• Non ci sono risultati.

N-doped TiO2

N/A
N/A
Protected

Academic year: 2022

Condividi "N-doped TiO2"

Copied!
13
0
0

Testo completo

(1)

N-doped TiO

2

: Theory and experiment

Cristiana Di Valentin

a,*

, Emanuele Finazzi

a

, Gianfranco Pacchioni

a

, Annabella Selloni

b

, Stefano Livraghi

c

, Maria Cristina Paganini

c

, Elio Giamello

c

aDipartimento di Scienza dei Materiali, Universita` di Milano-Bicocca, Via R. Cozzi, 53, 20125 Milano, Italy

bDepartment of Chemistry, Princeton University, Princeton, NJ 08540, USA

cDipartimento di Chimica IFM, Universita` di Torino and NIS, Nanostructured Interfaces and Surfaces Centre of Excellence, Via P. Giuria 7, I-10125 Torino, Italy

Received 16 April 2007; accepted 10 July 2007 Available online 27 July 2007

Abstract

Nitrogen doped titanium dioxide is attracting a continuously increasing attention because of its potential as material for environmen- tal photocatalysis. In this paper we review experimental and theoretical work done on this system in our groups in recent years. The analysis is largely based on electron paramagnetic resonance (EPR) spectra and on their interpretation based on high-level ab initio cal- culations. N-doped anatase TiO2contains thermally stable single N-atom impurities either as charged diamagnetic Nb centers or as neu- tral paramagnetic Nbcenters (b stays for bulk). The N-atoms can occupy both interstitial or substitutional positions in the solid, with some evidence for a preference for interstitial sites. All types of Nbcenters give rise to localized states in the band-gap of the oxide, thus accounting for the related reduction of absorption band edge. The relative abundance of these species depends on the oxidation state of the solid. In fact, upon reduction, oxygen vacancies form and transfer electrons from Ti3+ions to the Nbwith formation of Ti4+and Nb. EPR spectra measured under irradiation show that the Nbcenters are responsible for visible light absorption with promotion of electrons from the localized N-impurity states to the conduction band or to electron scavengers like O2adsorbed on the surface. These results provide an unambiguous characterization of the electronic states associated with N-impurities in TiO2and a realistic picture of the pro- cesses occurring in the solid under irradiation with visible light.

 2007 Elsevier B.V. All rights reserved.

Keywords: TiO2; Doping; Nitrogen; EPR; Experiments; DFT; Theory; Oxygen vacancy; Photocatalysis; Vis-light

1. Introduction

Worldwide concerns with environmental and energy- related issues have prompted an enormous interest in semi- conductor-based heterogeneous photocatalysis over the last decade. Photocatalysis allows the use of sunlight for the destruction of highly toxic molecules and remediation of pollutants; for the selective, synthetically useful redox transformations in specific organic compounds; for the production of hydrogen, and the conversion of solar energy to electric power[1–6].

The most widely used material in heterogeneous photo- catalysis is titanium dioxide (TiO2). It is inexpensive, non- toxic, resistant to photo-corrosion, and has high oxidative power. However, an important drawback of TiO2for pho- tocatalysis is that its band-gap is rather large, 3.0–3.2 eV, and thus only a small fraction of the solar spectrum (k < 380 nm, corresponding to the UV region) is absorbed.

To lower the threshold energy for photoexcitation, a great deal of research has focused on doping TiO2 with both transition metal and non-metal impurities. Doping with transition metals has shown both positive and negative effects. Indeed, several authors have reported that although metal ion doping decreases the photothreshold energy of TiO2, the metal ions may also serve as recombination

0301-0104/$ - see front matter  2007 Elsevier B.V. All rights reserved.

doi:10.1016/j.chemphys.2007.07.020

* Corresponding author. Tel.: +39 0264485235.

E-mail address:cristiana.divalentin@mater.unimib.it(C. Di Valentin).

www.elsevier.com/locate/chemphys

(2)

centers for electrons and holes, thus reducing the overall activity of the photocatalyst.

Stimulated by the report of Asahi et al. in 2001 [7], recently there has been an explosion of interest in TiO2

doping with non-metal ions, especially with nitrogen (see Fig. 1, from ISI). New papers on this topic appear every week, showing the importance of the subject but also mak- ing it difficult to keep track of the rapid developments.

Unfortunately, not all these studies are as innovative and original as one would like to see, sometimes contributing to the confusion more than to the understanding of the problem. Many authors have reported that N-doped tita- nium dioxide, hereafter N-TiO2, shows a significant cata- lytic activity in various reactions performed under visible light irradiation[8–19]. However, as summarized in a num- ber of very recent reviews [4–6] (to which we refer for a more exhaustive account), there is also an open debate on how doping achieves this, as well as disagreements in many of the conclusions drawn from the results. One of the reasons for this may be that, to incorporate nitrogen in titanium dioxide, different strategies are used, either based on chemical reactivity (sol–gel synthesis[8,9,13–15], chemical treatments of the bare oxide[7,10,11,18], oxida- tion of titanium nitride[20], etc.) or on physical methods (ion implantation [21–23], magnetron sputtering [24,25]).

These different procedures probably lead, at least in some cases, to materials with somewhat different properties.

Independent of the preparation procedure, there are a number of key questions which need to be answered to characterize the properties and behavior of N-TiO2. The first and most basic question concerns the chemical nature, the location in the solid, and the involvement in photo- activity of the nitrogen species. Different chemical species like NOx [8,9,15,26–28] substitutional N [7,12,14,21], or NHx [11] have been proposed. It is essential to know whether the species are primarily interstitial or substitu- tional, because the behavior of these two species is very dif- ferent and will affect the material properties accordingly. It is also important to determine whether the species tend to

segregate at the surface of the material or are preferentially incorporated in sub-surface or bulk sites, because this can strongly influence the surface reactivity and photocatalytic properties.

A second key issue concerns the electronic structure of the doped material and the change of the optical gap [4,7,29]. Many studies have addressed the question of the character, localized or delocalized, of the electronic states associated to the N-impurities, but no general consensus has been reached yet. While some authors claim that the band gap of the solid is reduced due to a rigid valence band shift upon doping [7,20,25], others attribute the observed absorption of visible light by N-TiO2 to the excitation of electrons from localized impurity states in the band-gap [9,10,12,15,16]. Interestingly, it appears that the N-doping induced modifications of the electronic structure may be slightly different for the anatase and rutile polymorphs of TiO2. In fact, while it is well established that N-doping low- ers the onset of optical absorption to the visible in anatase, a slight increase of the optical gap has been in some cases observed for rutile [21].

