Open access peer-reviewed chapter

Pulsed Laser-Deposited TiO2-based Films: Synthesis, Electronic Structure and Photocatalytic Activity

Written By

Oksana Linnik, Nataliia Chorna, Nataliia Smirnova, Anna Eremenko, Oleksandr Korduban, Nicolaie Stefan, Carmen Ristoscu, Gabriel Socol, Marimona Miroiu and Ion N. Mihailescu

Submitted: 16 September 2015 Reviewed: 22 February 2016 Published: 24 August 2016

DOI: 10.5772/62637

From the Edited Volume

Semiconductor Photocatalysis - Materials, Mechanisms and Applications

Edited by Wenbin Cao

Chapter metrics overview

1,809 Chapter Downloads

View Full Metrics

Abstract

Active under visible light, photocatalysts based on doped titania were obtained via pulsed laser deposition (PLD) method. To find out the crystalline structure, optical properties, and electronic structure, the following techniques such as X-ray diffraction, electronic spectroscopy, electrical conductivity measurements, and X-ray photoelectron spectroscopy (XPS) are used. Photocatalytic activity is monitored by applying the photoreduction of dichromate ions under UV and visible light. The influence of zirconium ions and its content and synthesis conditions on the efficiency of nitrogen incorporation into titania structure that, in turn, determines the electronic structure and photocatalytic ability of the semiconductive materials are discussed. A substitutional nitrogen (Ti–N) rather than an interstitial one (Ti–O–N) is mainly responsible for the observed photoactivity. It is pointed that substitutional nitrogen is responsible for bandgap narrowing or formation of intragap localized states within semiconductor bandgap. The bandgap energy values are sharply decreased, while the relative intensity of substitutional nitrogen XPS peaks is increased. Pulsed laser synthesis of TiO2 films in N2/CH4 atmosphere not only leads to nitrogen incorporation but also to the formation of defects including oxygen vacancies and Ti3+ states which are all contributing to light absorption. An appropriate ratio of gas mixture, optimum zirconia content, suitable pressure, and temperature during synthesis was found for the synthesis of highly active semiconductive films. The highest photocatalytic conversion yields are obtained for nitrogen-doped 10% ZrO2/TiO2 synthesized in N2:CH4 = 5:1 at 100 Pa and at 450°C under both UV and visible light.

Keywords

  • photocatalysis
  • semiconductor
  • bandgap
  • nitrogen-doped titania
  • zirconia
  • pulsed laser deposition
  • substitutional and interstitial nitrogen
  • X-ray photoelectron spectroscopy

1. Introduction

Among the ways leading to the intensive climate change, the widespread contamination of air, soil, and water sources are to be considered. Titania is recognized as the most prospective photocatalyst used in air purification, surface self-cleaning, and wastewater treatment [17]. Since solar radiation includes light of wavelengths from 280 to 4000 nm, the use of titania that absorbs the edge (3–5%) of solar energy is not efficient. Many studies are performed to design and develop photocatalysts based on titania to allow for their use under visible-light irradiation [8, 9]. Nitrogen, the most promising among nonmetal dopants, facilitates visible light absorption by nanomaterials based on TiO2 as a result of its incorporation into titania matrix due to its comparable atomic size with oxygen, small ionization energy, and high stability. Another approach to change the physical, optical, structural, and photocatalytic properties of titania includes employment of d-block metal ions (e.g., zinc, zirconium, iron, chromium, nickel, vanadium, or copper).

An incorporation of zirconium ions in TiO2 lattice was reported to enhance the specific surface area and the surface acidity and to modify the photo-electrochemical properties, leading to the improvement of the photocatalytic activity [10]. Ti1−xZrxO2 solid solution containing anatase led to the increase of bandgap energy with the anodic shift of the upper edge of the valence band position accelerating photocatalytic processes due to the improvement of charge separation. It was found an optimum content of ZrO2 in TiO2 that increased the quantity of the surface active sites such as OH-groups [11]. In the case of zinc-containing nanocomposites, their crystallinity slightly increased with Zn content and ZnTiO3 perowskite phase is formed. The films with low Zn content (1–5 wt%) exhibited superhydrophilicity. Direct photo-electrochemical measurements showed the cathodic shift of the flat band potential position and the increase of the photocurrent quantum yield for TiO2/ZnO in comparison with unmodified TiO2 electrodes that coincided with the increase of their photocatalytic activity [11]. More effective degradation was observed for Zn2+-doped samples with homogeneously distributed noble metal nanoparticles as a result of synergetic effect of new formed Zn2Ti3O8 phase and the charge separation promoted by the noble metal nanoparticles. It was found that a higher dispersion of noble metal crystallization centers did occur near zinc ions in titania matrix [12, 13]. Nanostructured iron titanate and ruthenium-modified titania thin films are reported to catalyze the photocatalytic dinitrogen reduction that is the second, after photosynthesis, most important chemical process in nature. A detailed mechanism investigation clarified the role of metal ions in all complicated steps of this process [14, 15]. However, d-block metals may also act as recombination sites for the photo-induced charge carriers lowering the activity of the semiconductors. Despite the fact that a decrease in bandgap energy was achieved by metal doping, photocatalytic activity has not been remarkably enhanced [16].

Since Sato [17] reported the visible-light activity of N-doped titania in the oxidation of carbon monoxide and ethane, and later, Asahi [18] explored the visible-light activity of N-doped TiO2, there have been many efforts dealing with nitrogen doping of TiO2. Numerous physical techniques such as sputtering [19, 20] and ion implantation [21], atomic layer deposition [22], pulse laser deposition [23], gas phase reaction [24], and the sol–gel method [16] have been successfully applied to obtain nitrogen-modified TiO2.

It is reported [18] that the doping of nonmetals could narrow the bandgap and might drive the response to visible light due to the similar energies of oxygen and nitrogen resulting in the hybridization of N 2p with O 2p states in anatase TiO2. On the other hand, it is stated [25] that isolated impurity energy levels above the valence band are formed as a result of oxygen atoms substitution by nitrogen ones. Irradiation with UV light excites electrons in both the valence band (VB) and the impurity energy levels, but illumination with visible light only excites electrons in the impurity energy level. An alternative point of view, where the visible-light activation of anion-doped TiO2 originates from the defects associated with oxygen vacancies, was also reported. As a result, the color centers appeared displaying these absorption bands, and not to a narrowing of the bandgap of TiO2 [26]. Whereas surface oxygen vacancies were reduced considerably after reheating at 400°C, the activity of N-doped TiO2 was enhanced fourfold suggesting no oxygen vacancies influence on visible-light activity [27]. In addition, nitrogen-modified titania powder obtained by mixing a commercially manufactured titania powder and the different nitrogen containing organic substances were applied for photomineralization of formic acid under visible light. It is concluded that neither nitridic nor NOx species nor defect states are responsible for the visible-light photocatalytic activity of modified titania prepared from urea, but higher melamine condensation products acting as visible-light sensitizers [28].

