Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2019 Mar 17;31(7):2577–2589. doi: 10.1021/acs.chemmater.9b00257

Dispelling the Myth of Passivated Codoping in TiO2

Benjamin A D Williamson †,, John Buckeridge †,, Nicholas P Chadwick §,, Sanjayan Sathasivam §, Claire J Carmalt §, Ivan P Parkin §, David O Scanlon †,‡,⊥,*
PMCID: PMC6483321  PMID: 31031526

Abstract

graphic file with name cm-2019-00257x_0007.jpg

Modification of TiO2 to increase its visible light activity and promote higher performance photocatalytic ability has become a key research goal for materials scientists in the past 2 decades. One of the most popular approaches proposed this as “passivated codoping”, whereby an equal number of donor and acceptor dopants are introduced into the lattice, producing a charge neutral system with a reduced band gap. Using the archetypal codoping pairs of [Nb + N]- and [Ta + N]-doped anatase, we demonstrate using hybrid density functional theory that passivated codoping is not achievable in TiO2. Our results indicate that the natural defect chemistry of the host system (in this case n-type anatase TiO2) is dominant, and so concentration parity of dopant types is not achievable under any thermodynamic growth conditions. The implications of passivated codoping for band gap manipulation in general are discussed.

Introduction

TiO2, is an earth abundant, chemically stable wide band gap semiconductor that is used in a broad range of applications. In particular, titanium dioxide is used as a photocatalyst for the degradation of dyes and pollutants,1,2 the antimicrobial destruction of Staphylococcus aureus and Escherichia coli(36) as well as the splitting of water into H2(g) and O2(g)79 to name a few. TiO2 is popular in industry due to the wide range of scalable synthesis techniques such as sol–gel and chemical vapor deposition (CVD) that can be used to produce it.10 The main drawback, however, is its large band gap (anatase = 3.2 eV) meaning that solely UV light is absorbed accounting for only ∼1–2% of the total solar irradiation that reaches the earth’s surface.11 In addition to this constraint, efficient photocatalysis demands that the conduction band minimum (CBM) and valence band maximum (VBM) must straddle the redox potentials of the species to be reduced/oxidized (e.g., for water splitting ∼1.23 eV) with a suitable overpotential of ∼0.2–0.3eV thereby requiring a band gap of 1.7–2.2 eV for suitable visible light photocatalysis.11 Although the CBM of TiO2 is energetically favorable for the reduction of protons,12 the VBM is very low in energy (due to the high ionization potentials (IP) typical of n-type wide band gap materials13,14 dominated by deep lying O 2p states) producing highly localized holes states.15 Within the scientific literature, efforts to increase the efficiency of photocatalysis in TiO2 via band gap reduction have focused on “raising” the VBM through acceptor doping.1620

This is typically achieved through nonmetal doping, i.e., with B,18,2123 C,19,2426 N,16,17,27,28 or P.16,20,29 In particular, nitrogen has been the “go-to” dopant for enhanced photocatalysis in TiO2 since the work of Asahi et al. in 2001.16 The popularity of N-doping in anatase is also due to the reproducibility of these results through a range of solution-processed and deposition-based techniques, such as sol–gel,30,31 CVD,3234 or sputtering.3537 When doping anatase with nitrogen, it is assumed that the N 2p states hybridize with the top of the valence band, lowering the band gap and ionization potential. However, despite over a decade of research, the exact nitrogen species that forms within the anatase lattice is still scrutinized. The position of N has been suggested to be either substitutional (NO or NTi), interstitial (Ni), or part of a complex species with the lattice ions or unintentional dopants, such as H (NH4).3843 Interestingly, although experimental thin films of anatase display a red shifting of the band gap when nitrogen doped, N-doped rutile thin films display a blue shift. This has been proposed to be due to the denser crystal structure of rutile forming a wider O 2p valence band, and upon nitrogen doping a reduction in Coulombic repulsion is observed.42 In reality, however, the nitrogen states form highly localized hole states just above the VBM corroborated through theoretical studies on both substitutional and interstitial nitrogen species.38,4043 Within the density functional theory (DFT) literature, multiple groups have managed to show that interstitial N displaces from the perfect interstitial site toward a lattice oxygen to form an N–O dumbbell which has also been corroborated using electron paramagnetic resonance (EPR) measurements.43 Unfortunately, these localized N states are a source of electron–hole recombination especially at the high concentrations typically required for “band gap reduction”.44,45 Theoretical studies have also proposed the inducement of oxygen vacancies due to nitrogen incorporation providing intrinsic compensation typical of the high ionization potential of TiO2 polymorphs.12,42

Reports of beneficial carbon doping also exist in the literature, albeit less explored than nitrogen-doped TiO2 in both experiment and theory. Carbon possesses small ionic radii (∼0.15–0.16 Å46) and generally prefers to exist as a 4+ cation (CH4, CO2 etc.) although it can also exist in the 4– oxidation state as in TiC. Experimental techniques, such as X-ray photoelectron spectroscopy,47,48 suggest substitutional carbon (CO) species or carbonate species (CO32–) adsorbed into the lattice. Two theory papers, in particular, one by Di Valentin et al.48 (using the standard Perdew–Burke–Ernzerhof functional) and a more recent hybrid DFT paper (HSE06) by Zhang and co-workers49 elucidate the possible locations and downsides to carbon doping in anatase. Both Di Valentin et al. and Zhang and co-workers recognize the existence of carbon replacing Ti in the lattice (CTi) as a dominant defect whereby it acts as an isovalent dopant neither adding or removing electrons but possibly reducing the band gap at higher doping concentrations.48,49,50 Due to the typical coordinations of carbon, it was found that CTi defects prefer a four-coordinate configuration adding strain to the system due to the six-coordinate environment of Ti (where Ti can hybridize 3d orbitals). Due to the small ionic radii, interstitial carbon possesses relatively low formation energies under both Ti-poor/O-rich and Ti-rich/O-poor conditions where it acts as a donor. CO is only expected to form toward highly O-poor regimes where it induces the formation of oxygen vacancies thereby compensation occurs as with nitrogen defects. The defects that are expected to give states in the band gap suitable for visible light excitation are the CO related species which do not form as readily as substitutional nitrogen species.49

Transition metals can also act as acceptor dopants in particular Ni,5153 Fe,5457 Co,5860 Cr,6164 and Cu,6568 for example. It has been proposed that transition metal dopants provide an additional service by acting as an electron (or hole) trap enhancing the efficiency.6972 Whether this increases or decreases electron–hole separation is still in debate with certain metal cations increasing the recombination rate (i.e., chromium) and others decreasing (i.e., copper and iron).45,73 Beneficial effects are not just limited to acceptor doping as donors such as Nb,10,74 Ta,7577 Sb,78,79 or W80,81 also displays an enhancement of the photoactivity of anatase despite no band gap reduction or states within the band gap. The enhanced effects could likely be due to the increased electron concentration and mobility realized through donor doping and thus an increase in reducing electrons or even an enhanced surface segregation as seen with Sb and Ta.82,83

To rectify the recombination issues encountered from doping with nonmetals, an approach to fully passivate the dopant states while retaining the reduced band gap was proposed by Wei and co-workers.44 Within this formalism, the additional incorporation of a donor to the anatase lattice should have the effect of annihilating the additional holes brought about by acceptor doping and is shown schematically in Figure 1. Thus, the beneficial band gap narrowing exhibited by N-doped TiO2 can be retained while alleviating the recombination centers. Due to the location of the anatase CBM relative to the reduction potential,12 a preferable resonant donor is required so that no downward shift of the CBM occurs. This makes Nb, Ta, Mo, and W ideal donors for this effect.44,84 The original study by Gai et al. proposed four different codoped systems, [V + N], [Nb + N], [Cr + C], and [Mo + C].44 Following from this study, the concept of “passivated codoping” became very popular, and each of the four proposed codoping pairs has been studied extensively experimentally ([Nb + N],8589 [Ta + N],90,91 [Cr + C],92 [Mo + C]9396) with notable examples such as [Nb + N] showing promising results such as the 7-fold increase in photocatalytic degradation of methylene blue or the 4-fold increase in decomposition of methyl orange with [Cr + C] codoped anatase.88,93 The term “full passivation” is used in this context to describe the concentration parity of acceptor and donor dopants.

Figure 1.