Another crucial question concerns the interaction between N-impurities and oxygen vacancies. Density func- tional theory (DFT) calculations by our group [28] have shown a large decrease in the formation energy for oxygen vacancies as a result of the presence of nitrogen atoms in the lattice. This decrease can be attributed to a ‘compensa- tion’ effect: electrons in vacancy related states near the bot- tom of the conduction band can be transferred to empty nitrogen impurity states near the top of the valence band, leading to a considerable energy gain. The large decrease in the oxygen vacancy formation energy also suggests that oxygen vacancies are most probably induced by N-doping of TiO2. Recent STM [22] and UPS [30] data appear to support these predictions, but a detailed understanding of the structure and electronic characteristics of the vacancy–impurity complex is still missing. It is important to determine the energetics and the character of the elec- tronic states of these complexes, e.g. how the occupied states of the vacancy–impurity complex are distributed, whether they are spread throughout the complex, and how close should the vacancy and the impurities be in order to form such a charge transfer complex.

The last and probably most important question for pho- tocatalysis is the effect of N-doping on the photocatalytic activity of TiO2. A large part of the existing literature on N-doped TiO2materials agrees that the addition of nitro- gen results in an improvement in visible light photocata- lytic activity, but contrasting findings have also been reported [4]. For instance, in a recent study on N-TiO2

films obtained by the addition of ammonia during chemical vapor deposition growth of TiO2[5], no evidence of photo- catalytic activity in the visible was found, in spite of clear evidence of the presence of N impurities at substitutional sites.

Significant advances towards a clarification of the above issues can only be achieved via multi-pronged efforts based

Fig. 1. Number of published articles on X-doped TiO2(X = N, C, S, F).

From ISI (February 26, 2007).

(3)

on the use of different experimental techniques together with theoretical calculations. Our group has used a combi- nation of electron paramagnetic resonance (EPR), X ray photoelectron spectroscopy (XPS), and DFT calculations to characterize the paramagnetic species in N-doped ana- tase TiO2powders obtained by sol–gel synthesis. The pur- pose of this paper is to present a critical review of this work together and some recent new results, such as the analysis of the EPR feature intensity during some red–ox cycles per- formed on the samples and the theoretical simulation of the N-doped anatase (1 0 1) surface.

2. Experimental and computational details 2.1. Sample preparation

N-TiO2 samples have been prepared via the sol–gel method using several nitrogen containing inorganic com- pound (NH4Cl, NH3, N2H4, NH4NO3 and HNO3) as nitrogen source. When necessary15N enriched compounds were employed. A solution of titanium (IV) isopropoxide in isopropylic alcohol was mixed with an aqueous solution of a nitrogen compound and kept upon constant stirring at room temperature (RT). The gel so obtained has been left ageing overnight at RT to ensure the completion of the hydrolysis and subsequently dried at 343 K. The dried compound was heated at 773 K in air for 1 h. The best results were obtained using ammonium chloride as nitro- gen source.

The calcination influences the final properties of the material depending on the temperature and heating rate employed in the treatment. After heating in air at 773 K and at a relatively slow heating rate (5 K/min) the final material exhibits a pale yellow colour and has the anatase structure. We will hereafter report results concerning this material which has shown the best catalytic performance among all materials examined.

2.2. Computational set-up

The computations have been done at the DFT level using the plane-wave-pseudopotential approach, together with the Perdew–Burke–Ernzerhof (PBE) [31] exchange- correlation functional, and ultrasoft pseudopotentials [32]

(with kinetic energy cut-offs of 25 ad 200 Ry for the smooth part of the electron wavefunctions and augmented electron density, respectively). We considered nearly cubic 2p

2· 2p

2· 1 and 2 · 2 · 3 supercells to model anatase and rutile, respectively. The optimized bulk lattice parameters, taken from previous calculations [33], are a = 3.786 A˚ , and c = 9.737 A˚ for anatase, and a = 4.634 A˚, and c = 2.963 A˚ for rutile. Two different codes within the Quantum ESPRESSO package [34]were employed: CP90, based on the Car-Parrinello (CP) approach [35,36], which allows for efficient structural optimizations with k-point sampling restricted to C, and PWscf, which was used for electronic structure calculations at a low-symmetry special k point

(hereafter denoted Kls). With CP90, all atoms of the super- cells were relaxed using a damped second-order dynamics until all components of the residual forces were less than 0.025 eV/A˚ . In the case of rutile, optimizations at Klshave been performed with the Broyden–Fletcher–Goldfarb–

Shanno technique.

Substitutional N-species (Ns) were modelled by replac- ing 1, 2, or 3 oxygen atoms in the 96-atoms anatase super- cell and 1 or 2 oxygens in the 72-atoms rutile supercell. The resulting stoichiometry is TiO2xNxwith 0.031 < x < 0.094 for anatase and 0.042 < x < 0.084 for rutile. These N con- centrations are comparable to those used in experiments.

The procedure of including more N-atoms in the same supercell is more accurate than using different smaller supercells as it allows a direct comparison of the various levels of doping on the band structure of the material.

Interstitial N-species (Ni) were modeled by adding one N-atom in the supercell. In the following, the paramagnetic species will be denoted as Nb, where b stands for bulk, when the distinction between Ni and Ns is irrelevant for the discussion. To study oxygen vacancy formation in pres- ence of N-impurities in anatase, one oxygen atom was removed while either two other oxygens were replaced with N-atoms or two N-atoms were added in interstitial posi- tions. The resulting stoichiometry is TiO23xN2x or TiO2xN2x, respectively, with x = 0.031. The supercell models are overall neutral despite the fact that the isolated defects may be charged (as described below) in order to reflect the experimental situation.

In order to compute atomic spin properties, such as hyperfine coupling constants of the electron spin with the nuclear spin of 14N and atomic spin densities, we used the CRYSTAL03 package[37], based on an atomic orbital basis set, together with the Becke-3 [38] and Lee–Yang–

Parr [39] (B3LYP) exchange and correlation functional.

Hybrid functionals are indeed known to give more accurate descriptions of spin-polarized systems [40,41]. We per- formed single point B3LYP calculations using the C point and the geometries given by the PBE calculations. The atoms were described with the following basis sets: Ti 86411(d41)[42], O 8411(d1) [43], N 7311(d1) [44], respec- tively. The hyperfine spin-Hamiltonian, Hhfc= S Æ A Æ I, is given in terms of the hyperfine matrix A which describes the coupling of the electron with the nuclear spin [45].

The A matrix can be decomposed in two parts: Aiso(isotro- pic component of the hyperfine coupling constants) and B (dipolar component of the hyperfine coupling constants).

3. Results and discussion

3.1. Catalytic and spectroscopic characterization

The anatase N-TiO2materials were compared with bare TiO2materials prepared and treated with the same proce- dure. The photocatalytic activity of N-TiO2in the reference reaction of methylene blue degradation performed under visible light (wavelength 437 nm) is definitely (though not

(4)

largely) higher than that of the bare oxide in the same con- ditions. This enhanced photocatalytic activity is in agree- ment with results previously reported in the literature.