The main technique proving the nitrogen incorporation into a semiconductor structure as substitutional nitrogen atom (Ns), where nitrogen substitutes oxygen lattice ions or interstitial one (Ni), and N binds to O lattice ions, is X-ray photoelectron spectroscopy (XPS). There is no definite opinion about the XPS measurements of N1s binding energy (BE). The peak at 395.7 eV is assignable to N3− bonded to three Ti atoms because the measured binding energy is very close to that of titanium nitrides (ca. 396.0 eV) [29, 30]; the values of 396–397 eV are assigned to the N–Ti–N [31] or O–Ti–N bonds [32]. The formation of N–N or N–C, N–O groups or chemisorbed dinitrogen was suggested for higher energies of (400–402 eV) [33]. Note that N1s binding energies of 399.1 and 400.5 eV measured for modified by urea commercially available TiO2 powder were assigned to carbon nitrides (399–400 eV, C=N–C), graphite-like phases (400.6 eV, N-Csp2), and to polycyanogen (399.0, 400.5 eV (–C=N–)x) [34]. XPS and electron paramagnetic resonance (EPR) measurements showed that the interstitial nitrogen atoms corresponded to BE of 400–401 eV are the photoactive species and can be prepared from discharge in molecular nitrogen in the presence of pure anatase [35]. However, the intensity of 400 eV peak belonged to the nitrogen atom from ammonium ions is reduced after rinsing as provided by XPS and solid-state nuclear magnetic resonance (NMR) [36].

The geometric structures of diamagnetic (Scheme 1, i) and paramagnetic (Scheme 1, ii–vi) nitrogen species incorporated into titania suggested from the literature were summarized in Ref. [29]. Nitrogen species with a binding energy of ca. 396 eV represent TiN species (nitridation of titania) described as structure i, where N3− species are directly bonded to three Ti3+ cations [29]. The N contained species corresponding to structure ii are related to a binding energy in the range of 397–399 eV [30, 37]. The formation of O–Ti–N–Ti during the synthesis could take place (structures iii and iv), when nitrogen substituted the lattice oxygen atom of O–Ti–O–Ti [38]. The structure v where the nitrogen atom connected with two oxygen of the TiO2 lattice depicted as imino-type species and may be formed as a result of the incomplete reaction of ammonia with surface oxygen atoms [39]. It is believed [40] that the imino-type species undergoes to nitrate-type species (structure vi) where the last is the photoactive nitrogen species.

Scheme 1.

Proposed structures of N species incorporated into titania as summarized in Ref. [29]. Reproduced with permission.

In light of numerous literature investigations, the applied synthesis conditions and the source of nitrogen are the turning points for the level of nitrogen doping and its states that change optical, electronic, and photocatalytic properties of a semiconductor.

The main principles of heterogeneous photocatalysis are presented as follows (Figure 1): (i) A semiconductor is characterized by an electronic band structure in which the highest occupied energy valence band (VB) and the lowest empty conduction band (CB) are separated by bandgap; (ii) only a photon of energy higher or equal to the bandgap energy (≥3.2 eV for anatase) promotes to a charge separation resulting in an electron (e) transferring from the VB to the CB with simultaneous generation of a hole (h+) in the VB (pathway 1); (iii) an electron and a hole can be trapped in surface states and react with donor (D) and acceptor (A) species, respectively, adsorbed or close to the surface of the particle resulting in the simultaneous oxidation and reduction (pathways 1.1 and 1.2); or (iv) a photoformed electron–hole pair can be recombined (pathway 2) on the surface or in the bulk of the particle in a few nanoseconds. The energy level at the bottom of the CB is actually the reduction potential of photoelectrons and the energy level at the top of the VB determines the oxidizing ability of photoformed holes. The flat-band potential presents the energy of both charge carriers at the semiconductor–electrolyte interface and depends on the nature of the material and the system equilibrium.

Figure 1.

Schematic presentation of possible pathways after light absorption by doped titania particle.

From a thermodynamic point of view, the trapped electron reduces an electron acceptor only if its redox potential is more positive than the flat-band potential of the CB, while the trapped hole oxidizes the adsorbed species having the redox potential more negative than the flat-band potential of the VB. The visible-light photoactivity of metal-ions-doped TiO2 can be explained by a new energy level produced in the bandgap where an electron can be excited from the defect state to the TiO2 conduction band by photon with the lower energy (pathway 3). Inhibition of charge carrier recombination rate resulting in enhanced photoactivity is attributed to the transition metal ions doping of TiO2. Visible-light illumination can generate a hole in the mid-gap level above the top of the VB for the N-doped TiO2 (pathway 4).

The nitrogen-modified TiO2 thin films exhibited photocurrents in the visible up to 700 nm as a result of occupied nitrogen-centered surface states above the valence band edge as noted in Ref. [41]. Moreover, the observed photocurrent under visible-light irradiation of N-doped TiO2 in the presence of reducing agent strongly depend on the reaction mechanism of oxidation rather than on the redox potential of the whole acceptors [42]. Photo-electrochemical measurements showed that the visible-light photogenerated holes in TiO2 modified by urea are able to oxidize formic acid comparing with ammonia modified one. An enhanced stabilization of the photogenerated hole against recombination with conduction band electrons for the former sample is explained [43]. In addition to the electrochemical investigations, preliminary computer simulations based on innovative cluster approaches demonstrated that nitrogen doping narrowed the optical bandgap of TiO2 generating N(2p) states [44]. Density functional theory (DFT) was employed to demonstrate the interstitial nitrogen character within anatase TiO2. No significant shift in the CB or VB of the TiO2 was recognized. The energy bonding states associated below the valence band and anti-bonding states present above the valence band. The anti-bonding nitrogen orbital between the TiO2 VB and CB is suggested to be responsible for the visible light absorption by acting as a stepping stone for excited electrons between conduction and valence bands [45].

Our investigation was directed to the synthesis of nitrogen-doped photocatalysts based on titania and/or mixed semiconductive oxides films and understanding the relationship between synthesis conditions, electronic structures, and photocatalytic activity. The influence of zirconium ions and its content as well as pulsed laser synthesis conditions on the efficiency of nitrogen incorporation into titania structure that, in turn, determines the photocatalytic ability of the semiconductive materials are obtained.

Advertisement

2. Correlation of electronic structure, optical, structural, and photocatalytic properties of TiO2-doped films

Pulsed laser deposition method used for the film synthesis allows obtaining high purity thin films of single or multicomponent materials [46]. PLD or laser ablation is a physical vapor deposition process based on the vaporization of condensed matter by means of photons [4650]. A highly intense short-pulsed high-power laser beam (usually a ns UV one) is focused under vacuum or a working gas atmosphere on a piece of material (on a material specimen) called “target.” If the laser fluence exceeds a specific threshold, one after one, each tiny quantity of ablated material by a laser pulse, is directed forward, in form of a plume, towards the deposition substrate, where the growing film is formed by recondensation [4750]. Due to the high energetic species arriving on the substrate, the synthesized films are very adherent. The plasma plume plays a significant piston-like role, pushing and carrying the removed matter from the target to the substrate [49]. In the most common setup geometry, “on axis,” the collector is placed parallel to the target, at a distance of some centimeters and, after being focused and transmitted by suitable optics, the laser pulses hit the target under 45°. The target is normally rotated in order to get a uniform ablation of its surface, avoiding drilling, thus obtaining a homogenous film. The collector, which is named usually “substrate,” may be heated if a high degree of crystallinity or even epitaxy is of interest for the obtained coating. A schematic of PLD setup is given in Figure 2.

Figure 2.

Schematic setup of a PLD system: The laser source is outside the deposition chamber, which is coupled to a vacuum system. Inside chamber, a classical “on axis” geometry is presented.