Figure 1

Schematic diagram for passivated codoping in TiO2 showing the positions of the band extrema relative to the redox potentials of water (green) and the acceptor (grey) and donor states (red).

Numerous computational studies have been carried out on codoped systems for band gap engineering,44,84,97109 in particular a computational screening by Yan and co-workers found that [Nb + N] and [Ta + N] were ideal at a “high alloying concentration” regime, and that both [Mo + 2N] and [W + 2N] were ideal in a “low alloying concentration” regime. Carbon codoped systems were suggested to reduce the band gap too much, which combined with the problems associated with doping with C mean that reproducibility will be an issue in this context, despite a few promising results.93,95

Within the codoping literature, there exists a lack of distinction between the increased photoactivity seen due to passivated codoping and that seen from individually doping with either nonmetal or metal aliovalent ions. Claims are also made about the increase in photocatalytic activity such as with the 7-fold increase and 4-fold increase of [Nb + N] and [Mo + C] systems of which the degradations of methylene blue and methyl orange rely just as much on reduction processes favouring an increased mobility and concentration of conduction band electrons.88,93 Methylene blue and methyl orange photocatalytic tests also rely on multistage degradation which can proceed via adsorption, therefore, making the difference between photocatalytic electron/hole transfer indistinguishable from other processes. Bartlett also remarked that stoichiometric pairing or full passivation of Nb + N was not achieved, even providing a follow up study on the composition dependence of Nb and N.89 This work concluded that in fact a larger incorporation of Nb both decreased the optical band gap and increased the photoactivity and is echoed by Chadwick et al.86,89 Zhang et al. also noted that although [Mo + C] showed an increase in the rate of methylene blue degradation compared to the undoped and carbon-doped TiO2 systems, they also found that C incorporates preferentially on the Ti site and thus cannot be deemed to provide full passivation.95 An analysis of the ionic radii of dopants in TiO2 shows that for Nb, Ta, W, Cr, Mo, and W, a range of 0.59–0.64 Å is observed (compared to 0.60 Å, for Ti) making these highly soluble dopants, therefore low formation energies would be expected.46

In this work, we demonstrate through theoretical calculations of the [Nb + N] and [Ta + N] codoped systems that charge compensation is simply not possible in anatase TiO2 under thermodynamic equilibrium due to its inherent n-type nature.

Computational Methodology

Density functional theory (DFT) calculations using the periodic code, VASP (Vienna Ab-initio Simulation Package),110113 and the hybrid HSE06 (Heyd–Scuseria–Ernzerhof)114 functional were performed on the bulk geometric and electronic relaxations as well as all studied intrinsic and extrinsic defects of anatase TiO2. The intrinsic defects and dopants (Ta, Nb, N) were simulated using 3 × 3 × 2 supercell expansions (∼108 atoms) of the geometrically relaxed conventional cell of anatase TiO2. The HSE06 hybrid functional115,116 provides excellent descriptions of the electronic and structural properties of all of the known polymorphs of TiO2,10,12,117123 and hybrid functionals are, in general, a marked improvement on local or semilocal DFT functionals as they better deal with the self-interaction error.124 The projector augmented wave125 method was utilized to describe the interactions between the core (Ti[Ar], O[He], Ta[Xe], Nb[Kr], N[He]) and valence electrons.

An initial geometry optimization of the bulk conventional cell was carried out minimizing the volume, lattice parameters, cell angles, and the ions within the cell until the forces acting on all of the atoms were less than 10 meV Å–1. A 700 eV plane-wave energy cutoff and a Γ-centered 7 × 7 × 5 k-point mesh were sufficient to calculate the electronic and structural properties. The bulk electronic properties can be found in ref (123). From this, a 3 × 3 × 2 supercell containing 108 atoms was formed to evaluate the defect properties of the material. Geometry optimizations of each defective supercell and its respective charge states involved calculating the relaxation of only the ions within the cell, keeping the cell volume, lattice parameters, and angles fixed. A plane wave energy cutoff of 450 eV, Γ-centered 2 × 2 × 2 k-point mesh, and spin polarization were used to relax the supercells to the same force convergence as for the conventional bulk. The anatase lattice and dopant states are restricted by the formation of limiting phases (as discussed in the Supporting Information, SI) and are tabulated in Table 1. Table 1 also displays the k-point meshes used in the geometry optimizations (force convergence criterion: <10 meV Å–1) and the calculated enthalpies of formation (ΔHf). These enthalpies of formation are in good agreement with the standard temperature and pressure experimental results, with differences expected due to the difference in temperature (DFT formation energies are calculated at 0 K) and may be due to the choice of the HSE06 functional for these limits.

Table 1. Tabulated Chemical Potential Limits as Calculated with the HSE06 Hybrid Functional Together with Their Associated Spacegroup Symmetry and Structural Relaxation Parametersa,b.

limit spacegroup k-point mesh ΔHf (eV)
TiN Fmm 8 × 8 × 8 –3.49 (−3.50)
Ti2N P42/mnm 5 × 5 × 8 –4.07 (N/A)
NO2 R 4 × 4 × 4 –0.26 (N/A)
N2O5 Cm 5 × 5 × 4 –0.13 (N/A)
Nb2O5 I4/mmm 2 × 2 × 2 –18.48 (−19.69)
NbO2 C2/c 3 × 3 × 3 –7.64 (−8.24)
NbO Pm3m 6 × 6 × 6 –4.10 (−4.35)
Ta2O5 I41/amd 6 × 6 × 1 –20.31 (−21.21)
a

All k-point meshes are Γ-centered.

b

The calculated enthalpies of formation ΔHf for the relaxed structures are also tabulated with experimental values126 in parentheses.

Defect Formalism

Equation 1 defines the enthalpy of formation for a defect (D) in charge state, q.

graphic file with name cm-2019-00257x_m001.jpg 1

Where

graphic file with name cm-2019-00257x_m002.jpg

ED,q is the total energy of the defective supercell in charge state q and EH is the total energy of the host supercell. Ei and μi refer to the elemental reference energy and chemical potentials of species “i” (Ti(s), O(g), Ta(s), Nb(s), and N(g)), respectively. ni is either negative or positive depending on whether the species is added to or removed from the system. The Fermi energy (EF) ranges from the valence band maximum (VBM) to the conduction band minimum (CBM). The band gap (Eg) of anatase has been determined to be 3.41 eV from our primitive cell calculations consistent with other ab initio works.12,127132 ϵVBMH refers to the eigenvalue of the VBM from the host supercell calculation, and ΔEpot is a potential alignment term to account between the difference between electrostatic potentials between the defective and host supercells. To account for the finite size of the supercell, two corrections are applied; Ecorr and EcorrBF. The former correction is an image-charge correction which is necessary due to the long-ranged nature of the Coulomb interaction,133,134 thus correcting for the interaction of the defect with its periodic images. The scheme used herein utilizes the formalism by Hine and Murphy135 based on the Lany and Zunger correction.136 The latter correction, Ecorr, corresponds to a band filling correction136,137 associated with the relatively high defect concentrations that give rise to an unphysical band gap renormalization present for the finite-sized supercell calculation.

Thermodynamic Limits

To determine the enthalpies of formation for a given defect over a chemical potential range relating to the equilibrium growth conditions of the host material (TiO2), competing phases must be calculated to create upper and lower bounds on the phase space.

The chemical potential bounds placed on the host can be calculated through the enthalpies of the formation of anatase TiO2

graphic file with name cm-2019-00257x_m003.jpg 2

Typically, the O-poor limit is determined by the formation of a secondary phase, Ti2O3, however, in practice, this is difficult to reach under general deposition temperatures and oxygen partial pressures in experiment. The O-poor limit is thus rationalized to be in the region of μO = −2 eV,138,139 which is what we will use as our Ti-rich/O-poor limit in this study. The Ti-poor/O-rich boundary is limited via the formation of oxygen gas (O2(g)); thus, the chemical potential ranges for Ti and O are

graphic file with name cm-2019-00257x_m004.jpg

and

graphic file with name cm-2019-00257x_m005.jpg

The limiting phases for Nb and N are shown in SI Figure 1 and demonstrate that under the chemical potential range, Nb2O5 is the limiting phase for μNb and that under Ti-rich/O-poor and Ti-poor/O-rich conditions, N2(g) and NO2 are the limiting phases, respectively. For μTa, the limiting phase is Ta2O5. Further limits are placed upon the dopant species added to the system. The dopant chemical potentials are thus: Ti-poor/O-rich

graphic file with name cm-2019-00257x_m006.jpg

Ti-rich/O-poor

graphic file with name cm-2019-00257x_m007.jpg

The thermodynamic transition levels (Figure 2) display the thermal equilibrium transition from charge state q to q′ at a specific Fermi level and can be calculated using the equation

graphic file with name cm-2019-00257x_m008.jpg 3

Such transitions may be seen using experimental techniques such as deep level transient spectroscopy.140

Figure 2.