N-TiO2were examined with a variety of spectroscopic techniques including infra-red spectroscopy, MAS-NMR, DR-UV–Vis spectroscopy, XPS and EPR. The most signif- icant results are those obtained by the latter three tech- niques and in particular by EPR.

The optical absorption of N-TiO2 materials is that of bare TiO2modified by an absorption in the visible region centred at about 430–450 nm. The intensity and width of this absorption depends on the preparation history of the material. In particular, for the N-TiO2material here con- sidered, the DR-UV–Vis spectrum,Fig. 2, is close to that first published by Sato[8]in 1986 and explains the pale yel- low colour of the solid.

3.1.1. EPR spectroscopy

An important turning point of our investigation on N- TiO2 is related to the observation that the sample is rich of paramagnetic centers [8,46]clearly detectable by EPR.

Due to its high sensitivity and to its capability of an accu- rate description of paramagnetic centers this latter tech- nique is particularly suited for the investigation of defects in the solid state. The analysis of a long series of EPR spec- tra performed on samples of N-TiO2prepared and treated in different ways has allowed us to identify two types of paramagnetic species, denoted as I and II. Type I consists of molecular nitrogen oxide species, NO and NO2, trapped in microvoids of the solid, provided the solid itself has a porous texture[47],Fig. 3.

Nitric oxide (NO) species is a product of the complex oxidation process of ammonium salts occurring upon calci- nations of the solid. NO2, on the contrary, is formed only when nitrates or nitric acid are used as nitrogen source and can be thus considered to derive from their decomposition.

The experimental and simulated EPR spectra of NO species are shown in Fig. 4, respectively obtained with

14NO and with a 15NO enriched (70%) mixture at 77 K.

At RT the EPR spectrum of this species is not observed.

The isotopic substitution was performed to unambiguously assign the EPR spectrum whose parameters are those expected for a NO molecule adsorbed on a positively charged site. The numerical values of the parameters are reported inTable 1.

The nitric oxide molecule is observed by EPR in the region of the free electron g value only when adsorbed on a surface[48,49]. When NO is in the adsorbed state, in fact, two effects allow the observation of an EPR spectrum. The former one is the orbital momentum quenching by the elec- tric field of the adsorption site which causes the2P1/2state of the gaseous molecule (complete coupling between the electron spin S and the angular momentum L) to become a paramagnetic, pure spin state. The second is the lifting of the degeneracy of the two p* orbitals producing an energy splitting which determines the gzz value of the adsorbed species. Since the NO spectrum is observed at low temperature only it was concluded that it must be a

300 450 600 750

Green

Blue

Reflectance %

b a

λ (nm)

Fig. 2. UV–Vis spectra of (a) bare TiO2, (b) N-TiO2.

Fig. 3. Schematic representation of NO molecules trapped in microvoids of N-TiO2crystal.

3300 3375 3450 3525

b a'

15N (70%)

14N

g3 g2

g1

b' a

B/Gauss

Fig. 4. EPR spectra of NO species: (a, a0) experimental and simulated spectra obtained with14N compounds, (b, b0) experimental and simulated spectra obtained with15N enriched (70%) compounds.

Table 1

EPR parameters of the three different paramagnetic species identified in N-TiO2

Species g1 g2 g3 A1(G) A2(G) A3(G)

NO 2.001 1.998 1.927 < 1 32.2 9.6

NO2 2.004 2.001 1.990 53.3 66.3 50.1

Nb 2.005 2.004 2.003 2.3 4.4 32.3

(5)

weakly adsorbed species, easily desorbed by raising the temperature. Since the EPR spectrum reversibly shows up upon lowering the temperature and the species is not removed upon pumping off up to 773 K, it was concluded that the species itself is trapped in a cavity of the bulk, adsorbed at low temperature on the walls of the microv- oids,Fig. 3.

A similar effect was observed for trapped NO2, see the spectrum ofFig. 5.

The highest concentration of NO2 was observed when nitric acid is used as a source of nitrogen but the lines of this species also appear in the case of samples prepared by NH4Cl and reoxidized at high temperature (773 K). In this case NO2lines are overlapped with those of NO. The generation of NO2and its segregation in bulk microvoids like for the NO molecule, is a result of the decomposition of HNO3during calcination:

2HNO3! 2NO2þ H2Oþ 1=2O2 ð1Þ

When NO and NO2species are simultaneously present in the system they are not independent as NO2concentration increases, upon thermal treatment, with decreasing that of NO. However, as the respective amounts of the two species are different (NO2concentration is two order of magnitude lower than that of NO) the reaction occurring is somehow complex and likely involves the formation of diamagnetic intermediates like nitrites and nitrates formed via reaction with oxygen ions of the solid. Due to their nature of trapped species, it is easy to conclude that both NO and NO2are by products of the nitrogen incorporation in the solid and do not directly influence the electronic structure of the system.

Much more interesting is the paramagnetic nitrogen spe- cies of type II, which always forms upon calcination of the system after gelification. The EPR signal of such Nbspecies is characterized by an orthorhombic g tensor whose main values are, however, very close one to the other [28]. For this reason the species has been investigated by a multifre- quency approach. InFig. 6the EPR spectra of Nb species measured at 9.5 GHz (X-band) and 35 GHz (Q-band), respectively are shown.

Both the g tensor and the hyperfine tensor are reported inTable 1. The hyperfine tensor is based on a quite large coupling constant (±32.3 G) in the direction of the g3ele- ment and two smaller constants in the other directions (A1= ±2.3 G and A2= ±4.4 G). The same set of spin Hamiltonian parameters was used to simulate both the X-band and the Q-band spectra obtaining a satisfactory fit between experimental and simulated traces in the two cases (Fig. 6). Nbspecies contains one N-atom which bears a total spin density of about 0.54, the larger contribution being due to a single p orbital of the nitrogen atom in the centre. The previous analysis indicates that the hyper- fine tensor does not account for the whole unpaired elec- tron spin density. In most cases of molecular radical centres the unaccounted spin density is usually localized on other atoms of the species having zero nuclear spin.

On the basis of the EPR spectra recorded in various condi- tions it was possible to understand, first of all, that Nbis a monomeric nitrogen species incorporated in the bulk of N- TiO2 as the species does not undergo dipolar broadening when oxygen is adsorbed at low temperature on the sur- face. The species is stable even at rather high temperature and, as shown in the following, sensitive to visible light irradiation.

3.1.2. X ray photoelectron spectroscopy (XPS)

InFig. 7the XPS spectrum of N-TiO2(in the region of binding energy typical of N(1s)) is reported. The signal deconvolution allows the identification of two peaks at 399.2 eV and 400.7 eV.

After a careful washing of the material in ultrasonic, the XPS peaks tend to decrease in intensity or in some cases disappear. XPS peaks in the range (396–404 eV) were observed by several authors and are typical of nitrogen

3250 3300 3350 3400 3450 3500 3550 a'

a

B/Gauss

Fig. 5. (a) EPR spectrum of NO2species and (a0) its simulation.