The PLD presents a series of advantages [4750], among which we mention: almost any material or materials’ combination may be congruently transferred as thin film on a suitable substrate after the identification of a selected deposition regime, a fine control of thickness is obtained by the number of the applied laser pulses (and the appropriate setting of various parameters), it is a clean, non-contaminating method, also allowing for an easy doping or multilayer structures. These and others made PLD a tool of remarkable performance, versatility, and simplicity for the growth of various types of thin films: complex oxides, high-temperature superconducting, protective or ultrahard coatings, polymer or organic thin films [5052].

ZrO2 % in TiO2 Ratio of gases Sample code P, Pa t, °C Ebg, eV
0 O2 TiO/20 20 450 3.3
0 N2:O2 = 1:1 TiNO/10 10 450 2.8
0 N2 TiN/20 20 450 2.6
0 СН4 TiC/5 5 450
0 N2:СН4 = 5:1 TiN5C1/100 100 450 3.3
0 N2:СН4 = 5:1 TiN5C1/3 3 450
2.5 O2 2.5ZrTiO/100 100 450 3.3
5 O2 5ZrTiO/100 100 450 3.4
10 O2 10ZrTiO/100 100 450 3.5
2.5 N2:СН4 = 9:1 2.5ZrTiN9C1/100 100 450 3.0
5 N2:СН4 = 9:1 5ZrTiN9C1/100 100 450 2.9
2.5 N2:СН4 = 5:1 2.5ZrTiN5C1/100 100 450 3.4
5 N2:СН4 = 5:1 5ZrTiN5C1/100 100 450 3.3
10 N2:СН4 = 5:1 10ZrTiN5C1/100 100 450 3.1
50 N2:O2 = 1:1 50ZrTiN1O1/5 5 600 3.5
50 O2 50ZrTiO/10 10 600 3.7
50 N2 50ZrTiN/10 10 600 3.0
50 CH4 50ZrTiC/10 10 600 3.2
50 N2:CH4 = 10:1 50ZrTiN10C1/3 3 600
50 CH4 50ZrTiC/5 5 600
50 N2:CH4 = 5:1 50ZrTiN5C1/10 10 600 3.6

Table 1.

PLD conditions and bandgap values of TiO2-based films.

PLD experiments were performed using a KrF* laser source (COMPexPro 205 Coherent, τFWHM ≤ 25 ns). Pristine TiO2 or ZrO2-mixed TiO2 targets with different content of zirconium oxide were laser ablated to grow coatings on glass plates. The films were synthesized at substrate temperatures of 450 or 600°C. The samples were deposited at different pressure of pure oxygen, nitrogen, methane, and mixed N2:CH4 atmosphere with 1:1, 5:1, 9:1, and 10:1 ratios.

Photocatalytic activity of the films was assessed via dichromate ions reduction reaction. The film was immersed in 15 ml of an aqueous solution of potassium dichromate (in all experiments, the initial concentration of dichromate ions was 2*10−4 M) and the reducing agent (disodium salt of ethylenediaminetetraacetic acid) in the molar ratio 1:1 adjusted to pH ≥ 2 by perchloric acid. The mass of the films was found to be 0.40 ± 0.02 mg (the concentration of the catalyst in aqueous solution of redox couple was 2.7 × 10−2 g/L, 37 times lower than usually used suspension 1 g/L). The reaction temperature was kept constant (25°C) during the experimental procedure. The change of Cr(VI) ions concentration was monitored with a Lambda 35 UV–Vis spectrophotometer (PerkinElmer) every 20 min. The film was immersed in the solution until complete adsorption in the dark occurred and then irradiated by 1000 W middle-pressure mercury lamp for 120 min. The distance lamp reactor was set at 90 cm. Two blank experiments were carried out: the catalytic reduction of dichromate ions (dark condition) and photoreduction reaction (a bare glass was used instead of film). The conversion percentage of every photocatalytic run is given after subtraction of the corresponding value of the last blank. For testing the visible-light sensitivity, a filter transmitting light with λ > 380 nm was introduced in the photocatalytic setup.

The sample codes that will appear over the chapter can be easily understand as x(Zr)TiYy(Zz)/P, where x is the ZrO2 percentage, (Zr)Ti is a target composition, Y and Z mean gas atmosphere (O is O2, N is N2, C is CH4), y and z are the ratio of Y and Z, P is the gas pressure during PLD procedure. The synthesis conditions and the measured value of bandgap energy (Ebg) are collected in Table 1.

2.1. Nitrogen-doped TiO2 films

Pulsed laser synthesis of TiO2 films in oxygen, nitrogen, methane and their mixtures was performed (Table 1) [53]. Excluding the UV part of spectrum by cut-of-filter, the films obtained at different atmospheres exhibited the visible light absorption (Figure 3). Compared to bare TiO2, nonmetal-doped films show a change in optical absorption that was originally associated to a bandgap narrowing. Comparing the values of bandgap energy calculated by extrapolating the linear parts of the (αhν)1/2 ~ f(hν) curves, that is assuming an indirect electronic transition (Table 1), the decrease in its values for the doped samples is obvious proving that the electronic structure of a photocatalyst affects its optical absorption. The sharp decrease of Ebg was demonstrated when pure dinitrogen or the mixture of dinitrogen and oxygen has been applied.

Figure 3.

Absorption spectra of TiO (1), TiN (2), TiN5C1 (3), TiNO (4), and cut-of-filter (dashed line).

As revealed by X-ray diffraction (XRD) spectra (Figure 4), anatase phase formation has been obtained for all films excepting TiC/5. Moreover, the samples TiNO/10 and TiO/20 present a preferential texture along the <004> plane. On the other hand, the peaks at 37.80°, 43.30°, and 62.80° visible in the diffractogram of the TiO2 film deposited in CH4 (TiC/5 film) could be related to <004>, <002>, and <202> planes of TiO phase indicating the formation of a sub-stoichiometric compound [53] or even Ti crystalline phase [29, 54, 55]. It is obvious that pure methane used in PLD prevents anatase formation providing oxygen atoms deficiency.

Figure 4.

XRD spectra of TiO2 films deposited at 450°C in oxygen, nitrogen, methane, and mixture of nitrogen and oxygen (*—anatase phase and #—TiO or Ti phase). Reproduced with permission from [53].

Intensity belonging to substitutional Ns (N–Ti) lines at 395.2 and 395.8 eV for TiYy(Zz)/P is varied as clearly seen from Figure 5 and Table 1. In addition, the appearance of the line at 397.8 and 398.1 eV accompanied by characteristic lines in the O1s and Ti2p spectra (not shown here) reveals a more complex structure of nitrogen in the doped TiO2 films. The peak at BE ~ 457.5 eV in the Ti2p XPS spectrum of this film can be attributed to either the Ti atoms in the N–Ti–O linkages or Ti species interacting with VO, whose formation is highly favored in the oxygen-poor conditions (Table 2) [57].

Figure 5.

XPS spectra of N1s binding energy for samples TiNO/10 (a) and TiN/20 (b) adopted from [9] and TiN5C1/100 (c) and TiN5C1/3 (d). Adapted with permission from [56].

Replacing dinitrogen or the mixture of dinitrogen/oxygen by the mixture of dinitrogen/methane leads to the change not only the line intensities but also some peaks position. Quite surprisingly, the intensity of N1s peak of Ns is sharply raised and reached ca. 90% for TiN5C1/3. It points that no other nitrogen contained species are present in the lattice structure of titania for this materials. A new additional peak of Ti2p3/2 region at lower BE (456.3 eV) is observed for TiN5C1/3. Similarly, the Ti2p3/2 lines corresponded to 455.6 and 456.9 eV appeared with increasing nitridation temperature leading to the formation of TiN phase as proven by XRD [29]. Taking into account that the peak at 456.3 eV is assigned by the published literature to TiN phase [30], it could be suggested the deeper nitrogen doping into titania at the low pressure (N–Ti–N and the dative-bonded N atoms). Thus, the deposition atmosphere and pressure have strong influence on the efficiency of Ns and Ni atoms incorporation.