Figure 2

Calculated thermodynamic transition levels for both Nb, Ta, and N-doped TiO2 under Ti-rich/O-poor growth conditions and under the Ti-poor/O-rich limit growth conditions. The Fermi energy ranges from the VBM to the CBM (∼3.41 eV).

Equilibrium Concentrations

The equilibrium defect and carrier concentrations were determined through calculation of the self-consistent Fermi energy, as implemented in the code SC-FERMI.141,142 The concentration of defect D in charge state q, [D,q] is given by

graphic file with name cm-2019-00257x_m009.jpg 4

where ND is the density of sites on which the defect in question can form, gD,q is the degeneracy of the defect state, k is Boltzmann’s constant, and T is the temperature. Electron (n0) and hole (p0) concentrations are determined using the equations

graphic file with name cm-2019-00257x_m010.jpg 5
graphic file with name cm-2019-00257x_m011.jpg 6

where ρ(E) is the density of states (DOS), Eg is the band gap, fe(E) = [exp((EFE)/kT) + 1]−1 is the Fermi-Dirac distribution function, and fh(E) = 1 – fe(E) (the energy scale is defined so that the VBM is zero). The defect and carrier concentrations are functions of EF via ΔH(D,q), fe(E), or fh(E); given the constraint of overall charge neutrality, it is therefore possible to determine the self-consistent value of EF and hence the equilibrium concentrations at T.

Results

Defect Thermodynamics

To understand the thermodynamic defect chemistry of codoped anatase TiO2, density functional theory using the hybrid functional, HSE06114 was utilized. HSE06 has previously been shown to accurately reproduce the electronic and geometric structures of TiO2 polymorphs relative to experiment.10,12,118123 An analysis of the density of states (DOS), geometry, and band structure of anatase TiO2 has been carried out previously in ref (123). The thermodynamic transition level diagrams for the intrinsic and extrinsic defects are shown in Figure 2. These are depicted under two different growth regimes, an experimentally viable Ti-rich/O-poor and an Ti-poor/O-rich limit.

Intrinsic Defects

The dominant acceptor (VTi) and donor (VO) defect species were calculated within the anatase lattice.118,143147 Under the n-type “favorable” conditions of Ti-rich/O-poor conditions, VO is the lowest formation energy intrinsic defect (ΔHf(VO) = ∼2.71eV), and under Ti-poor/O-rich growth conditions (p-type favorable), this raises to ∼4.71 eV. In TiO2, VO is a resonant two-electron donor where the 2+/0 transition level exists ∼0.12 eV above the CBM (∼3.41 eV). This is in contrast to SnO2, ZnO, and BaSnO3 where the 2+/0 transition level is in the band gap acting as a deep donor and is not expected to contribute to the electronic conductivity of these materials.14,148155

The titanium vacancy (VTi), under Ti-poor/O-rich conditions where its formation is most conducive still has a large formation energy in the neutral charge state of ∼7.23 eV. This value rises to ∼11.24 eV under Ti-rich/O-poor conditions. Under both sets of conditions, the 0/1– transition level occurs ∼0.87 eV above the VBM, making VTi a very deep acceptor. Under O-rich conditions, VTi4– begins to compensate VO ∼2.28 eV above the VBM trapping the Fermi level around this point. The position of compensation by VTi4– under Ti-rich/O-poor conditions occurs ∼0.06 eV below the CBM.

To assess the ability of fully passivated codoping in TiO2, the defect thermodynamics of the TiO2:[Nb + N] and TiO2:[Ta + N] systems were analyzed and compared. These systems were chosen due to their abundance in the literature and from previous theory calculations identifying them as ideal candidates for water splitting.44,8491,100,156,157 Both Nb and Ta have been shown to easily dope into TiO210,7477,158 and success has been seen in N-doped TiO2.16,17,27,28 The calculations of these defects were carried out alongside the possible codoped clusters in TiO2 using the hybrid HSE06115,116 functional.

Mono Doping of Anatase with N, Nb, and Ta

All mono-doped anatase TiO2 systems were analyzed to determine the thermodynamic viability of the chosen defects and any points of unintentional compensation. Figure 2 displays the transition levels for the N-, Nb-, and Ta-related defects in anatase.

Nitrogen Doping

Within the anatase lattice, N can exist both substitutionally and interstitially. Substitutional nitrogen, NO, acts as an amphoteric defect under both growth regimes with a relatively high formation energy (neutral charge state) of 3.08 and 5.34 eV under Ti-rich/O-poor and Ti-poor/O-rich conditions, respectively. Although O-rich conditions typically favor the formation of acceptor defects, NO is the highest in the formation energy under this growth regime. This is due to the higher μO and thus lack of available oxygen sites reflected in the higher formation energy of VO. NO is both an ultradeep acceptor (0/1– transition ∼1.70 eV above the VBM) and donor defect (1+/0 transition ∼2.24 eV below the CBM) contributing no source of p or n-type conductivity in anatase. Due to the intrinsic n-type nature of anatase, the Fermi level is likely to exist toward the CBM and, as such, NO will act as an acceptor. Figure 3a displays the partial charge density for the neutral charge state of NO showing the hole localized on the N 2p orbital. These results are consistent with other HSE06 studies on TiO2:N.38,39 Intrinsic compensation occurs from VO just around 0.34 and 0.25 eV below the CBM for Ti-rich/O-poor and Ti-poor/O-rich conditions, respectively. In reality, under O-rich conditions, this point would not be realizable due to the Fermi level being pinned in the band gap around 2.28 eV above the VBM from the crossing of VTi4– and VO2+.

Figure 3.

Figure 3

Partial charge densities of (a) NO0 and (b) (NO)O as viewed along the [100] direction and (c) Nb(Ta)Ti0 as viewed along the [001] direction. The anatase lattice (Ti = gray, O = black) is depicted by the wire-frame lattice, and the nitrogen and Nb(Ta) species, by the colored orbs (color coded to Figure 2). The charge isosurfaces for holes are shown in blue and plotted from 0 to 0.02 eV Å–1 and the electron charge density in orange and plotted from 0 to 0.002 eV Å–1.

Nitrogen relaxes from the “ideal” interstitial point to form a “split-interstitial” producing an N–O dumbbell situated on an O atom site (denoted (NO)O). This behavior is observed in the previous theoretical literature and is supported by the presence of NO paramagnetic species in electron paramagnetic resonance (EPR).38,39,4143,159 The formation energies of (NO)O0 remain stable across the entirety of the chemical potential range with similar values of 4.48 and 4.73 eV under Ti-rich/O-poor and Ti-poor/O-rich conditions, respectively. (NO)O, like NO, can act as an electron acceptor and as a donor. At Fermi energies close to the VBM, (NO)O acts as a three-electron donor and becomes the dominant nitrogen species in TiO2. (NO)O undergoes, a 3+/2+ and a 2+/0 transition level around 2.82 and 1.63 eV below the CBM, respectively. In anatase, however, the Fermi level is likely to exist toward the CBM at which point (NO)O acts as an acceptor. Despite having the ability to produce three holes in TiO2, (NO)O only undergoes one transition level in the band gap (0/1– transition at 2.07 eV above the VBM). (NO)O can, therefore, be said to only accept one hole, consistent with the previous theory.38,41,42Figure 3b shows neutral charge state of (NO)O displaying the localized hole bound to the N–O dumbbell. Under Ti-poor/O-rich conditions, it is likely that (NO)O will be the dominant nitrogen species across the entirety of the band gap and explains the defect’s appearance in the experimental literature. Under Ti-rich/O-poor, however, it is likely that a greater concentration of substitutional nitrogen will be seen when the Fermi level is trapped just below the CBM.

