriguez, J. A. J. Am. 132, 024711.
Chem. Soc. 2010, 132, 356. (71) Paz-Borbon, L. O.; Barcaro, G.; Fortunelli, A.; Levchenko, S. V.
(37) Zhou, J.; Kang, Y. C.; Ma, S.; Chen, D. A. Surf. Sci. 2004, 562, Phys. Rev. B 2012, 85, 155409.
113. (72) Heard, C. J.; Heiles, S.; Vajda, S.; Johnston, R. L. Nanoscale
(38) Chun, W.-J.; Koike, Y.; Ijima, K.; Fujikawa, K.; Ashima, H.; 2014, 6, 11777.
Nomura, M.; Iwasawa, Y.; Asakura, K. Chem. Phys. Lett. 2007, 433, (73) Giordano, L.; Pacchioni, G.; Bredow, T.; Sanz, J. F. Surf. Sci.
345. 2001, 471, 21.
(39) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, (74) Jiang, M.; Zeng, Q.; Zhang, T.; Yang, M.; Jackson, K. A. J. Chem.
Phys. 2012, 136, 104501.
3865.
(75) Bader, R. F. W. Atoms in Molecules: A Quantum Theory; Oxford
(40) Anisimov, V. I.; Korotin, M. A.; Mylnikova, A. S.; Kozhevnikov,
University Press: New York, 1990.
A. V.; Korotin, D. M.; Lorenzana, J. Phys. Rev. B 2004, 70, 172501.
(76) Seriani, N. J. Phys.: Condens. Matter 2010, 22, 255502.
(41) Di Valentin, C.; Pacchioni, G.; Selloni, A. Phys. Rev. Lett. 2006,
(77) Resta, R. Nature 2008, 453, 735.
97, 166803. (78) Seriani, N.; Jin, Z.; Pompe, W.; Ciacchi, L. Colombi. Phys. Rev. B
(42) Morgan, B. J.; Watson, G. W. Surf. Sci. 2007, 601, 5034. 2007, 76, 155421.
(43) Camellone, M. F.; Kowalski, P. M.; Marx, D. Phys. Rev. B 2011, (79) Dianat, A.; Seriani, N.; Bobeth, M.; Cuniberti, G. J. Mater. Chem.
84, 035413. A 2013, 1, 9273.
(44) Finazzi, E.; di Valentin, C.; Pacchioni, G.; Selloni, A. J. Chem. (80) Löwdin, P.-O. Phys. Rev. 1955, 97, 1474.
Phys. 2008, 129, 154113. (81) Seriani, N. J. Phys. Chem. C 2012, 116, 22974.
(45) Deák, P.; Aradi, B.; Frauenheim, T. Phys. Rev. B 2011, 83, (82) Islam, M. S.; Fisher, C. A. J. Chem. Soc. Rev. 2014, 43, 185.
155207. (83) Giordano, L.; Pacchioni, G.; Bredow, T.; Sanz, J. F. Surf. Sci.
(46) Deskins, N. A.; Dupuis, M. Phys. Rev. B 2007, 75, 195212. 2001, 471, 21.
(47) Shibuya, T.; Yasuoka, K.; Mirbt, S.; Sanyal, B. J. Phys. Chem. C (84) Bredow, T.; Aprá, E.; Catti, M.; Pacchioni, G. Surf. Sci. 1998,
2014, 118, 9429. 418, 150.
(48) Finazzi, E.; Di Valentin, C.; Pacchioni, G. J. Phys. Chem. C 2009, (85) Mellan, T. A.; Maenetja, K. P.; Ngoepe, P. E.; Woodley, S. M.;
113, 3382. Catlow, C. R. A.; Grau-Crespo, R. J. Mater. Chem. A 2013, 1, 14879.
(49) Deskins, N. A.; Rousseau, R.; Dupuis, M. J. Phys. Chem. C 2009, (86) Crespo, Y.; Seriani, N. J. Mater. Chem. A 2014, 2, 16538.
113, 14583. (87) Henrich, V. E.; Dresselhaus, G.; Zeiger, H. J. Phys. Rev. Lett.
(50) Pan, H.; Gu, B.; Zhang, Z. J. Chem. Theory Comput. 2009, 5, 1976, 36, 1335.