BE, eV TiNO/10 TiN/20 TiN5C1/100 TiN5C1/3
I, %
N1s
395.2 26.4
395.8 25.2 9.7 90.1
397.8 27.4
398.1 7.3
399.1 12.9 33.2 39.6 5.6
400.3 61.9 18.0
400.8 23.3 4.3
409.2 15.1
Ti2p 3/2
456.3 15.3
457.5 2.6 16.8 24.4 20.3
458.1 60.6 48.1
458.3 62.9 67.9
458.8 34.5 15.3 15.0 16.3

Table 2.

N1s and Ti2p3/2 binding energy values and their relative intensity.

Comparing the bandgap energy values of TiN/20 (2.6 eV), TiNO/10 (2.8 eV), and TiN5C1/100 (3.3 eV) with their intensity of Ns peak 26.4, 25.2, and 9.7, respectively, the bandgap narrowing of anatase could result in the incorporation of substituted nitrogen.

Sample UV, % vis, %
TiC/5 15 3
TiO/20 24 3
TiNO/10 3
TiN/20 9
TiN5C1/100 25 9
TiN5C1/3 1

Table 3.

Conversion percentage of dichromate ions after 120 min under UV and visible light.

The catalytic response under UV light exposure was observed for TiC/5, TiO/20, and TiN5C1/100 samples (Table 3). TiNO/10 and TiN/20 films were completely inactive under UV light, while the activity of TiN5C1/100 is comparable with pure TiO2 (TiO/20). The low activity of former N-doped materials could be explained by the nitrogen-induced levels formation in the banggap of TiO2, acting as recombination centers for charge carriers under irradiation with UV light. TiN/20 and TiN5C1/100 exhibited an identical level of photocatalytic conversion under visible light, while the relative intensities of Ns (26.4 and 9.7 for TiN/20 and TiN5C1/100, respectively) and Ni (7.3 and 27.4 for TiN/20 and TiN5C1/100, respectively) are different for these samples. No photocatalytic activity in the presence of TiN5C1/3 is observed.

2.2. Nitrogen-doped ZrO2/TiO2 films

2.2.1. Low content of ZrO2 in the target

Incorporation of transition metal oxides into titania improved the photocatalytic activity due to the mutual action of metal ions and anatase resulting in increased specific surface area, surface acidity, and changed photo-electrochemical properties. Coupling of two semiconductors is useful to achieve a more efficient separation of photogenerated electron–hole pair that led to the improvement of the photoactivity. It is found that the photocatalytic activity depends on the phase composition and crystalline size that modify the TiO2 bandgap [11]. Zirconium incorporation into TiO2 lattice by sol–gel synthesis was achieved with the formation of Ti1−xZrxO2 solid solution for the film with a Zr content up to 10 mol% as evidenced by XRD and XPS. It was demonstrated that the formation of Zr–O–Ti bonds has an influence on the thermal stability during sintering of the mesoporous structure of the films, surface texture, and optical properties as well as in the changes of number of surface active sites for nanocomposite films [10, 58].

Pulsed laser-deposited (2.5, 5, and 10%) ZrO2/TiO2 thin films were synthesized under various gas atmospheres (see Table 1). We inferred the characteristic Ebg values for pure titania and zirconia of 3.3 and 4.2 eV. All nonmetal-doped films showed more intensive absorption extending into the visible-light region comparing xZrTiO/100. A shift to the longer wavelength from lower to higher contents of ZrO2 results in decrease of bandgap energy, as oppositely observed for the samples synthesized in O2. The bandgap values of the films synthesized in a gas atmosphere with a higher ratio of N2 to CH4 (9:1) are lowered when increasing zirconia content.

For films synthesized at 3 Pa, the calculation of Ebg was impossible because of the strong absorption in the visible region (Figure 6 inset) [56]. Similar broad absorption was reported for N-doped TiO2 after the nitridation by ammonia at 650°C or above and explained by the metallic property of samples as indicated by the XRD and transmission electron microscopy (TEM) result [29].

The N1s line at 395.8 eV attributed to Ns is present for all nonmetal-doped samples. The XPS spectra of 2.5ZrTiN9C1/100 and 2.5ZrTiN5C1/100 films showed almost the same relative intensity of Ns lines as TiN5C1/100 (Figures 5c and 7b, d) and Tables 2 and 4). On the other hand, the increased intensity of Ns lines is fixed for the samples containing 5% ZrO2, especially for the film synthesized at higher ratio of N2 to CH4 (5ZrTiN9C1/100). It shows that the higher the dinitrogen content in the gas mixture, the more substitutional nitrogen is incorporated in the semiconductive structure. Comparing the intensity of N1s line at 395.8 eV of all samples synthesized at 100 Pa, it is evident that highest peak intensity (~43%) corresponds to the sample with the largest content (10%) of zirconia, while the lowest intensity at 397.8 eV is observed.

Figure 6.

Absorption spectra of 2.5ZrTiN5C1/100 (1), 5ZrTiN5C1/100 (2), and 10ZrTiN5C1/100 (3). Inset: samples 5ZrTiN5C1/3—a, 5ZrTiN9C1/3—b, 5ZrTiN5C1/100—c. Reproduced with permission from [56].

Figure 7.

XPS spectra of N1s binding energy for 2.5ZrTiN9C1/100 (a), 2.5ZrTiN5C1/100 (b), 5ZrTiN9C1/100 (c), and 5ZrTiN5C1/100 (d) samples. Reproduced with permission from [56].

BE, eV 2.5ZrTiN9C1/100 2.5ZrTiN5C1/100 5ZrTiN9C1/100 5ZrTiN5C1/100
I, %
N1s
395.8
397.8 3.8 14.4 11.6 18.1
399.2 60.3 55.5 44.5 47.0
400.8 27.9 21.4 22.5 21.5
Ti2p 3/2
457.5 3.2 7.2 11.2 5.1
458.3 68.3 57.4 72.9 63.5
458.8 28.5 35.4 15.9 31.4

Table 4.

N1s and Ti2p3/2 binding energy values and their relative intensity.

Analogous to TiN5C1/3, Ns peak for 10ZrTiN5C1/3 films contains the single high-intensity N1s line at 395.7 eV. No other deconvoluted peaks were found (Figure 8; Table 5). Thus, an incorporation of substitutional nitrogen ions without any other forms can be achieved by carrying the synthesis at 3 Pa pressure.

Figure 8.

XPS spectra of N1s binding energy for 10ZrTiN5C1/100 (a) and 10ZrTiN5C1/3 samples (b).

Such effective nitrogen incorporation in oxide matrix can be the effect of an interaction between carbon (and/or hydrogen) and laser ablated oxygen atoms forming CO2 during PLD that leads to the oxygen deficiency and the formation of nonstoichiometric TiO2−xNx structure. This can be confirmed by the lower ratio (1.6) of OTi1s/Ti2p for samples synthesized at 3 Pa in comparison with pristine TiO2 where this ratio equals 2 [56].