Nb and Ta Doping

NbTi and TaTi are the lowest formation energy defects under both growth regimes having formation energies of 0.38 and 0.22 eV, respectively, under Ti-rich/O-poor conditions. Under a Ti-poor/O-rich growth regime, the formation energies rise to 1.38 and 1.22 eV for NbTi and TaTi, respectively. This is likely due to the similar ionic radii of Nb and Ta (∼64 pm) to Ti (∼61 pm)46 causing minimal strain to the anatase lattice upon incorporation of these dopants (the six surrounding O atoms move ∼1–1.5% away from the dopant center). Both substitutional Nb and Ta act as resonant donors donating one electron to the conduction band with the 1+/0 transition levels occurring above the CBM. The electron density is, therefore, delocalized over the Ti 3d orbitals that make up the CBM (Figure 3c). These results are indicative of the high conductivities seen in experiment TiO210,158 and is reproduced in the theoretical literature.119,160,161 Under Ti-poor/O-rich conditions, Nb/TiTi is compensated for by VTi4– ca. 0.66 and 0.63 eV below the CBM for Nb and Ta, respectively. Under a Ti-rich/O-poor growth regime, however, no intrinsic compensation occurs in the band gap.

Realization of Fully Passivated Codoping?

To assess the viability of fully passivated codoping in anatase TiO2, two different dopant clusters were tested in this study. It is clear from doping with N that under O-poor conditions, NO is the dominant nitrogen defect across the band gap, and under O-rich conditions, the interstitial (NO)O dumbell is preferred. The dopant clusters proposed are thus: [MTi + NO] and [MTi + (NO)O] (where M = Nb or Ta). Despite interstitial nitrogen formally inducing three holes, due to the nature of the N–O species, only one hole is available and acts as an acceptor at Fermi levels toward the CBM (as (NO)O is in the 1– charge state). For full passivation of both clusters, only one substitutional Nb(Ta) defect is, therefore, required. Both defect clusters showed a preferential “near” configuration (by ∼0.13, 0.04, 0.18, and 0.07 eV for [NbTi + NO], [TaTi + NO], [NbTi + (NO)O], and [TaTi + (NO)O], respectively) as is to be expected due to the Coulombic attraction. The thermodynamic transition levels show that Nb and Ta clusters possess very similar formation energies, which is to be expected considering the ionic radii46 and formation energies of the NbTi and TaTi.

Under O-poor conditions, the [Nb(Ta)Ti + NO] clusters possess a formation energy of 1.60 eV (Ta = 1.53 eV) in the neutral charge state. The density of states (DOS) in SI Figure 2a,b shows the spin-up and spin-down occupied states around the Fermi level (for the cluster supercell) (0.36 and 0.41 eV above the VBM for [Nb + N] and [Ta + N], respectively) which gives rise to the 1+ and 2+ charge states whereby each electron is removed. The two transition levels (2+/1+ and 1+/0) for [Nb(Ta)Ti + NO] occur in the band gap: the 2+/1+ at ∼2.93 eV (2.63 eV) and the 1+/0 ∼1.95 eV (1.89 eV) below the CBM for Nb and (Ta), respectively. Under O-rich conditions, [Nb(Ta)Ti + NO] possesses a relatively high formation energy of 4.86 eV (4.79 eV).

[Nb(Ta)Ti + (NO)O] possesses relatively high formation energies under both O-poor and O-rich conditions of 3.47 eV (3.41 eV) and 4.73 eV (4.67 eV), respectively, when in the neutral charge state. As with [Nb(Ta)Ti + NO], [Nb(Ta)Ti + (NO)O] undergoes two transition levels in the band gap. The DOS for each cluster is shown in SI Figure 2c,d displaying a similar situation as with [MTi + NO] whereby the N and O states appear around 0.61 eV above the VBM for both clusters, respectively. This can also be thought of as removing electrons from the antibonding π orbital in an N−O molecular orbital diagram. For the [Nb(Ta)Ti + (NO)O] cluster, the 2+/1+ and 1+/0 transitions occur at 2.93 eV (2.94 eV) and 1.71 eV (1.73 eV) below the CBM, respectively, for the Nb and (Ta) systems.

Figure 4a,b shows the partial charge densities for both [MTi + NO](q = 0) and [MTi + (NO)O](q = 0) (where M = Nb and Ta), respectively, for the filled states, as shown in the DOS in SI Figure 2. In Figure 4a, the localized hole is now passivated with the donated electron. The likely improvement in visible light activity in TiO2 is possibly due to the excitation of the electron in the filled N orbital to the CBM which occurs in visible light.38 A similar situation occurs in Figure 4b whereby the delocalized hole state on the N–O dumbbell is filled by the donated electron from Nb(Ta)Ti.

Figure 4.

Figure 4

Partial charge densities for (a) [Nb(Ta)Ti + NO] and (b) [Nb(Ta)Ti + (NO)O] as viewed along the [100] direction. The anatase lattice is portrayed through the wire-frame model (Ti = gray, O = black) and the dopant species by the colored spheres (colored coded to Figure 2). The electron partial charge density is shown in orange and is plotted from 0 to 0.02 eV Å–3.

Under Ti-rich/O-poor conditions, it is clear that when comparing the neutral charge states, the formation energy of nitrogen defects is reduced in the presence of substitutional Nb and Ta defects. Despite this, the defect clusters retain high formation energies compared to NbTi and TaTi and at Fermi levels at the CBM both NO and (NO)O are both preferable (by 0.23 and 0.34 eV, respectively, for [Nb + N] and by 0.16 and 0.28 eV, respectively, for [Ta + N]). Under Ti-poor/O-rich growth conditions, this effect is more pronounced and NO(q = 1−) and (NO)O(q = 1−) were lower in energy by 1.23 and 1.34 eV, respectively, for [Nb + N] and by 1.16 and 1.28 eV for [Ta + N], respectively. This suggests that cluster formation will be negligible under thermodynamic equilibrium under all conditions.

Binding energies (EBE) for the clusters were calculated using the equations

graphic file with name cm-2019-00257x_m012.jpg 7

and

graphic file with name cm-2019-00257x_m013.jpg 8

Where ΔHf[Nb(Ta)+N]p, ΔHf, and ΔHfNOn or(NO)O are the enthalpies of formation of the cluster and substitutional Nb(Ta) and N, respectively. p, m, and n refer to the charge states of each defect state and EF is the Fermi energy. As EF occurs around the CBM, the last term in this case is equal to 0 as p = 0, m = +1, and n = −1, respectively. For each case, positive binding energies indicate the favorable formation of the defect clusters. The binding energies for [NbTi + NO] and [TaTi + NO] are 0.17 and 0.08 eV, respectively, indicating a small preference for complex formation, however, the formation energy difference between NO and the substitutional n-type defects is likely to significantly negate this effect such that no 1:1 compensation will occur. For [NbTi + (NO)O], a similarly small binding energy of 0.07 eV exists; however for [TaTi + (NO)O], the binding energy is −0.04 eV indicating that the complex formation is unfavorable.

Defect and Carrier Concentrations

The calculated equilibrium defect and carrier concentrations as a function of temperature (T) are shown in Figure 5 for the cases of Nb and N codoping (left panel) and Ta and N codoping (right panel). As can be seen from Figure 2, NbTi and TaTi have very similar formation energies, as do the complexes involving the different metal dopants with N impurities. As a consequence, the concentrations for both cases are quite similar, although the TaTi have slightly higher concentrations at each T than the NbTi due to their lower formation energy (by 0.16 eV). From Figure 5, we see that in each case, the NbTi or TaTi impurities dominate, as expected from their low formation energies. Under O-poor conditions, the dopants are compensated predominantly by electron carriers, resulting in a strongly n-type system, with EF close to the conduction band minimum (see the insets). For the Nb case (Ta case) at T > 350 K (T > 250 K), [NbTi] > n0 ([TaTi] > n0), indicating that the activation of the donor dopants reduces somewhat counterintuitively at higher T. The reason for this reduction is that, although at low T the 1+ charge state is most favorable and we have [MTi] ≈ [MTi1+] = n0 (where M = Nb or Ta), at higher T the 0 and 1– charge states become populated (as the (1+/0) and (0/1−) transitions occur close to the CBM), such that [MTi] = [MTi] + [MTi0] + [MTi], and [MTi1+] is charge balanced by n0 + [MTi]. In O-rich conditions, the metal dopants are compensated by VTi4–, and EF becomes trapped close to where the formation energies of the dopants and the vacancies are equal, at about EF = 2.7 eV (see insets in the bottom panels of Figure 5). The system is, therefore, insulating under these conditions.