3320 3340 3360 3380

12040 12060 12080 12100

c' c b' b a'

a 14N + 15N

14N

14N

(X-Band) B/Gauss

(Q-B and) B/Gauss

Fig. 6. EPR spectra and related simulations of species Nb: (a) X-band spectrum of the species containing 15N (70%) and 14N, (b) X-band spectrum of the species containing14N, (c) Q-band spectrum of the species containing14N. a0, b0 and c0 are the related simulations (figure adapted from Ref.[46]).

(6)

doped titanium dioxide [16,26]. The assignment of XPS features in N-TiO2, however is still under debate and con- troversial hypotheses are found in the literature. Peaks at 396–397 eV were attributed to substitutional nitrogen because of the proximity to the typical binding energy of 396 eV observed for N in TiN, while peaks at higher bind- ing energies (400 eV like those found in our case) were observed and usually ascribed to a generic interstitial site.

In the case of the sample here described and prepared via sol–gel, peaks at 396 eV were never observed.

The abovementioned reduction or elimination of the XPS peaks observed after a careful washing of the material suggests a relation between the peak at 400 eV and the residual ammonium ions which are indeed easily removed from the surface by washing. Being Nb species unambigu- ously located in bulk a doubt arises regarding a clear-cut correlation of the XPS nitrogen peaks in N-TiO2with this species. A variety of species including Nb, and not a unique one, is probably responsible for the XPS features.

3.2. Theoretical models of N-TiO2

We have used DFT calculations to study two models capable of explaining the observed properties of the nitro- gen species, namely substitutional (Ns) and interstitial (Ni) nitrogen impurities in the anatase bulk structure,Fig. 8. In the substitutional model, the nitrogen atom is bound to three Ti atoms and replaces a lattice oxygen in TiO2. The Ti–N bond lengths, 1.964 and 2.081 A˚ , are only slightly longer than the Ti–O ones, 1.942 and 2.002 A˚ . This nitro- gen atom is in a negative oxidation state. In the lowest energy interstitial model, the nitrogen atom is bound to one lattice oxygen and therefore is in a positive oxidation state. The resulting NO species (d(N–O) = 1.36 A˚ ) interacts with the lattice Ti atoms through its p bonding states (Fig. 8).

Electronic structure calculations indicate that for both Nsand Nithere is virtually no shift of the upper and lower edges of the O 2p valence band, as well as of the conduc-

tion band, with respect to the undoped material. For sub- stitutional N, a few occupied N 2p localized states appear slightly above the valence band edge (Fig. 9). For intersti- tial N, the NO bond generates localized states with p char- acter (Fig. 9). The two bonding states are deep in energy and lie below the top of O 2p band. The two antibonding (but still occupied) states lie above the O 2p band and are higher in the gap than the N 2p states generated by sub- stitutional nitrogen: the highest localized state for the inter- stitial species is 0.73 eV above the top of the valence band, while for the substitutional species it is 0.14 eV above it, Fig. 9. The computed band-gap for these systems is 2.6 eV, obtained as the difference between conduction and valence band energy eigenvalues (band gaps are noto- riously underestimated by standard DFT methods, such as the present PBE functional) [28]. Thus, the conclusion is that both Nsand Niinduce formation of localized occupied states in the gap, accounting for the visible light activity of N-doped samples. The localized nature of the N-induced states has the consequence that the hole generated by

Fig. 7. Deconvolution of the N-TiO2XPS spectrum. A and B indicate the two peaks used to convolute the spectrum.

Fig. 8. Model structures for substitutional and interstitial N-dopants in anatase TiO2phase (figure adapted from Ref.[28]).

VB CB

N 2p

VB CB

π* N-O

π N-O SUBSTITUTIONAL

0.14 eV 0.73 eV

2.6 eV 2.6 eV

INTERSTITIAL

Fig. 9. Electronic structure computed with PBE density functional for substitutional and interstitial models (figure adapted from Ref.[28]).

(7)

vis-irradiation is less mobile than that generated by UV irradiation in the O 2p band and has also a lower direct oxidation potential for photocatalytic applications[50,51].

Even though core level binding energy values for the N 1s state, computed within the initial state approximation, cannot be directly compared to experimental XPS data because of the absence of final state relaxation effects, the relative values of substitutional vs interstitial configura- tions are qualitatively relevant. The calculated values for these two species differ by 1.6 eV, with higher core level binding energy for the interstitial configuration. This result qualitatively correctly correlates with the two experimen- tally observed peaks for N-doped TiO2, one at 397 eV (attributed to substitutional N) and the other at >399 eV (interstitial N).

A more detailed analysis can be performed for the spin properties, by comparing the computed and measured hyperfine coupling constants (A matrix in Table 1). The agreement can be considered satisfactory for both Ns and Nispecies, even though in the latter case it is slightly better.

In the Ns case, the unpaired electron is almost entirely (0.87) localized on the substitutional N-atom, more pre- cisely on the p state perpendicular to the plane defined by the three Ti atoms to which nitrogen is bound (Fig. 10).

In the case of interstitial nitrogen, the unpaired electron is shared between the N and O atoms of the NO species and the localization on N is reduced to 0.67. Here the spin density is localized on the p system of the NO species (Fig. 10) involving only N-atomic p states. Despite the dif- ferent calculated spin densities on Nsand Ni(0.87 and 0.67, respectively), the hyperfine coupling constants are quite similar in the two cases, and rather close to the experimen- tal values as well (see Table 1). As mentioned above, the experimental estimate of the spin density on N, 0.54, obtained as the ratio of the experimentally measured B val- ues with tabulated atomic B, might not be strictly correct, so that it cannot be used to discriminate between the two model species under investigation. Therefore, the above comparative experimental–computational EPR analysis indicates that both proposed models, Ns and Ni, may account for the paramagnetic species observed in the TiO2powders.

To summarize, stable bulk N-species can be obtained by sol–gel synthesis. They give rise to localized states in the gap with paramagnetic character. Identification of the loca- tion of the N-impurities is not straightforward, but the results are more in favor of interstitial nitrogen species.

3.3. Anatase versus rutile: differences and similarities For the rutile TiO2polymorph, the structural variations following O replacement with N in the 72-atom supercell are found to be slightly more pronounced than for anatase.

The equatorial Tieq–N bonds are asymmetrically stretched, from 1.956 to 1.988 and 2.036 A˚ , with respect to the corre- sponding bonds in undoped rutile, while the axial Tiax–N bond is slightly stretched, from 1.999 to 2.021 A˚ (see Fig. 11). As for anatase, the spin density is largely localized on the N-atom. The electronic structure shows both analo- gies and differences with respect to anatase. Not only the top of the valence band, Ev, is lowered by 0.4 eV with respect to the undoped case (seeFig. 12), but also the N 2p localized states are slightly lower in energy (by E3= 0.05 eV) than Evin pure rutile,Fig. 12. Thus N-doped rutile is predicted to show a small blue-shift of 0.1 eV in the optical absorption rather than a red-shift, as observed for anatase. This conclusion is consistent with measurements by Diwald et al.[21]on N-doped rutile TiO2single crystals.