BE, eV 10ZrTiN5C1/100 10ZrTiN5C1/3
(a) (b)
I, %
N1s
395.7 43.2 86.8
397.8 2.8
399.1 26.2 4.6
400.8 27.9 8.7
Ti2p 3/2
456.3 8.1
457.5 21.0 37.2
458.1 72.80 47.1
458.8 6.24 7.6

Table 5.

Relative intensity of N1s and Ti2p3/2 binding energy.

A significant intensity of peak at 457.5 eV assigned to the Ti–N bonds was observed for 10ZrTiN5C1/100 and 10ZrTiN5C1/3, assuming also the presence of Ti3+ states. According to the literature [59], such titanium species have oxidation states between +3 and +4. Moreover, the appearance of additional Ti2p3/2 peaks with lower BE similar to TiN5C1/3 (Section 2.1) is assigned to the change of surrounding titanium, where more oxygen atoms are substituted by nitrogen. The Zr3d5/2 line (Ebg = 182.3 eV) was recorded for all zirconia-containing samples, meaning the formation of Zr4+ ions surrounded by O2− ions [10]. No signal pointing on the formation of Ti–O–Zr bonds was obtained (not shown here).

Additionally, the ratio of N1s line at 395.8 eV to Ti2p3/2 line at 457.5 eV corresponds as 1:2 and 2:1 for TiN5C1/100 and 10ZrTiN5C1/100, respectively. It is supposed that the distortion of Ті4+О2 lattice with the advent of Ti3+ states occurs due to larger size of zirconium ions (RZr4+ = 0.720 Å vs. RTi4+ = 0.650 Å). The relatively high stability of Ti3+ states of the tested samples is associated to the presence of Zr4+ ions. Note that the non-equilibrium character of the PLD process is favorable to the generation of strains in the lattice. XPS data pointed to the correlation between zirconium contents and ratio of number of nitrogen atoms linked in Ti–N bonds to the number of Ti3+ states as well as the relative contents of N1s at 395.8 eV. This can be indicative for the formation of Ті–N fragments in the coordination sphere of Ti3+ ions. In other words, the Ns states (395.8 eV) responsible for photocatalytic activity of doped samples could be also connected with Ті3+ states of Ті4+О2 lattice.

Certainly, the non-doped ZrO2/TiO2 films (columns 7, 8, and 9 in Figure 9) exhibit much lower activity under both UV and visible light compared to nitrogen-doped synthesized at 100 Pa. The highest yield under both UV (51% reduced ions) and visible light (14%) was demonstrated by 10ZrTiN5C1/100. The activity of this film is twice higher than TiO/20 and TiN5C1/100 under UV light and 1.5 times higher comparing with TiN/20 and TiN5C1/100 under visible light. We have shown that the conversion percentage under UV light is strongly correlated with the contribution of N1s at 395.8 eV which in turn depends on the zirconia content (Tables 4 and 5). Hence, the photocatalytic activity strongly depends on the efficiency of substitutional nitrogen incorporation inside titania matrix that is, in turn, affected by zirconia content and the gas ratio used during PLD synthesis.

Figure 9.

Conversion of dichromate ions after 120 min under visible (a) and UV (b) light: 1—TiN5C1/100; 2—2.5ZrTiN5C1/100; 3—5ZrTiN5C1/100; 4—10ZrTiN5C1/100; 5—2.5ZrTiN9C1/100; 6—5ZrTiN9C1/100; 7—2.5ZrTi/100; 8—5ZrTi/100; 9—10ZrTi/100.

It must be emphasized that all films obtained at lowest pressure are completely inactive under both UV and visible-light irradiation (not shown here). In order to understand in depth this phenomenon, the electrical conductivity (Figure 10) measurements were performed. The samples obtained in the range of (3–10) Pa are black colored (Figure 11) and exhibit the electrical conductivity pointing on the appearance of oxygen vacancies, which results in metallic character of these materials. The formation of nonstoichiometric TiO2−xNx containing in excess of Ti–N bonds exhibiting a low semiconductive behavior is concluded.

Figure 10.

Electrical conductivity of 10ZrTiN5C1/X: 3 (1), 6 (2), 10 (3), 50 (4), and 100 Pa (5).

Moreover, no electrical conductivity (Figure 10, curves 4 and 5) was observed for the samples obtained at 50 and 100 Pa providing their semiconductive behavior. This indicates that anatase in conjunction with a certain numbers of Ti–N bonds is beneficial on the effective photocatalytic transformations.

Figure 11.

Images of 10ZrTiN5C1/X obtained under different gas pressures.

As seen from Figure 11, the level of nitrogen incorporation strongly affects the optical properties of the films which turned from black to dark yellow color.

2.2.2. High weight content of ZrO2 in the sample

The next series of the films used in our investigation was based on the mixed titania–zirconia films with the component ratio 50:50 wt%. The synthesis conditions are detailed in Table 1. TiO2/ZrO2 films prepared in ambient N2 and/or CH4 atmosphere had different absorption spectra. The shift of the absorption band edge to the red region of the spectra with the increase of N2 content in the gas mixtures was observed (Figure 12). The influence of the doping agent nature and its content was observed in the variation of bandgap energy values (Table 1).

Figure 12.

Absorption spectra of 50ZrTiO/10 (1), 50ZrTiN1O1/5 (2), 50ZrTiN5C1/10 (3), and 50ZrTiN/10 (4) films.

Similar to the bandgap values of low zirconia content films (Section 2.2.1), Ebg values correlate with dinitrogen content in the gas mixture and applied pressure. For non-doped ZrO2/TiO2 (50ZrTiO/10), the bandgap energy value is higher (3.7 eV) than for anatase, pointing on the presence of wide bandgap semiconductor as zirconia. In particular, the Ebg values are decreased with the increase of dinitrogen content in a gas mixture, reaching 3.0 eV for 50ZrTiN/10. On the other hand, for the doped composite samples 50ZrTiN10C1/3 and 50ZrTiC/5, the calculation of Ebg was impossible because of the strong absorption in the visible region.

Unexpectedly, no crystallization of titania but incipient crystallization corresponding to the (101) line of ZrO2 were confirmed by XRD investigation (not shown here) [60]. It can be explained by the existence of undetectable very small TiO2 particles or by the excess of ZrO2 in the film matrix inhibiting the titania crystallization.

Figure 13.

XPS spectra of N1s energy for the samples 50ZrTiN/10 (a), 50ZrTiN5C1/10 (b), 50ZrTiN10C1/3 (c).

The sample synthesized at the high content of N2 in the gas mixture (50ZrTiN10C1/3) contains the most prominent peaks of Ti2p3/2 at 457.5 eV and N1s at 395.8 eV (Figure 13; Table 6), indicating the effective Ns incorporation at low pressure similar to low zirconia content films (Section 2.2.1). Existence of other nitrogen species (400 eV) could be explained by high content of zirconia that caused the distortion of the crystalline lattice and easier incorporation of other N species. Almost equal intensity of the peaks at 395.8 and 400.0 eV was observed for sample deposited in pure N2. The Ti2p3/2 signal at 457.5 eV was weaker for sample obtained in pure N2 than in 5:1 N2:CH4 at 10 Pa. The binding energy of Zr3d5/2 peak for all samples (182.3 eV) corresponded to Zr4+ species in ZrO2.