Figure 5.

Figure 5

Equilibrium concentrations of metal impurities MTi (red line, where M = Nb or Ta), titanium vacancies (VTi, blue line), substitutional N on O sites (NO, green line), complexes of substitutional nitrogen and metal impurities (MTi + NO, dotted teal line, where M = Nb or Ta) and electron carriers (n0, dashed orange line) calculated as a function of temperature. The left panel gives the concentrations for Nb doping, the right for Ta doping. The upper and lower panels correspond to O-poor and O-rich conditions, respectively. The insets show the self-consistent Fermi level (EF, magenta line) as a function of temperature, with the zero of the energy scale set equal to the valence band maximum and the dashed line indicating the position of the conduction band minimum.

The most notable feature from Figure 5 is the lack of significant concentrations of N impurities across all values of T, which is a consequence of the relatively high formation energies of the related defects. This can, in part, be related back to equation 1 and the chemical potential limits imposed on the defects by TiO2. For the nitrogen-related defects, the cost of N–N bond breaking is high, thus the change in μN within the chemical potential limits of TiO2 is low compared to μTa or μNb. Concentrations above 1017 cm–3 are only observed for T > 1200 K, and we predict that many of those N atoms will be incorporated as complexes, but with significantly smaller concentrations than isolated Nb or Ta dopants, which act as shallow n-type donors. Note that in this analysis, we assume all defects are at the dilute limit, and we, therefore, do not include defect–defect interactions, but considering the low binding energies we calculate for the complexes, we expect that such interactions will have a minor effect on the computed concentrations. The key point is the relatively high formation energies of N-related defects, whether they be isolated or in complexes. Our results clearly demonstrate that Nb or Ta dopants will dominate, which results in an n-type system under O-poor conditions or an insulating system under O-rich conditions, due to the compensation by VTi. Assuming defect formation occurs under thermodynamic equilibrium, the N dopants will only play a minor role in the defect chemistry.

Discussion

The realization of fully passivated codoping in [Nb + N] of [Ta + N] codoped TiO2, however, is certainly unobtainable due to the inherent preference for n-type defects and dopants in anatase TiO2. As seen in Figures 2 and 5, under both growth regimes, substitutional Nb(Ta) are the dominant defects and are far lower in energy than any of the nitrogen or cluster species. The implication of this work is that although it is possible for the codoped clusters to form within the lattice (as shown by the binding energies), there will always be a higher concentration of n-type dopant and thus there will be no possibility to produce a fully passivated system. The lack of quantitative data on dopant concentrations in the previous experimental literature makes it impossible to ascertain whether full compensation has occurred.87,90,91,100,156,157 Certain studies do, however, acknowledge the lack of dopant parity such as those by Bartlett et al. and Chadwick et al.86,88,89 It should be noted that the majority of synthesis studies, which have attempted to produce codoped anatase TiO2, have utilized growth conditions that are nonequilibrium in nature, and where kinetic effects can play a large role. None of these methods, however, has managed to reach full passivation. Notably, CVD grown films of [Nb + N] codoped TiO2, where kinetics are known to be hugely important, have failed to come close to full passivation.86

These results open up a further question: is compensated codoping of TiO2 possible at all? In terms of the other potential co-dopant pairs, [W + C] and [Mo + C], we consider that in both of these cases, the formation energy of the donor dopant will be much lower than that of any acceptor-based C-related defect in anatase TiO2. Indeed, previous theoretical results have indicated that carbon interstitials are the dominant C-related defect under most growth conditions, and hence C will act as a donor and not as an acceptor in a compensated codoping pair which would raise the VBM of TiO2. These insights beg the question of whether passivated codoping as a band gap modification strategy in any material is achievable?

The answer to this question perhaps lies most simply in an analysis of the absolute band alignments of the band edges of the host system to be doped, relative to the vacuum level, Figure 6. TiO2 is a natively n-type semiconductor, with a large electron affinity and a large ionization potential (IP). Systems that share this type of band alignment are almost always n-type in nature, such as SnO2, ZnO, In2O3, etc. It has been demonstrated repeatedly by both theory and experiment that p-type doping in these systems is not achievable.13,14,162,163 It follows that efforts to produce passivated codoping in these and similar systems would be fighting against thermodynamics, as the n-type dopant will always be dominant over the p-type dopant.

Figure 6.

Figure 6

Band alignment of anatase TiO2164 compared to a selection of n-type and p-type semiconductors: ZnO,165 SnO2,166 In2O3,167 Cu2O,168 CuAlO2,169 SnO.170 The ionization potentials for each compound are displayed relative to the vacuum level (0 eV).

Materials with a lower IP display a higher probability of effective incorporation of acceptor type defects, as is seen for natively p-type materials, such as Cu2O, CuAlO2, and SnO.171175 These type of systems often possess large band gaps, however, which means that although VBM positions are perhaps more ideal relative to the redox potential for water splitting, the same problems of weak light absorption exist as in n-type TiO2, and effective passivated codoping to lower the CBM while retaining the VBM position would face the inverse of the problems that we run into for codoping of TiO2 i.e. p-type defects will dominate over n-type defects. It is, therefore, likely that in practice only a truly bipolar material, which can be successfully doped both n-type and p-type, could approach full passivation. An additional benefit of bipolar materials is that they generally possess band gaps that are already within the visible light active range, and so passivated codoping is not likely to be necessary. Therefore, the future for passivated codoping is quite unclear.

Conclusions

Using hybrid density functional theory, an analysis of the thermodynamic properties of two archetypal systems: [Ta + N] and [Nb + N] codoped anatase TiO2, has been undertaken demonstrating that passivated codoping of anatase TiO2 is not possible. Analyzing the defect chemistry of these systems demonstrates that a severe doping asymmetry arises due to the inherent n-type nature of anatase. This asymmetry always favors incorporation of a higher concentration of the electron donating dopant, meaning that full passivation is not achievable for a system grown at thermodynamic equilibrium.

Acknowledgments

This work made use of the ARCHER U.K. National Supercomputing Service (http://www.archer.ac.uk) via our membership of the U.K.’s HEC Materials Chemistry Consortium, which is also funded by the EPSRC (EP/L000202). B.A.D.W. would like to acknowledge the use of the “bapt” python-package by A. Ganose (https://github.com/utf/bapt) to plot the band alignment in Figure 6. The UCL Legion and Grace HPC Facilities (Legion@UCL and Grace@UCL) were also used in the completion of this work. D.O.S, C.J.C., and I.P.P. would like to acknowledge support from the EPSRC (EP/N01572X/1). D.O.S. acknowledges support from the European Research Council, ERC, (grant no. 758345). We are grateful to the U.K. Materials and Molecular Modelling Hub for computational resources, which is partially funded by EPSRC (EP/P020194/1).

Supporting Information Available

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.chemmater.9b00257.

  • Details of the limiting phases pertaining to the chemical potential limits, density of states (DOS) of the codoped clusters and formation energies under each growth regime together with transition levels relative to the VBM (PDF)

The authors declare no competing financial interest.