As discussed in detail in Ref.[52], the different behavior of rutile and anatase can be rationalized on the basis of their different structures and densities and of their different O 2p band width.

3.4. Role of other defects: O vacancies 3.4.1. Thermodynamic aspects

An unexpected but relevant aspect of the interaction between N-centers and oxygen vacancies (VO) is that the cost of VO formation in bulk anatase TiO2 is drastically reduced in the presence of N-impurities, from 4.2 eV in

O N O

O T i O T i

T i

O T i O T i O O

O

O N O

O T i O T i

T i

O O

O O

T i T i O O

SUBSTITUTIONAL INTERSTITIAL

Fig. 10. Electron spin density contour lines for substitutional and inter- stitial N-dopants (figure adapted from Ref.[28]).

Fig. 11. Supercell models for anatase and rutile N-doped TiO2 (figure adapted from Ref.[52]).

(8)

pure TiO2to 0.6 eV in N-doped TiO2, as indicated by the computed energetics for the following processes[28]:

TiO2! TiO2xþ xVOþ 1=2xO2 DE¼ 4:2 eV ð2Þ TiO22xN2x! TiO23xN2xþ xVOþ 1=2xO2 DE¼ 0:6 eV

ð3Þ This result is very important and explains why in oxygen poor conditions (low values of the oxygen chemical poten- tial) substitutional N impurities in the presence of VO

become very favorable, as clearly indicated by the stabil- ity diagram inFig. 13 [53]. By contrast, under oxygen-rich conditions (higher values of the oxygen chemical poten- tial), interstitial nitrogen species (Ni) become favored.

The energies of formation are evaluated according to the formula:

Eform ¼ 1=nEtot ðN-dopedÞ  1=n½Etot ðpureÞ þ nlN mlO ð4Þ

where n is the number of N-atoms and m is the number of oxygen vacancies per supercell in the model considered. We use a fixed value of the nitrogen chemical potential, lN= 1/2 lN2, while for oxygen we take lO¼ 1=2lO2þ l0O (lN2 and lO2 are the total energies of N2 and O2, respectively).

3.4.2. Electronic aspects

Experiments show that O vacancies in undoped TiO2

introduce defect states in the band-gap at energy of about 0.8 eV below the conduction band; these are assigned to Ti3+ 3d [1] states on the basis of EPR experiments [54–

56]. The theoretical description of these states is a delicate issue which we have analyzed in detail in a recent paper [41]. In particular we have shown that the electronic struc- tures of hydroxylated and reduced TiO2 are very similar and, in both systems, the localized defect states on Ti3+ions can be well described only by use of hybrid functionals.

In Fig. 14we compare the total and projected (on two selected Ti ions) density of states (DOS) for the hydroxyl- ated system calculated using (a) the B3LYP functional for both geometry optimization and DOS calculations; (b) the B3LYP functional for geometry optimization and the PBE functional for the DOS calculation; (c) the PBE functional for geometry and B3LYP for DOS calculation; (d) the PBE functional for both geometry and DOS calculations. It is clear from this picture that the PBE functional (d) underes- timates the band-gap and leads to a very delocalized solu- tion for the defects states. The band-gap problem is amended by use of the B3LYP functional with the PBE geometry, but the defect state is still delocalized (c). In order to achieve the correct structural deformation (pola- ronic distortion) for the electron localized on the Ti3+ions it is necessary to perform the atomic relaxation with the hybrid functional, as it is clear from Fig. 14a and b. In Fig. 14b, however, the band-gap description by PBE func- tional is still poor and thus, although the states are rather localized (because of the proper structural relaxation with B3LYP), they are still too close to the conduction band.

Only the full B3LYP calculation correctly describes the localization and energy position of the impurity states in the band-gap, Fig. 14a.

Unlike the case of undoped TiO2, in N-TiO2 oxygen vacancies can be satisfactorily described even without using hybrid functionals. There are two main reasons for this: (1) the position of the N-impurity states is not altered when computed with the B3LYP hybrid functional [28] which correctly describes the band-gap of TiO2; (2) when consid- ering the concomitant presence of N-impurities and oxygen vacancies, there are no excess electrons in Ti3+ states (poorly described by standard DFT), as they are trans- ferred to the empty N-states (see below). Even if the origi- nal position of the Ti3+states is not accurately predicted by standard DFT, there is no doubt that these states are higher in energy than those associated with the N-impuri- ties. Thus, internal charge transfer occurs independently of the exchange-correlation functional used.

Fig. 13. Stability diagram for the various bulk N-doped anatase models considered.

Fig. 12. Schematic electronic structure for pure and N-doped anatase and rutile TiO2 polymorphs (substitutional N) computed with the PBE functional (figure adapted from Ref.[52]).

(9)

To study N-doping in presence of oxygen vacancies, DFT calculations have been performed on a bulk anatase supercell of 96-atoms containing either two Ni or two Ns paramagnetic species plus one oxygen vacancy located as far as possible from the N-centers, so as to avoid any direct defect-impurity interaction. Due to the presence of unpaired spins at various sites, various spin configurations were considered. In particular, the high-spin configuration corresponds to four unpaired electrons (quintet state), two on the two N-impurities and two on two Ti3+ions, while a low-spin closed shell configuration (singlet state) results from the charge transfer of the two unpaired electrons occupying the high lying Ti3+ states to the low lying Ni and/or Ns singly occupied levels, just above the top of the valence band (Fig. 5). This process leads to the forma- tion of charged diamagnetic Ni and Ns defects and reox- idized Ti4+ions. The latter low-spin configuration is found to be about 3 eV more stable than the high-spin quintet one. The formation of charged defects in the material is

accompanied by a local distortion: the Ns center is charac- terized by shorter N–Ti bond lengths than Ns; Ni by a longer NO bond length than Ni.

Similar to their neutral counterparts, the diamagnetic Ni and Ns charged defects give rise to localized levels in the gap of the material which lie few tenths of an eV above the valence band edge. Because of the increased Coulomb repulsion, the N-induced defect states for the diamagnetic charged species are slightly higher in energy than the corre- sponding singly occupied levels for the paramagnetic neu- tral species: 0.59 eV for Ns and 0.75 eV for Ni ðDEd in Fig. 15), to be compared to 0.14 eV for Ns and 0.73 eV for Ni (DEp in Fig. 15). The Ns Ns shift of about 0.4 eV for substitutional nitrogen is caused by the fact that in this case the orbital where the extra electron is accom- modated is a localized atomic-like orbital; instead the shift is almost negligible for interstitial nitrogen because the added electron occupies a more extended p molecular orbi- tal. On the basis of these data, no significant variations in the absorption spectrum of the doped material are expected as a consequence of the internal charge transfer from the oxygen vacancies derived states to the N-impurity states, in particular if the species present are interstitial N-atoms, Nior Ni.