BE, eV 50ZrTiN/10 (a) 50ZrTiC/5 (b) 50ZrTiN5C1/10 (c) 50ZrTiN10C1/3 (d)
I,%
N1s
395.8 42.5 27.6 67.6
400.0 57.5 100 68.5 32.4
Ti2p 3/2
457.5 7.5 8.1 23.5
458.3 36.9 9.0 31.7 27.8
55.6 91.0 60.2

Table 6.

Relative intensity of N1s and Ti2p3/2 binding energy from Figure 13.

All the films were inert under UV and visible-light irradiation showing an activity similar to that of blank sample. The photocatalytic performance of 50ZrTiN10C1/3 film was comparable to TiO2 one under UV light. The highest activity for this series of films was reached for the structure with the largest number of Ns [60]. As clearly seen, no favorable effect of half doping level of ZrO2 in titania was recognized as a result of the low crystallization degree of anatase, while the synthesis conditions for the effective nitrogen substitution were chosen.

Advertisement

3. Conclusions

In summary, an appropriate selection of gas mixture, pressure, and temperature during PLD synthesis is identified for the optimum photocatalytic activity of semiconductive materials. The highest photocatalytic conversion yields in the dichromate ions reduction are obtained for nitrogen-doped 10% ZrO2/TiO2 synthesized in N2:CH4 = 5:1 at 100 Pa and at 450°C under both UV and visible light. This is the result of effective nitrogen atom substitution into titania lattice as confirmed by N1s XPS line at 395.8 eV. When the films were synthesized at low pressure of 3 Pa, the only Ns peak with the binding energy 395.8 eV was observed and photocatalytically inactive nonstoichiometric TiO2−xNx structures were deposited. The absence of semiconductive properties of the films obtained at the low pressures is a direct evidence of no anatase formation and, respectively, their photocatalytic performance remained at the level of blank experiments. It can be suggested that carbon and hydrogen radicals/atoms forming as a result of laser action react with oxygen species causing the subsequent oxygen substitution by nitrogen atoms. This mechanism is supported by the XPS evidence, according to which the intensity of N–Ti peak at 395.9 eV is lower for the samples synthesized in a pure N2 atmosphere. Thus, the effective incorporation of substitutional nitrogen in oxide matrix can be caused by a direct interaction of carbon or hydrogen atom with oxygen species during PLD resulting in an oxygen deficient lattice, taking into account that pure methane used in PLD prevents anatase formation providing oxygen atoms deficiency. At high pressure, a number of collisions between C or H species and ablated oxygen atoms are less probable owing to the screening effect of other atoms/species (Figure 14a). When pressure is decreased, the deposition rate is enhanced in addition to much higher probability for carbon/oxygen and hydrogen/oxygen interactions (Figure 14b). This can result in the higher efficiency of Ti–N as compared with Ti–O fragments formation.

Figure 14.

Schematic representation of pulse laser deposition in a mixture of N2/CH4 at 100 Pa (a) and 3 Pa (b).

As seen from Figure 15, the percentage of reduced dichromate ions under UV light is correlated with the contributions of N1s at 395.8 eV which is in turn dependent on the zirconia content. The ratio of line intensities of N1s at 395.8 eV to Ti2p3/2 at 457.5 eV is pointing to the large number of O–Ti3+–N bonds with the increase of ZrO2 contents. It is obvious that photocatalytic activity under UV light strongly depends on the efficiency of substitutional nitrogen incorporation inside titania matrix. The high ZrO2 content (50%) induces the deceleration of anatase crystallization rate leading to the activity decrease, while a much lower zirconia amount (2.5 and 5%) is still not enough to stimulate an appropriate quantity of substituted nitrogen embedded in titania matrix. The correlation between the zirconia content and the efficiency of substitutional N incorporation is established. The distortion of Ті4+О2 lattice with the advent of Ti3+ states occurred due to the larger radius of zirconium ions (RZr4+ = 0.720 Å vs. RTi4+ = 0.650 Å). The relative high stability of Ti3+ states is assigned to the presence of Zr4+ ions.

Figure 15.

Influence of zirconia contents in (xZr)TiN5C1/100 films on the photocatalytic conversion under UV (black column), ratio of N1s at 395.8 eV to Ti2p at 457.5 eV (red column) and N1s at 395.8 eV to total N1s (blue column). Reproduced with permission from [56].

Figure 16.

Dependence of bandgap energy of the semiconductive materials with the relative intensity of N1s XPS signal at 395.2–395.8 eV.

In general, the low photocatalytic efficiency is explained by the limited photo-excitation of electrons in the intragap localized states, the very low mobility of photogenerated holes [2], and the high recombination rate due to the creation of oxygen vacancies by doping [61]. Hence, the photoreactivity of doped TiO2 appeared to be a complex function of the doping agent concentration, the energy level of doping agent within the TiO2 lattice, their d electronic configurations, the distribution of doping agent, the electron donor concentrations, and light intensity.

The debates concerning the preferred N substitutional or interstitial sites which induce the highest photocatalytic action is still alive in the literature [6264]. Our investigation suggests that the substitutional nitrogen (Ti–N) that belongs to XPS lines at 395.2–395.8 eV is basically responsible for the observed photoactivity. Additionally, it can be proved by the next two observations. (i) The peak at 397.8 eV has the negligible relative intensity (2.8%) for most active 10ZrTiN5C1/100, and (ii) the peaks in the range of 398–401 eV are observed for the other samples with low reactivity. We also underline that Ns species with 395.2–395.8 eV binding energy are responsible for bandgap narrowing or formation of intragap localized states of the doping agent within bandgap (Fig. 16). It is pointed that the bandgap energy values are sharply declined when the relative intensity of this XPS peaks is increased, while no dependence between 397.8 or 398–401 eV peaks and Ebg is observed. PLD synthesis of TiO2 films in N2/CH4 atmosphere not only leads to nitrogen incorporation but also to the formation of defects including oxygen vacancies and Ti3+ states which are all contributing to light absorption.

Advertisement

Acknowledgments

All authors acknowledge the collaboration between Romanian Academy and National Academy of Sciences of Ukraine. Romanian authors acknowledge with thanks the support of CNCS-UEFISCDI under the contract ID304/2011 and Core Programme 2016–2017 LAPLACE-IV.