Supplementary Material

cm9b00257_si_001.pdf (81.5KB, pdf)

References

  1. Tayade R. J.; Surolia P. K.; Kulkarni R. G.; Jasra R. V. Photocatalytic Degradation of Dyes and Organic Contaminants in Water using Nanocrystalline Anatase and Rutile TiO2. Sci. Technol. Adv. Mater. 2007, 8, 455–462. 10.1016/j.stam.2007.05.006. [DOI] [Google Scholar]
  2. Orlov A.; Chan M. S.; Jefferson D. A.; Zhou D.; Lynch R. J.; Lambert R. M. Photocatalytic Degradation of Water-Soluble Organic Pollutants on TiO2 Modified with Gold Nanoparticles. Environ. Technol. 2006, 27, 747–752. 10.1080/09593332708618685. [DOI] [PubMed] [Google Scholar]
  3. Yu J. C.; Ho W.; Lin J.; Yip H.; Wong P. K. Photocatalytic Activity, Antibacterial Effect, and Photoinduced Hydrophilicity of TiO2 Films Coated on a Stainless Steel Substrate. Environ. Sci. Technol. 2003, 37, 2296–2301. 10.1021/es0259483. [DOI] [PubMed] [Google Scholar]
  4. Fu G.; Vary P. S.; Lin C.-T. Anatase TiO2 Nanocomposites for Antimicrobial Coatings. J. Phys. Chem. B 2005, 109, 8889–8898. 10.1021/jp0502196. [DOI] [PubMed] [Google Scholar]
  5. Dunnill C. W.; Aiken Z. A.; Pratten J.; Wilson M.; Morgan D. J.; Parkin I. P. 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 2009, 207, 244–253. 10.1016/j.jphotochem.2009.07.024. [DOI] [Google Scholar]
  6. Page K.; Wilson M.; Parkin I. P. Antimicrobial surfaces and their potential in reducing the role of the inanimate environment in the incidence of hospital-acquired infections. J. Mater. Chem. 2009, 19, 3819–3831. 10.1039/b818698g. [DOI] [Google Scholar]
  7. Walter M. G.; Warren E. L.; McKone J. R.; Boettcher S. W.; Mi Q.; Santori E. A.; Lewis N. S. Solar Water Splitting Cells. Chem. Rev. 2010, 110, 6446–6473. 10.1021/cr1002326. [DOI] [PubMed] [Google Scholar]
  8. Kudo A.; Miseki Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. 10.1039/B800489G. [DOI] [PubMed] [Google Scholar]
  9. Xu C.; Killmeyer R.; Gray M. L.; Khan S. U. Enhanced carbon doping of n-TiO2 thin films for photoelectrochemical water splitting. Electrochem. Commun. 2006, 8, 1650–1654. 10.1016/j.elecom.2006.07.018. [DOI] [Google Scholar]
  10. Bhachu D. S.; Sathasivam S.; Sankar G.; Scanlon D. O.; Cibin G.; Carmalt C. J.; Parkin I. P.; Watson G. W.; Bawaked S. M.; Obaid A. Y.; Al-Thabaiti S.; Basahel S. N. Solution Processing Route to Multifunctional Titania Thin Films: Highly Conductive and Photcatalytically Active Nb:TiO2. Adv. Funct. Mater. 2014, 24, 5075–5085. 10.1002/adfm.201400338. [DOI] [Google Scholar]
  11. Walsh A.; Ahn K.-S.; Shet S.; Huda M. N.; Deutsch T. G.; Wang H.; Turner J. A.; Wei S.-H.; Yan Y.; Al-Jassim M. M. Ternary cobalt spinel oxides for solar driven hydrogen production: Theory and experiment. Energy Environ. Sci. 2009, 2, 774–782. 10.1039/b822903a. [DOI] [Google Scholar]
  12. Buckeridge J.; Butler K. T.; Catlow C. R. A.; Logsdail A. J.; Scanlon D. O.; Shevlin S. A.; Woodley S. M.; Sokol A. A.; Walsh A. Polymorph Engineering of TiO2: Demonstrating How Absolute Reference Potentials Are Determined by Local Coordination. Chem. Mater. 2015, 27, 3844–3851. 10.1021/acs.chemmater.5b00230. [DOI] [Google Scholar]
  13. Walsh A.; Buckeridge J.; Catlow C. R. A.; Jackson A. J.; Keal T. W.; Miskufova M.; Sherwood P.; Shevlin S. A.; Watkins M. B.; Woodley S. M.; Sokol A. A. Limits to Doping of Wide Band Gap Semiconductors. Chem. Mater. 2013, 25, 2924–2926. 10.1021/cm402237s. [DOI] [Google Scholar]
  14. Scanlon D. O.; Watson G. W. On the possibility of p-type SnO2. J. Mater. Chem. 2012, 22, 25236–25245. 10.1039/c2jm34352e. [DOI] [Google Scholar]
  15. Williamson B. A. D.; Buckeridge J.; Brown J.; Ansbro S.; Palgrave R. G.; Scanlon D. O. Engineering Valence Band Dispersion for High Mobility p-Type Semiconductors. Chem. Mater. 2016, 29, 2402–2413. 10.1021/acs.chemmater.6b03306. [DOI] [Google Scholar]
  16. Asahi R. Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides. Science 2001, 293, 269–271. 10.1126/science.1061051. [DOI] [PubMed] [Google Scholar]
  17. Sakthivel S.; Janczarek M.; Kisch H. Visible Light Activity and Photoelectrochemical Properties of Nitrogen-Doped TiO2. J. Phys. Chem. B 2004, 108, 19384–19387. 10.1021/jp046857q. [DOI] [Google Scholar]
  18. Finazzi E.; Di Valentin C.; Pacchioni G. Boron-Doped AnataseTiO2: Pure and Hybrid DFT Calculations. J. Phys. Chem. C 2009, 113, 220–228. 10.1021/jp8072238. [DOI] [Google Scholar]
  19. Wu X.; Yin S.; Dong Q.; Guo C.; Li H.; Kimura T.; Sato T. Synthesis of high visible light active carbon doped TiO2 photocatalyst by a facile calcination assisted solvothermal method. Appl. Catal., B 2013, 142–143, 450–457. 10.1016/j.apcatb.2013.05.052. [DOI] [Google Scholar]
  20. Sotelo-Vazquez C.; Noor N.; Kafizas A.; Quesada-Cabrera R.; Scanlon D. O.; Taylor A.; Durrant J. R.; Parkin I. P. Multifunctional P-Doped TiO2 Films: A New Approach to Self-Cleaning, Transparent Conducting Oxide Materials. Chem. Mater. 2015, 27, 3234–3242. 10.1021/cm504734a. [DOI] [Google Scholar]
  21. Quesada-González M.; Boscher N. D.; Carmalt C. J.; Parkin I. P. Interstitial Boron-Doped TiO2 Thin Films: The Significant Effect of Boron on TiO2 Coatings Grown by Atmospheric Pressure Chemical Vapor Deposition. ACS Appl. Mater. Interfaces 2016, 8, 25024–25029. 10.1021/acsami.6b09560. [DOI] [PubMed] [Google Scholar]
  22. Wang B.; Zhao F.; Du G.; Porter S.; Liu Y.; Zhang P.; Cheng Z.; Liu H. K.; Huang Z. Boron-Doped Anatase TiO2 as a High-Performance Anode Material for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 16009–16015. 10.1021/acsami.6b03270. [DOI] [PubMed] [Google Scholar]
  23. Grabowska E.; Zaleska A.; Sobczak J.; Gazda M.; Hupka J. Boron-doped TiO2: Characteristics and photoactivity under visible light. Procedia Chem. 2009, 1, 1553–1559. 10.1016/j.proche.2009.11.003. [DOI] [Google Scholar]
  24. Yang Y.; Ni D.; Yao Y.; Zhong Y.; Ma Y.; Yao J. High photocatalytic activity of carbon doped TiO2 prepared by fast combustion of organic capping ligands. RSC Adv. 2015, 5, 93635–93643. 10.1039/C5RA19058D. [DOI] [Google Scholar]
  25. Wu G.; Nishikawa T.; Ohtani B.; Chen A. Synthesis and Characterization of Carbon-Doped TiO2 Nanostructures with Enhanced Visible Light Response. Chem. Mater. 2007, 19, 4530–4537. 10.1021/cm071244m. [DOI] [Google Scholar]
  26. Park Y.; Kim W.; Park H.; Tachikawa T.; Majima T.; Choi W. Carbon-doped TiO2 photocatalyst synthesized without using an external carbon precursor and the visible light activity. Appl. Catal., B 2009, 91, 355–361. 10.1016/j.apcatb.2009.06.001. [DOI] [Google Scholar]