Our finding that N-doping favors O vacancy formation is fully supported by real-time transmission electron microscopy (TEM) experiments [57]. Similarly, recent experiments have found that N-implantation reduces the features associated to Ti3+ions[22]. All these data suggest that N-TiO2contains, besides a given number of paramag- netic Nb centers, a fraction of diamagnetic Nb impurities, which depends on the level of oxygen deficiency in the sam- ple. A particularly clear proof of this dependence comes from the observation that the intensity of Nb is deeply affected by the redox state of the solid.Fig. 16shows the intensity variation of the EPR spectrum upon reductive and oxidative cycles of the solid. After the very first cycle which is affected by the presence of surface hydroxyl groups, the successive cycles show that the Nb EPR inten- sity vanishes in parallel with thermovacuum reductive treatments at 773 K (in such conditions O2 depletion occurs in parallel with formation of anion vacancies and enrichment of the electron population of the solid). Restor- ing the initial situation by reoxidation at the same temper- ature, the EPR intensity is fully recovered, thus confirming the predicted connection between the presence of Nb and the redox state of the solid.

Fig. 14. Total and projected (on relevant Ti atoms) density of states for the hydroxylated TiO2(1 1 0) rutile surface obtained with B3LYP and PBE density functionals. (a) B3LYP functional for both geometry and DOS calculations; (b) B3LYP functional for geometry and PBE functional for the DOS calculation; (c) the PBE functional for geometry and B3LYP for DOS calculation; (d) the PBE functional for both geometry and DOS calculations. Vertical dashed lines indicate the highest occupied level.

Nb

.

3d Ti3+

VB CB

Nb

- 3d Ti4+

VB CB

ΔEp ΔEd

Fig. 15. Electronic band structure modifications resulting from the interactions between Nb (Ns or Ni) and Ti3+ (oxygen vacancy) defects (figure adapted from Ref.[59]).

(10)

3.5. From bulk to surface: anatase (1 0 1)

Doping of the surface layers may induce new and/or special features in the geometry and electronic structure of TiO2 not present in the bulk. Therefore, it is relevant to determine whether the nitrogen species is on the surface, either in bridging or in a three-coordinated position, or prefers to lie in the second layer of TiO2 (sub-surface).

We have studied this issue in the case of the anatase (1 0 1) surface,Fig. 17.

InFig. 18, we report the total and projected (on the N- atoms) DOS for the most stable substitutional and intersti- tial configurations. From the comparison of the DOS, we can see the occurrence of the N-impurity states near the edge of the valence band for Ns and higher in energy for the Nispecies. The projection onto the N 2p states provides evidence for the localized character of these states. There- fore, N-doping induces a reduction of some tenths of eV of the energy gap between highest occupied and lowest unoccupied states which is not due to a real shift of the valence band edge, but to the appearance of localized states

in the gap, in complete analogy with the bulk case. The presence of an empty spin-down state in the middle of the gap is also evident for both configurations. These states can act as a deep electron trap in the material.

The stability diagram of some of the investigated surface species as a function of the oxygen chemical potential, Fig. 19, is qualitatively similar to that obtained for the bulk species in Fig. 13. However, after careful inspection, one can note that the stability range for the N-atom bound to a lattice O in a surface position is larger than that of the typical interstitial species in the bulk (Fig. 13).

For both surface and sub-surface doping, it is found that at low oxygen concentrations substitutional N-impurities accompanied by oxygen vacancies are favored. The substi- tutional N-impurities are found to be more stable in the sub-surface layers while only surface bridging oxygen vacancies have been considered. The stabilization effect associated to the simultaneous presence of oxygen vacan- cies and nitrogen impurities is a consequence of the elec- tron transfer from the reduced Ti3+ atoms to states deriving from the N substitutional impurities, as observed in the bulk. This stabilization almost cancels the energy cost to form surface bridging O vacancies (from 4.6 to 0.3 eV). Thus, N-doping is expected to introduce many more oxygen vacancies on the surface than commonly observed for the anatase (1 0 1) surface and therefore to lar- gely increase adsorption and reaction properties, com- monly associated to the presence of such defects. Some

as produced

red 1 ox 1 red 2 ox 2 red 3 ox 3 red 4 ox 4

cycle 4 cycle 3

cycle 2 cycle 1

I (a.u.)

Thermal treatment

Nb

.

Fig. 16. EPR intensity of species Nb upon different thermal treatments (intensity in arbitrary units).

Fig. 17. Model of anatase (1 0 1) TiO2surface.

-60 -40 -20 0 20 40 60

-1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5

DOS

E-EV (eV)

Total DOS PDOS, N 2p Total DOS, pure

-60 -40 -20 0 20 40 60

-1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5

DOS

E-EV (eV)

Total DOS PDOS, N 2p Total DOS, pure

Fig. 18. Total (continuous line) and projected (filled line) DOS of (a) Ns- doped and (b) Ni-doped (bound to a surface bridging O atom) anatase (1 0 1) TiO2 surface. The DOS of the undoped system is reported as a reference (dotted line). The origin is set at the top of the valence band of the undoped system. The vertical dotted line defines the uppermost occupied state energy. DFT results with PBE functional.

(11)

experimental evidences of this effect have already been reported in the literature[58].

3.6. Dynamics of N-TiO2under irradiation

Conclusive evidence of the crucial role of Nb in the photoactivity of N-TiO2 in visible light was obtained by irradiating the material with visible light having wave- length corresponding to the absorption maximum [59]

(437 nm, seeFig. 2).

Fig. 20 reports the EPR signals of Nb observed under various conditions.

The first spectrum,Fig. 20a was recorded at room tem- perature using a N-TiO2sample kept in the dark. The sec- ond one, Fig. 20b, was recorded during irradiation of the same sample in visible light (437 nm, blue). The concentra- tion of Nb increases in these conditions but the intensity

immediately recovers the initial value as soon as the light is switched off. The irradiation cycles in vacuo can be repeated indefinitely. The phenomenon is not observed if a different wavelength is used (es. 500 nm, green). The increase of Nb intensity is even higher in the presence of oxygen in the gas phase,Fig. 20c. In such condition, how- ever, a second paramagnetic species is formed which can be best observed at 77 K and which is visible in the EPR spec- trum at room temperature, though overlapped with the sig- nal of Nb, provided that excess oxygen is removed from the atmosphere,Fig. 20d. This species is the well known super- oxide O2 anion, whose formation is due to a photo- induced electron transfer from the solid to the surface adsorbed oxygen molecule.