References

  1. 1. Fujishima A., Rao TN, Tryk DA., Titanium dioxide photocatalysis. J. Photochem. Photobiol. C: Photochem. Rev. 2000; 1(1): 1–21.
  2. 2. Hashimoto K., Irie H., Fujishima A., TiO2 photocatalysis: a historical overview and future prospects. Jpn. J. Appl. Phys. 2005; 44: 8269–8285.
  3. 3. Thompson TL, Yates JTJr, Surface science studies of the photoactivation of TiO2 new photochemical processes. Chem. Rev. 2006; 106: 4428–4453.
  4. 4. Fujishima A., Honda K., Electrochemical photolysis of water at a semiconductor electrode. Nature. 1972; 238(5358): 37–38.
  5. 5. Frank SN, Bard AJ, Heterogeneous photocatalytic oxidation of cyanide ion in aqueous solutions at titanium dioxide powder. J. Am. Chem. Soc. 1977; 99(1): 303–304.
  6. 6. Wang R., Hashimoto K., Fujishima A., Chikuni M., Kojima E., Kitamura A., Shimohigoshi M., Watanabe T., Light-induced amphiphilic surfaces. Nature. 1997; 388: 431–432.
  7. 7. O’Regan B., Gratzel M., Low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991; 353: 737–739.
  8. 8. Fujishima A., Zhang X., Tryk D.A., TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008; 63 (12): 515–582.
  9. 9. Sakthivel S., Kisch H., Daylight photocatalysis by carbon-modified titanium dioxide. Angew. Chem. Int. Ed. 2003; 42 (40): 4908–4911.
  10. 10. Gnatyuk Y., Smirnova N., Eremenko A., Ilyinl V., Design and photocatalytic activity of mesoporous TiO2/Zr O2 thin films, Adsorp. Sci. Technol. 2005; 23(6); 497–408.
  11. 11. Smirnova N., Gnatyuk Y., Vityuk N., Linnik O., Eremenko A., Vorobets V., Kolbasov G., Nanosized TiO2-based mixed oxide films: sol–gel synthesis, structure, electrochemical characteristics and photocatalytic activity. Int. J. Mater. Eng. 2013; 3(6): 124–135.
  12. 12. Smirnova N., Vorobets V., Linnik O., Manuilov E., Kolbasov G., Photoelectrochemical and photocatalytic properties of mesoporous TiO2 films modified with silver and gold nanoparticles. Surf. Interface. Anal. 2010; 42: 1205–1208.
  13. 13. Linnik O., Smirnova N., Korduban O., Eremenko A., Gold nanoparticles into Ti1−xZnxO2 films: synthesis, structure and application. Mater. Chem. Phys. 2013; 142 (1): 318–324.
  14. 14. Linnik O., Kisch H., On the mechanism of nitrogen photofixation at nanostructured iron titanate films. Photochem. Photobiol. Sci. 2006; 5: 938–942.
  15. 15. Linnik O., Kisch H., Dinitrogen photofixation at ruthenium-modified titania films. Mendeleev Commun. 2008; 18(1): 10–11.
  16. 16. Shestopal N., Linnik O., Smirnova N., Influence of metal and non-metal ions doping on the structural and photocatalytic properties of titania films. Chem. Phys. Technol. Surf. 2015; 6(2): 203–210.
  17. 17. Sato S., Photocatalytic activity of NOx-doped TiO2 in the visible light region, Chem. Phys. Lett. 1986; 123: 126–128.
  18. 18. Asahi R., Morikawa T., Ohwaki T., Aoki K., Taga Y., Visible-light photocatalysis in nitrogen-doped titanium dioxide. Science. 2001; 293: 269–271.
  19. 19. Abadias G., Paumier F., Eyidi D., Guerin P., Girardeau T., Structure and properties of nitrogen-doped titanium dioxide thin films produced by reactive magnetron sputtering. Surf. Interface. Anal. 2010; 42: 970–973.
  20. 20. Nakano Y., Morikawa T., Ohwaki T., Yaga Y., Deep-level optical spectroscopy investigation of N-doped TiO2 films. Appl. Phys. Lett. 2005; 86: 132104.
  21. 21. Li Jinlong, Xinxin M., Mingren S., Li X., Zhenlun S., Fabrication of nitrogen-doped mesoporous TiO2 layer with higher visible photocatalytic activity by plasma-based ion implantation. Thin. Solid. Films. 2010; 519 (1): 101–105.
  22. 22. Pore V., Heikkila M., Ritala M., Leskela M., Arev S., Atomic layer deposition of TiO2−xNx thin films for photocatalytic applications. J. Photochem. Photobiol. A Chem. 2006; 177 (1); 68–75.
  23. 23. Zhao L., Jiang Q., Lian J., Visible-light photocatalytic activity of nitrogen-doped TiO2 thin film prepared by pulsed laser deposition. Appl. Surf. Sci. 2008; 254 (15): 4620–4625.
  24. 24. Kafizas A., Crick C., Parkin IP, The combinatorial atmospheric pressure chemical vapour deposition (c APCVD) of a gradating substitutional/interstitial N-doped anatase TiO2 thin-film; UVA and visible light photocatalytic activities. J. Photochem. Photobiol. A: Chem. 2010; 216 (2–3): 156–166.
  25. 25. Irie H., Watanabe Y., Hashimoto K., Nitrogen-concentration dependence on photocatalytic activity of TiO2−xNx powders. J. Phys. Chem. B 2003; 107(23): 5483–5486.
  26. 26. Serpone N., Is the band gap of pristine Tio2 narrowed by anion- and cation-doping of titanium dioxide in second-generation photocatalysts? J. Phys. Chem. B 2006; 110 (48): 24287–24293.
  27. 27. Chen X., Wang X., Hou Y., Huang J., Wu L., Fu X., The effect of postnitridation annealing on the surface property and photocatalytic performance of N-doped Ti O2 under visible light irradiation. J. Catal. 2008; 255 (1): 59–67.
  28. 28. Mitoraj D., Kisch H., The nature of nitrogen-modified titanium dioxide photocatalysts active in visible light. Angew. Chem. Int. Ed. 2008; 47(51): 9975–9978.
  29. 29. Zhang Z., Wang X., Long J., Gu Q., Ding Z., Fu X., Nitrogen-doped titanium dioxide visible light photocatalyst: spectroscopic identification of photoactive centers. J. Catal. 2010; 276 (2): 201–214.
  30. 30. Saha NC, Tompkins HG, Titanium nitride oxidation chemistry: an x‐ray photoelectron spectroscopy study. J. Appl. Phys. 1992; 72: 3072–3079.
  31. 31. Nosaka Y., Matsushita M., Nishino J., Nosaka AY, Nitrogen-doped titanium dioxide photocatalysts for visible response prepared by using organic compounds. Sci. Technol. Adv. Mater. 2005; 6(2): 143–148.
  32. 32. Cong Y., Zhang J., Chen F., Anpo M., Synthesis and characterization of nitrogen-doped TiO2 nanophotocatalyst with high visible light activity. J. Phys. Chem. C 2007; 111 (19): 6976–6982.
  33. 33. Beranek R., Kisch H., Tuning the optical and photoelectrochemical properties of surface-modified TiO2. Photochem. Photobiol. Sci. 2008; 7: 40–48.
  34. 34. Mitoraj D., Kisch H., On the mechanism of urea-induced titania modification. Chem. Eur. J. 2010; 16: 261–269.
  35. 35. Napoli F., Chiesa M., Livraghi S., Giamello E., Agnoli S., Granozzi G., Pacchioni G., Di Valentin C., The nitrogen photoactive centre in N-doped titanium dioxide formed via interaction of N atoms with the solid. Nature and energy level of the species. Chem. Phys. Lett. 2009; 477 (1–3): 135–138.
  36. 36. Livraghi S., Giamello MRCE, Magnacca G., Paganini MC, Cappelletti G., Bianchi CL, Nitrogen-doped titanium dioxide active in photocatalytic reactions with visible light: a multi-technique characterization of differently prepared materials. J. Phys. Chem. C 2008; 112 (44): 17244–17252.