  27. Sathish M.; Viswanathan B.; Viswanath R. P.; Gopinath C. S. Synthesis, Characterization, Electronic Structure, and Photocatalytic Activity of Nitrogen-Doped TiO2 Nanocatalyst. Chem. Mater. 2005, 17, 6349–6353. 10.1021/cm052047v. [DOI] [Google Scholar]
  28. 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, 6976–6982. 10.1021/jp0685030. [DOI] [Google Scholar]
  29. Gopal N. O.; Lo H.-H.; Ke T.-F.; Lee C.-H.; Chou C.-C.; Wu J.-D.; Sheu S.-C.; Ke S.-C. Visible Light Active Phosphorus-Doped TiO2 Nanoparticles: An EPR Evidence for the Enhanced Charge Separation. J. Phys. Chem. C 2012, 116, 16191–16197. 10.1021/jp212346f. [DOI] [Google Scholar]
  30. Shi J.; Chen S.; Wang S.; Ye Z. Sol-Gel Preparation and Visible Light Photocatalytic Activity of Nitrogen Doped Titania. Procedia Eng. 2012, 27, 564–569. 10.1016/j.proeng.2011.12.488. [DOI] [Google Scholar]
  31. Ai H.-Y.; Shi J.-W.; Duan R.-X.; Chen J.-W.; Cui H.-J.; Fu M.-L. Sol-Gel to Prepare Nitrogen Doped TiO2 Nanocrystals with Exposed {001} Facets and High Visible-Light Photocatalytic Performance. Int. J. Photoenergy 2014, 2014, 1–9. 10.1155/2014/724910. [DOI] [Google Scholar]
  32. Quesada-Cabrera R.; Sotelo-Vázquez C.; Quesada-González M.; Melián E. P.; Chadwick N.; Parkin I. P. On the apparent visible-light and enhanced UV-light photocatalytic activity of nitrogen-doped TiO2 thin films. J. Photochem. Photobiol., A 2017, 333, 49–55. 10.1016/j.jphotochem.2016.10.013. [DOI] [Google Scholar]
  33. Kim S. J.; Xu K.; Parala H.; Beranek R.; Bledowski M.; Sliozberg K.; Becker H.-W.; Rogalla D.; Barreca D.; Maccato C.; Sada C.; Schuhmann W.; Fischer R. A.; Devi A. Intrinsic Nitrogen-doped CVD-grown TiO2 Thin Films from All-N-coordinated Ti Precursors for Photoelectrochemical Applications. Chem. Vap. Deposition 2013, 19, 45–52. 10.1002/cvde.201206996. [DOI] [Google Scholar]
  34. Sarantopoulos C.; Gleizes A. N.; Maury F. Chemical vapor deposition and characterization of nitrogen doped TiO2 thin films on glass substrates. Thin Solid Films 2009, 518, 1299–1303. 10.1016/j.tsf.2009.04.070. [DOI] [Google Scholar]
  35. Chen S.-Z.; Zhang P.-Y.; Zhuang D.-M.; Zhu W.-P. Investigation of nitrogen doped TiO2 photocatalytic films prepared by reactive magnetron sputtering. Catal. Commun. 2004, 5, 677–680. 10.1016/j.catcom.2004.08.011. [DOI] [Google Scholar]
  36. Prabakar K.; Takahashi T.; Nezuka T.; Takahashi K.; Nakashima T.; Kubota Y.; Fujishima A. Visible light-active nitrogen-doped TiO2 thin films prepared by DC magnetron sputtering used as a photocatalyst. Renewable Energy 2008, 33, 277–281. 10.1016/j.renene.2007.05.018. [DOI] [Google Scholar]
  37. Pandiyan R.; Delegan N.; Dirany A.; Drogui P.; Khakani M. A. E. Probing the Electronic Surface Properties and Bandgap Narrowing of in situ N, W, and (W,N) Doped Magnetron-Sputtered TiO2 Films Intended for Electro-Photocatalytic Applications. J. Phys. Chem. C 2016, 120, 631–638. 10.1021/acs.jpcc.5b08057. [DOI] [Google Scholar]
  38. Varley J. B.; Janotti A.; de Walle C. G. V. Mechanism of Visible-Light Photocatalysis in Nitrogen-Doped TiO2. Adv. Mater. 2011, 23, 2343–2347. 10.1002/adma.201003603. [DOI] [PubMed] [Google Scholar]
  39. Chen H.; Dawson J. A. Nature of Nitrogen-Doped Anatase TiO2 and the Origin of Its Visible-Light Activity. J. Phys. Chem. C 2015, 119, 15890–15895. 10.1021/acs.jpcc.5b03587. [DOI] [Google Scholar]
  40. Finazzi E.; Di Valentin C.; Selloni A.; Pacchioni G. First Principles Study of Nitrogen Doping at the Anatase TiO2 (101) Surface. J. Phys. Chem. C 2007, 111, 9275–9282. 10.1021/jp071186s. [DOI] [Google Scholar]
  41. Livraghi S.; Paganini M. C.; Giamello E.; Selloni A.; Di Valentin C.; Pacchioni G. Origin of Photoactivity of Nitrogen-Doped Titanium Dioxide under Visible Light. J. Am. Chem. Soc. 2006, 128, 15666–15671. 10.1021/ja064164c. [DOI] [PubMed] [Google Scholar]
  42. Di Valentin C.; Finazzi E.; Pacchioni G.; Selloni A.; Livraghi S.; Paganini M. C.; Giamello E. N-doped TiO2: Theory and experiment. Chem. Phys. 2007, 339, 44–56. 10.1016/j.chemphys.2007.07.020. [DOI] [Google Scholar]
  43. Di Valentin C.; Pacchioni G.; Selloni A.; Livraghi S.; Giamello E. Characterization of Paramagnetic Species in N-Doped TiO2 Powders by EPR Spectroscopy and DFT Calculations. J. Phys. Chem. B 2005, 109, 11414–11419. 10.1021/jp051756t. [DOI] [PubMed] [Google Scholar]
  44. Gai Y.; Li J.; Li S.-S.; Xia J.-B.; Wei S.-H. Design of Narrow-Gap TiO2: A Passivated Codoping Approach for Enhanced Photoelectrochemical Activity. Phys. Rev. Lett. 2009, 102, 036402 10.1103/PhysRevLett.102.036402. [DOI] [PubMed] [Google Scholar]
  45. Bharti B.; Kumar S.; Lee H.-N.; Kumar R. Formation of oxygen vacancies and Ti3+ state in TiO2 thin film and enhanced optical properties by air plasma treatment. Sci. Rep. 2016, 6, 32355 10.1038/srep32355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Shannon R. D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1976, 32, 751–767. 10.1107/S0567739476001551. [DOI] [Google Scholar]
  47. Li Y.; Hwang D.-S.; Lee N. H.; Kim S.-J. Synthesis and characterization of carbon-doped titania as an artificial solar light sensitive photocatalyst. Chem. Phys. Lett. 2005, 404, 25–29. 10.1016/j.cplett.2005.01.062. [DOI] [Google Scholar]
  48. Di Valentin C.; Pacchioni G.; Selloni A. Theory of Carbon Doping of Titanium Dioxide. Chem. Mater. 2005, 17, 6656–6665. 10.1021/cm051921h. [DOI] [Google Scholar]
  49. Sun Y.-Y.; Zhang S. Kinetics stabilized doping: computational optimization of carbon-doped anatase TiO2 for visible-light driven water splitting. Phys. Chem. Chem. Phys. 2016, 18, 2776–2783. 10.1039/C5CP07109G. [DOI] [PubMed] [Google Scholar]
  50. Sakthivel S.; Kisch H. Daylight Photocatalysis by Carbon-Modified Titanium Dioxide. Angew. Chem., Int. Ed. 2003, 42, 4908–4911. 10.1002/anie.200351577. [DOI] [PubMed] [Google Scholar]
  51. Sharma S. D.; Singh D.; Saini K.; Kant C.; Sharma V.; Jain S.; Sharma C. Sol–gel-derived super-hydrophilic nickel doped TiO2 film as active photo-catalyst. Appl. Catal., A 2006, 314, 40–46. 10.1016/j.apcata.2006.07.029. [DOI] [Google Scholar]
  52. Janisch R.; Gopal P.; Spaldin N. A. Transition metal-doped TiO2 and ZnO—present status of the field. J. Phys.: Condens. Matter 2005, 17, R657–R689. 10.1088/0953-8984/17/27/R01. [DOI] [Google Scholar]
  53. Tian J.; Gao H.; Deng H.; Sun L.; Kong H.; Yang P.; Chu J. Structural, magnetic and optical properties of Ni-doped TiO2 thin films deposited on silicon(100) substrates by sol–gel process. J. Alloys Compd. 2013, 581, 318–323. 10.1016/j.jallcom.2013.07.105. [DOI] [Google Scholar]