The formation of superoxide is irreversible at RT. Upon turning the lamp off the intensity value of species Nb remains higher than the initial value recorded in the dark prior to oxygen adsorption.

The experimental results discussed above allowed us to draw a schematic picture of the mechanism connected to the interaction of N-TiO2 with visible light [59]. Indepen- dently of its specific lattice position (interstitial or substitu- tional species), the Nbcenter introduces localized electronic states in the gap which can be either doubly occupied (dia- magnetic Nb) or singly occupied (paramagnetic Nb).

Irradiation under vacuo with visible light 437 nm (2.84 eV) selectively promotes electrons from these states to the conduction band according to the following process:

Nb !hm Nbþ e Nb!hmNþb þ e

ð5Þ

The energy of the visible light in fact is not sufficient to ex- cite electrons from the valence band, as the TiO2anatase band-gap is 3.2 eV. Diamagnetic Nb species at equilibrium are expected to be definitely more abundant than Nb be- cause, under irradiation, an increase of the concentration of the paramagnetic centres Nb is observed, Fig. 20. The equilibrium conditions in the dark are recovered instanta- neously when irradiation is stopped.

The presence of oxygen in the gas phase modifies the sit- uation as a fraction of photoexcited electrons is scavenged by O2producing adsorbed O2:

Nb þ O2ðgasÞ!hm Nbþ O2ðsurfÞ Nbþ O2ðgasÞ!hm Nþb þ O2ðsurfÞ

ð6Þ

This interaction shifts the equilibrium with formation of a higher amount of paramagnetic Nb centres with respect to simple irradiation in vacuo. In other words, photoinduced charge separation has occurred. The generation of surface adsorbed O2 causes the formation of a negative layer at the interface. This, in turn, is expected to cause band bending which limits further electron transfer. Stopping irradiation, the electrons scavenged by oxygen remain in the ad-layer so that the initial concentration of Nb centers

3350 3375 3400 3425

d

c

b

a

B/Gauss

Fig. 20. EPR spectra of species Nb (a) as synthesized in dark, (b) under irradiation in vacuo with visible light of 437 nm (c) under irradiation with visible light of 437 nm in the presence of oxygen (d) after removal of the excess of oxygen.

Fig. 19. Stability diagram for the various N-doped models considered for the anatase (1 0 1) TiO2surface.

(12)

is not recovered. Fig. 21 schematically describes the two processes.

4. Summary and conclusions

In this paper we have reviewed recent work of our group, based on a combination of electron paramagnetic resonance (EPR), X ray photoelectron spectroscopy (XPS), and DFT calculations to characterize the paramag- netic species in N-doped anatase TiO2powders obtained by sol–gel synthesis. EPR measurements have provided evi- dence for the existence of a paramagnetic nitrogen species intimately interacting with the TiO2lattice. The spin-Ham- iltonian parameters of the observed species are found to be consistent with calculations for both substitutional (Ns) and interstitial (Ni) N-impurities. Calculations also show that the electronic states of these species are localized, and their energies are just and somewhat above the top of the valence band for Nsand Ni,respectively[52]. Anal- ysis of the interaction between nitrogen impurities and oxy- gen vacancies and of the relative stability of the different species as a function of the oxygen chemical potential (oxy- gen partial pressure) shows that under oxygen-rich (poor) conditions Ni (Ns accompanied by oxygen vacancies) is most stable. Finally, by investigating the behavior of the sol–gel N-TiO2samples under irradiation with photons of different energy and in the presence of adsorbates, we show that the EPR-active N-species is responsible for the absorp- tion of visible radiation and for the photo-induced electron transfer from the solid to a surface adsorbed electron scav- enger as molecular oxygen. In conclusion, with the present study we have been able to identify the role played by impurities (dopants, oxygen vacancies, hydroxyl groups) in the photo-activity of N-doped TiO2 samples. On the basis of this work we believe that the photocatalytic prop- erties of TiO2, both in UV and Vis-regions, can be largely improved by tailoring the number and type of defect pres- ent in the photocatalyst.

Acknowledgements

This work is supported by the Italian MIUR through a Cofin2005 Project. A.S. acknowledges the Department of Energy for financial support (DE-FG02-05ER15702).

References

[1] K. Hashimoto, H. Irie, A. Fujishima, Japanese J. Appl. Phys. 44 (2005) 8269.

[2] M. Anpo, Bull. Chem. Soc. Jpn. 77 (2004) 1427.

[3] A. Kudo, H. Kato, I. Tsuji, Chem. Lett. 33 (2004) 1534.

[4] T.L. Thompson, J.T. Yates Jr., Chem. Rev. 106 (2006) 4428.

[5] H.M. Yates, M.G. Nolan, D.W. Sheel, M.E. Pemble, J. Photochem.

Photobiol. A 179 (2006) 213.

[6] M. Ni, M.K.H. Leung, D.Y.C. Leung, K. Sumathy, Renew. Sustain.

Energy Rev. 11 (2007) 401.

[7] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001) 269.

[8] S. Sato, Chem. Phys. Lett. 123 (1986) 126.

[9] S. Sakthivel, M. Janczarek, H. Kisch, J. Phys. Chem. B 108 (2004) 19384.

[10] H. Irie, Y. Watanabe, K. Hashimoto, J. Phys. Chem. B 107 (2003) 5483.

[11] O. Diwald, T.L. Thompson, T. Zubkov, E.G. Goralski, S.D. Walck, J.T. Yates Jr., J. Phys. Chem. B 108 (2004) 6004.

[12] M. Miyauchi, A. Ikezawa, H. Tobimatsu, H. Irie, K. Hashimoto, Phys. Chem. Chem. Phys. 6 (2004) 865.

[13] J.L. Gole, J.D. Stout, C. Burda, Y. Lou, X. Chen, J. Phys. Chem. B 108 (2004) 1230.

[14] Z. Lin, A. Orlov, R.M. Lambert, M.C. Payne, J. Phys. Chem. B 109 (2005) 20948.

[15] S. Sato, R. Nakamura, S. Abe, Appl. Catal. B 284 (2005) 131.

[16] M. Sathish, B. Viswanathan, R.P. Viswanath, C.S. Gopinath, Chem.

Mater. 17 (2005) 6349.

[17] M. Alvaro, E. Carbonell, V. Forne´s, H. Garcia, ChemPhysChem 7 (2006) 200.

[18] Y. Nosaka, M. Matsushita, J. Nasino, A.Y. Nosaka, Sci. Technol.

Adv. Mat. 6 (2005) 143.

[19] Y. Irokawa, T. Morikawa, K. Aoki, S. Kosaka, T. Ohwaki, Y. Taga, Phys. Chem. Chem. Phys. 8 (2006) 1116.

[20] T. Morikawa, R. Asahi, T. Ohwaki, K. Aoki, Y. Taga, Jpn. J. Appl.

Phys. 40 (2001) L561.