  37. 37. György E., Pérez del Pino A., Serra P., Morenza JL, Depth profiling characterisation of the surface layer obtained by pulsed Nd:YAG laser irradiation of titanium in nitrogen. Surf. Coat. Technol. 2003; 173 (2–3): 265–270.
  38. 38. Chen X., Burda C., Photoelectron spectroscopic investigation of nitrogen-doped titania nanoparticles. J. Phys. Chem. B 2004; 108 (40): 15446–15449.
  39. 39. Diwald O., Thompson TL, Zubkov T., Goralski EGO, Walck SD, Yates JT, Photochemical activity of nitrogen-doped rutile TiO2(110) in visible light. J. Phys. Chem. B 2004; 108 (19): 6004–6008.
  40. 40. Reyes-Garcia EA, Sun Y., Reyes-Gil K., Raftery D., 15N solid state NMR and EPR characterization of N-doped TiO2 photocatalysts. J. Phys. Chem. C 2007; 111 (6): 2738–2748.
  41. 41. Beranek R., Kisch H., Surface-modified anodic TiO2 films for visible light photocurrent response. Electrochem. Commun. 2007; 9 (4): 761–766.
  42. 42. Nakamura R., Tanaka T., Nakato Y., Mechanism for visible light responses in anodic photocurrents at N-doped TiO2 film electrodes. J. Phys. Chem. B 2004; 108 (30): 10617–10620.
  43. 43. Mitoraj D., Berґanek R., Kisch H., Mechanism of aerobic visible light formic acid oxidation catalyzed by poly(tri-s-triazine) modified titania. Photochem. Photobiol. Sci. 2010; 9: 31–38.
  44. 44. Melhem H., Simon P., Wang J., Di Bin C., Ratier B., Leconte Y., Herlin-Boime N., Makowska-Janusik M., Kassiba A., BoucléJ., Direct photocurrent generation from nitrogen doped TiO2 electrodes in solid-state dye-sensitized solar cells: towards optically-active metal oxides for PV applications. Solar Energy Mater. Solar Cells. 2012; 117: 624–631.
  45. 45. Di Valentin C., Finazzi E., Pacchioni G., Selloni A., Livraghi S., Paganini MC, Giamello E., N-doped TiO2: theory and experiment. Chem. Phys. 2007; 339 (1–3): 44–56.
  46. 46. Mihailescu IN, Ristoscu C., Bigi A., Mayer I., Advanced biomimetic implants based on nanostructured coatings synthesized by pulsed laser technologies, Chapter 10. In: Laser-Surface Interactions for New Materials Production Tailoring Structure and Properties, Series: Springer Series in Materials Science, Vol. 130, Miotello A.; Ossi P.M. (Eds.), 2010, pp. 235–260.
  47. 47. Schneider CW, Lippert T., Laser ablation and thin film deposition, Chapter 5. In: Laser Processing of Materials Fundamentals, Applications and Developments, Schaaf P. (Ed.), Springer Series in Materials Science 139, Berlin, 2010, pp. 89–112.
  48. 48. Bauerle D., Laser Processing and Chemistry, Springer-Verlag, Berlin Heidelberg, 3rd ed., 2000.
  49. 49. Mihailescu IN, Gyorgy E., Pulsed laser deposition: an overview. In: International Trends in Optics and Photonics, Asakura T. (Ed.), Heidelberg, Springer, 1999, pp. 201–2014.
  50. 50. Eason R. (Ed.), Pulsed Laser Deposition of Thin Films: Applications-led Growth of Functional Materials, Wiley & Sons, 2007.
  51. 51. Craciun D., Bourne G., Zhang J., Siebein K., Socol G., Dorcioman G., Craciun V. Thin and hard ZrC/TiN multilayers grown by pulsed laser deposition. Surf. Coat. Technol., 2011; 205 (23–24): 5493–549.
  52. 52. Ristoscu C., Mihailescu IN, Biomimetic coatings by pulsed laser deposition, Chapter 7. In: Laser Technology in Biomimetics, Schmidt V., Belegratis M.R. (Eds.), 2013, Basics and Applications, Series: Biological and Medical Physics, Biomedical Engineering, Springer-Verlag, Heidelberg, New York, Dordrecht, London, pp. 163–191.
  53. 53. Socol G., Gnatyuk Y., Stefan N., Smirnova N., Sutan C., Malinovski V., Stanculescu A., Korduban O., Mihailescu I.N., Photocatalytic activity of pulsed laser deposited TiO2 thin films in N2, O2 and CH4. Thin Solid Films 2010; 518: 4648–4653.
  54. 54. Oh S., Moon K.-S., Moon J.-H., Bae Ji-M., Jin S., Visible light irradiation-mediated drug elution activity of nitrogen-doped Ti O2 nanotubes. J. Nanomater. 2013; 2013: Article ID 802318.
  55. 55. Wang H., Hu Y., The photocatalytic property of nitrogen-doped TiO2 nanoball film. Int. J. Photoenergy 2013; 2013: Article ID 179427.
  56. 56. Linnik O., Shestopal N., Smirnova N., Eremenko A., Korduban O., Kandyba V., Kryshchuk T., Socol G., Stefan N., Popescu-Pelin G., Ristoscu C., Mihailescu IN, Correlation between electronic structure and photocatalytic properties of non-metal doped TiO2/ZrO2 thin films obtained by pulsed laser deposition method. Vacuum. 2015; 115: 166–171.
  57. 57. Nakano Y., Morikawa T., Ohwaki T., Taga Y., Origin of visible-light sensitivity in N-doped TiO2 films. Chem. Phys. 2007; 339: 20–26.
  58. 58. Gnatuk Y., Smirnova N., Korduban O., Eremenko A., Effect of zirconium incorporation on the stabilization of TiO2 mesoporous structure. Surf. Interface Anal. 2010; 42: 1276–1280.
  59. 59. Chen H., Nambu A., Wen W., Graciani J., Zhong Z., Hanson JC, Fujita E., Rodriguez JA, Reaction of NH3 with titania: N-doping of the oxide and TiN formation. J. Phys. Chem. C 2007; 111 (3): 1366–1372.
  60. 60. Linnik O., Petrik I., Smirnova N., Kandyba V., Korduban O., Eremenko A., Socol G., Stefan N., Ristoscu C., Mihailescu I. N., Sutan C., Malinovski V., Djokic V., Janakovic D., TiO2/ZrO2 thin films synthesized by PLD in low pressure N-, C- and/or O-containing gases: structural, optical and photocatalytic properties. Digest J. Nanomater. Biostruct. 2012; 7(3): 1343–1352.
  61. 61. Katoh R., Furube A., Yamanaka K.-i, Morikawa T., Charge separation and trapping in N-doped TiO2 photocatalysts: a time-resolved microwave conductivity study. J. Phys. Chem. Lett. 2010; 1 (22): 3261–3265.
  62. 62. Dunnill CWH, Aiken ZA, Pratten J., Wilson M., Morgan DJ, Parkin IP, Enhanced photocatalytic activity under visible light in N-doped TiO2 thin films produced by APCVD preparations using t-butylamine as a nitrogen source and their potential for antibacterial films. J. Photochem. Photobiol. A: Chem. 2009; 207 (2–3): 244–253.
  63. 63. Oropeza FE, Harmer J., Egdell RG, Palgrave RG, A critical evaluation of the mode of incorporation of nitrogen in doped anatase photocatalysts. Phys. Chem. Chem. Phys. 2010; 12: 960–969.
  64. 64. Braun A., Akurati KK, Fortunato G., Reifler FA, Ritter A., Harvey AS, Vital A., Graule T., Nitrogen doping of TiO2 photocatalyst forms a second eg state in the oxygen 1s NEXAFS pre-edge. J. Phys. Chem. C 2010; 114 (1): 516–519.

Written By

Oksana Linnik, Nataliia Chorna, Nataliia Smirnova, Anna Eremenko, Oleksandr Korduban, Nicolaie Stefan, Carmen Ristoscu, Gabriel Socol, Marimona Miroiu and Ion N. Mihailescu

Submitted: 16 September 2015 Reviewed: 22 February 2016 Published: 24 August 2016