3074. (88) Diebold, U. Surf. Sci. Rep. 2003, 48, 53.
(51) Janotti, A.; Varley, J. B.; Rinke, P.; Umezawa, N.; Kresse, G.; van (89) Ganduglia-Pirovano, M. V.; Hofmann, A.; Sauer, J. Surf. Sci. Rep.
de Walle, C. G. Phys. Rev. B 2010, 81, 085212. 2007, 62, 219.
(52) Deák, P.; Kullgren, J.; Frauenheim, T. Phys. Status Solidi R 2014, (90) Berger, T.; Sterrer, M.; Diwald, O.; Knözinger, E.; Panayotov,
8, 583. D.; Thompson, T. L.; Yates, J. T. J. Phys. Chem. B 2005, 109, 6061.
(53) Kowalski, P. M.; Camellone, M. F.; Nair, N. N.; Meyer, B.; (91) Wang, Y.; Sun, H.-J.; Tan, S.-J.; Feng, H.; Cheng, Z.-W.; Zhao,
Marx, D. Phys. Rev. Lett. 2010, 105, 146405. J.; Zhao, A.; Wang, B.; Luo, Y.; Yang, J.-L.; Hou, J.-G. Nat. Commun.
(54) Li, S.; Jena, P. Phys. Rev. B 2009, 79, 201204(R). 2013, 4, 2214.
(55) Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; (92) Serpone, N. J. Phys. Chem. B 2006, 110, 24287.
Cavazzoni, C.; Ceresoli, D.; Chiarotti, G. L.; Cococcioni, M.; Dabo, I.; (93) Nowotny, J.; Bak, T.; Nowotny, M. K.; Sheppard, L. R. J. Phys.
Dal Corso, A.; de Gironcoli, S.; Fabris, S.; Fratesi, G.; Gebauer, R.; Chem. B 2006, 110, 18492.
Gerstmann, U.; Gougoussis, C.; Kokalj, A.; Lazzeri, M.; Martin-Samos, (94) Lin, Z.; Orlov, A.; Lambert, R. M.; Payne, M. C. J. Phys. Chem. B
L.; Marzari, N.; Mauri, F.; Mazzarello, R.; Paolini, S.; Pasquarello, A.; 2005, 109, 20948.
Paulatto, L.; Sbraccia, C.; Scandolo, S.; Sclauzero, G.; Seitsonen, A. P.; (95) Xing, M.; Fang, W.; Nasir, M.; Ma, Y.; Zhang, J.; Anpo, M. J.
Smogunov, A.; Umari, P.; Wentzcovitch, R. M. J. Phys.: Condens. Catal. 2013, 297, 236.
Matter 2009, 21, 395502. (96) Hamdy, M. S.; Amrollahi, R.; Mul, G. ACS Catal. 2012, 2, 2641.
(56) Scandolo, S.; Giannozzi, P.; Cavazzoni, C.; de Gironcoli, S.;
Pasquarello, A.; Baroni, S. Z. Kristallogr. 2005, 220, 574.
(57) Vanderbilt, D. Phys. Rev. B 1990, 41, 7892.
(58) Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.;
Sutton, A. P. Phys. Rev. B 1998, 57, 1505.
(59) Cococcioni, M.; De Gironcoli, S. Phys. Rev. B 2005, 71, 035105.
(60) Scanlon, D. O.; Walsh, A.; Morgan, B. J.; Watson, G. W.; Payne,
D. J.; Egdell, R. G. Phys. Rev. B 2009, 79, 035101.
(61) Monkhorst, H. J.; Pack, J. D. Phys. Rev. B 1976, 13, 5188.
(62) Li, G.; Gray, K. A. Chem. Phys. 2007, 339, 173.
(63) Lazzeri, M.; Vittadini, A.; Selloni, A. Phys. Rev. B 2001, 63,
155409.
(64) Cheng, H.; Selloni, A. Phys. Rev. B 2009, 79, 092101.
(65) Balducci, G. Chem. Phys. Lett. 2010, 494, 54.

6702 DOI: 10.1021/acs.jpcc.5b00846


J. Phys. Chem. C 2015, 119, 6696−6702

You might also like