[21] O. Diwald, T.L. Thompson, E.G. Goralski, S.D. Walck, J.T. Yates Jr., J. Phys. Chem. B 108 (2004) 52.

[22] M. Batzill, E.H. Morales, U. Diebold, Phys. Rev. Lett. 96 (2006) 026103.

[23] A. Ghicov, J.M. Macak, H. Tsuchiya, J. Kunze, V. Haeublein, L.

Frey, P. Schmuki, Nano Lett. 6 (2006) 1080.

[24] J.M. Mwabora, T. Lindgren, E. Avendano, T.F. Jaramillo, J. Lu, S.E.

Lindquist, C.G. Granqvist, J. Phys. Chem. B 108 (2004) 20193.

[25] Y. Nakano, T. Morikawa, T. Ohwaki, Y. Taga, Appl. Phys. Lett. 86 (2005) 132104.

[26] X. Chen, C. Burda, J. Phys. Chem. B 108 (2004) 15446.

[27] Y. Sakatani, J. Nunoshige, H. Ando, K. Okusako, H. Koike, T.

Takata, J.N. Kondo, M. Hara, K. Domen, Chem. Lett. 32 (2003) 1156.

[28] C. Di Valentin, G. Pacchioni, A. Selloni, S. Livraghi, E. Giamello, J.

Phys. Chem. B 109 (2005) 11414.

[29] N. Serpone, J. Phys. Chem. B 110 (2006) 24287.

[30] A. Nambu, J. Graciani, J.A. Rodriguez, Q. Wu, E. Fujita, Sanz J.

Fdez, J. Chem. Phys. 125 (2006) 094706.

[31] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865.

[32] D. Vanderbilt, Phys. Rev. B 41 (1990) 7892.

[33] M. Lazzeri, A. Vittadini, A. Selloni, Phys. Rev. B 63 (2001) 155409.

[34] S. Baroni, et al.<http://www.democritos.it>.

C.B.

V.B.

Ef

C.B.

V.B.

Ef Lamp ON

e-

Lamp Off

C.B.

V.B.

Ef

C.B.

V.B.

Ef Lamp ON

e

Lamp Off O2

.

-

Fig. 21. Energy states involved during N-TiO2 visible irradiation. (a) Vacuum condition. (b) O2atmosphere.

(13)

[35] R. Car, M. Parinello, Phys. Rev. Lett. 55 (1985) 2471.

[36] K. Laasonen et al., Phys. Rev. B 47 (1993) 10142.

[37] V.R. Saunders, R. Dovesi, C. Roetti, R. Orlando, C.M. Zicovich- Wison, N.M. Harrison, K. Doll, B. Civalleri, I.J. Bush, P.h. D’Arco, M. Llunell, CRYSTAL03.

[38] A.D. Becke, J. Chem. Phys. 98 (1993) 5648.

[39] C. Lee, W. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785.

[40] G. Pacchioni, F. Frigoli, D. Ricci, J.A. Weil, Phys. Rev. B 63 (2001) 054102.

[41] C. Di Valentin, G. Pacchioni, A. Selloni, Phys. Rev. Lett. 97 (2006) 166803.

[42] C.M. Zicovich-Wilson, R. Dovesi, J. Phys. Chem. B 102 (1998) 1411.

[43] E. Ruiz, M. Llunlell, P. Alemany, J. Solid State Chem. 176 (2003) 400.

[44] R. Pandey, J.E. Jaffe, N.M. Harrison, J. Phys. Chem. Solids 55 (1994) 1357.

[45] J.A. Weil, J.R. Bolton, J.E. Wertz, Electron Paramagnetic Reso- nance, Wiley, New York, 1994.

[46] S. Livraghi, A. Votta, M.C. Paganini, E. Giamello, Chem. Commun.

4 (2005) 498.

[47] S. Livraghi, M.C. Paganini, M. Chiesa, E. Giamello, Res. Chem.

Intermed. (in press).

[48] J.H. Lundsford, J. Chem. Phys. 72 (1968) 2141.

[49] M.C. Paganini, M. Chiesa, P. Martino, E. Giamello, J. Phys. Chem. B 106 (2002) 12531.

[50] R. Nakamura, T. Tanaka, Y. Nakato, J. Phys. Chem. B 108 (2004) 10617.

[51] M. Mrowetz, W. Balcerski, A.J. Colussi, M.R. Hoffmann, J. Phys.

Chem. B 108 (2004) 17269.

[52] C. Di Valentin, G. Pacchioni, A. Selloni, Phys. Rev. B 70 (2004) 085116.

[53] C. Di Valentin, G. Pacchioni, A. Selloni, Chem. Mater. 17 (2005) 6656.

[54] E. Serwicka, M.W. Schlierkamp, R.N.Z. Schindler, Naturforsch. 36a (1981) 226.

[55] T. Berger, M. Sterrer, O. Diwald, E. Kno¨zinger, D. Panayotov, T.L.

Thompson, J.T. Yates Jr., J. Phys. Chem. B 109 (2005) 6061.

[56] D.C. Hurum, A.G. Agrios, K.A. Gray, T. Rajh, M.C. Thurnauer, J. Phys. Chem. B 107 (2003) 4545.

[57] E. Sutter, P. Sutter, E. Fujita, J. Muckerman, Abstracts of the European Materials Research Society Spring Meeting, Nice, France, 2006, p. M-14.

[58] T. Tachikawa, Y. Takai, S. Tojo, M. Fujitsuka, H. Irie, K.

Hashomoto, T. Majima, J. Phys. Chem. B 110 (2006) 13158.

[59] S. Livraghi, M.C. Paganini, E. Giamello, A. Selloni, C. Di Valentin, G. Pacchioni, J. Am. Chem. Soc. 128 (2006) 15666.

Riferimenti

Documenti correlati

If we were able to prove the convergence of the solution toward a dynamic sharp-interface evolution, then the dynamic energy-dissipation balance for the damped wave equation in

di evitare quote di risultato e trattamenti preferenziali in grado di assicurare una misura effettiva di partecipazione delle donne alla competizione elettorale. Non a caso, è

In conclusion, we have evaluated T-cell responses directed at a HSV-1 in lung transplant patients

Secondly, we conducted a case-control study, comparing M1dG levels detected in subjects affected by colon, gastric or breast cancer (respectively 51, 87 and 75 subjects) and those

Giovanna Siedina (a cura di), Essays on the Spread of Humanistic and Renaissance Literary Civilization in the.. similarities

In this chapter, with the aim to verify whether the presence of graphene changes the adsorption geometry of the oxygen atoms on Ir(111), we studied the oxygen intercalation

Aim of the work: to resolve the ambiguities in the assignment of the red-ox state of the iron atoms in the [FeFe]-hydrogenase active site and its biomimetic models, we compare

room temperature because of the short-range magnetic exchange interaction of deep impurity states in wide band gap semiconductors.. Here, the double exchange is the mechanism