  54. Al-Jawad S. M.; Taha A. A.; Salim M. M. Synthesis and characterization of pure and Fe doped TiO2 thin films for antimicrobial activity. Optik 2017, 142, 42–53. 10.1016/j.ijleo.2017.05.048. [DOI] [Google Scholar]
  55. Mathews N. R.; Jacome M. A. C.; Morales E. R.; Antonio J. A. T. Structural and spectroscopic study of the Fe doped TiO2 thin films for applications in photocatalysis. Phys. Status Solidi C 2009, 6, S219–S223. 10.1002/pssc.200881319. [DOI] [Google Scholar]
  56. Wang X.; Gong W. Bactericidal and Photocatalytic Activity of Fe3.-TiO2 Thin Films Prepared by the Sol-Gel Method. J. Wuhan Univ. Technol., Mater. Sci. Ed. 2008, 23, 155–158. 10.1007/s11595-006-2155-x. [DOI] [Google Scholar]
  57. Alexandrescu R.; Morjan I.; Scarisoreanu M.; Birjega R.; Popovici E.; Soare I.; Gavrila-Florescu L.; Voicu I.; Sandu I.; Dumitrache F.; Prodan G.; Vasile E.; Figgemeier E. Structural investigations on TiO2 and Fe-dopedTiO2 nanoparticles synthesized by laser pyrolysis. Thin Solid Films 2007, 515, 8438–8445. 10.1016/j.tsf.2007.03.106. [DOI] [Google Scholar]
  58. Sadanandam G.; Lalitha K.; Kumari V. D.; Shankar M. V.; Subrahmanyam M. Cobalt doped TiO2: A stable and efficient photocatalyst for continuous hydrogen production from glycerol: Water mixtures under solar light irradiation. Int. J. Hydrogen Energy 2013, 38, 9655–9664. 10.1016/j.ijhydene.2013.05.116. [DOI] [Google Scholar]
  59. Iwasaki M.; Hara M.; Kawada H.; Tada H.; Ito S. Cobalt Ion-Doped TiO2 Photocatalyst Response to Visible Light. J. Colloid Interface Sci. 2000, 224, 202–204. 10.1006/jcis.1999.6694. [DOI] [PubMed] [Google Scholar]
  60. Miao Y.; Zhai Z.; Jiang L.; Shi Y.; Yan Z.; Duan D.; Zhen K.; Wang J. Facile and new synthesis of cobalt doped mesoporous TiO2 with high visible-light performance. Powder Technol. 2014, 266, 365–371. 10.1016/j.powtec.2014.06.019. [DOI] [Google Scholar]
  61. Ould-Chikh S.; Proux O.; Afanasiev P.; Khrouz L.; Hedhili M. N.; Anjum D. H.; Harb M.; Geantet C.; Basset J.-M.; Puzenat E. Photocatalysis with Chromium-Doped TiO2: Bulk and Surface Doping. ChemSusChem 2014, 7, 1361–1371. 10.1002/cssc.201300922. [DOI] [PubMed] [Google Scholar]
  62. Peng Y.-H.; Huang G.-F.; Huang W.-Q. Visible-light absorption and photocatalytic activity of Cr-doped TiO2 nanocrystal films. Adv. Powder Technol. 2012, 23, 8–12. 10.1016/j.apt.2010.11.006. [DOI] [Google Scholar]
  63. Dholam R.; Patel N.; Adami M.; Miotello A. Hydrogen production by photocatalytic water-splitting using Cr- or Fe-doped TiO2 composite thin films photocatalyst. Int. J. Hydrogen Energy 2009, 34, 5337–5346. 10.1016/j.ijhydene.2009.05.011. [DOI] [Google Scholar]
  64. Jun T. H.; Lee K. S. Cr-doped TiO2 thin films deposited by RF-sputtering. Mater. Lett. 2010, 64, 2287–2289. 10.1016/j.matlet.2010.07.069. [DOI] [Google Scholar]
  65. You M.; Kim T. G.; Sung Y.-M. Synthesis of Cu-Doped TiO2 Nanorods with Various Aspect Ratios and Dopant Concentrations. Cryst. Growth Des. 2010, 10, 983–987. 10.1021/cg9012944. [DOI] [Google Scholar]
  66. Colón G.; Maicu M.; Hidalgo M.; Navío J. Cu-doped TiO2 systems with improved photocatalytic activity. Appl. Catal., B 2006, 67, 41–51. 10.1016/j.apcatb.2006.03.019. [DOI] [Google Scholar]
  67. Choudhury B.; Dey M.; Choudhury A. Defect generation, d-d transition, and band gap reduction in Cu-doped TiO2 nanoparticles. Int. Nano Lett. 2013, 3, 25 10.1186/2228-5326-3-25. [DOI] [Google Scholar]
  68. Park H. S.; Kim D. H.; Kim S. J.; Lee K. S. The photocatalytic activity of 2.5 wt % Cu-doped TiO2 nano powders synthesized by mechanical alloying. J. Alloys Compd. 2006, 415, 51–55. 10.1016/j.jallcom.2005.07.055. [DOI] [Google Scholar]
  69. Choi W.; Termin A.; Hoffmann M. R. The Role of Metal Ion Dopants in Quantum-Sized TiO2: Correlation between Photoreactivity and Charge Carrier Recombination Dynamics. J. Phys. Chem. 1994, 98, 13669–13679. 10.1021/j100102a038. [DOI] [Google Scholar]
  70. Fagan R.; McCormack D. E.; Dionysiou D. D.; Pillai S. C. A review of solar and visible light active TiO2 photocatalysis for treating bacteria, cyanotoxins and contaminants of emerging concern. Mater. Sci. Semicond. Process. 2016, 42, 2–14. 10.1016/j.mssp.2015.07.052. [DOI] [Google Scholar]
  71. Zaleska A. Doped-TiO2: A Review. Recent Pat. Eng. 2008, 2, 157–164. 10.2174/187221208786306289. [DOI] [Google Scholar]
  72. Schneider J.; Matsuoka M.; Takeuchi M.; Zhang J.; Horiuchi Y.; Anpo M.; Bahnemann D. W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. 10.1021/cr5001892. [DOI] [PubMed] [Google Scholar]
  73. Linsebigler A. L.; Lu G.; Yates J. T. Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 1995, 95, 735–758. 10.1021/cr00035a013. [DOI] [Google Scholar]
  74. Kafizas A.; Dunnill C. W.; Parkin I. P. Combinatorial atmospheric pressure chemical vapour deposition (cAPCVD) of niobium doped anatase; effect of niobium on the conductivity and photocatalytic activity. J. Mater. Chem. 2010, 20, 8336–8349. 10.1039/c0jm01244k. [DOI] [Google Scholar]
  75. Alim M. A.; Bak T.; Atanacio A. J.; Ionescu M.; Kennedy B.; Price W. S.; Plessis J. D.; Pourmahdavi M.; Zhou M.; Torres A.; Nowotny J. Photocatalytic properties of Ta-doped TiO2. Ionics 2017, 23, 3517–3531. 10.1007/s11581-017-2162-2. [DOI] [Google Scholar]
  76. Znad H.; Ang M. H.; Tade M. O. Ta/Ti-and Nb/Ti-Mixed Oxides as Efficient Solar Photocatalysts: Preparation, Characterization, and Photocatalytic Activity. Int. J. Photoenergy 2012, 2012, 1–9. 10.1155/2012/548158. [DOI] [Google Scholar]
  77. Kelly P.; West G.; Ratova M.; Fisher L.; Ostovarpour S.; Verran J. Structural Formation and Photocatalytic Activity of Magnetron Sputtered Titania and Doped-Titania Coatings. Molecules 2014, 19, 16327–16348. 10.3390/molecules191016327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Moon J.; Takagi H.; Fujishiro Y.; Awano M. Preparation and characterization of the Sb-doped TiO2 photocatalysts. J. Mater. Sci. 2001, 36, 949–955. 10.1023/A:1004819706292. [DOI] [Google Scholar]
  79. Luo L.; Li T.; Ran X.; Wang P.; Guo L. Probing Photocatalytic Characteristics of Sb-DopedTiO2under Visible Light Irradiation. J. Nanomater. 2014, 2014, 1–6. 10.1155/2014/947289. [DOI] [Google Scholar]
  80. Sathasivam S.; Bhachu D. S.; Lu Y.; Chadwick N.; Althabaiti S. A.; Alyoubi A. O.; Basahel S. N.; Carmalt C. J.; Parkin I. P. Tungsten Doped TiO2 with Enhanced Photocatalytic and Optoelectrical Properties via Aerosol Assisted Chemical Vapor Deposition. Sci. Rep. 2015, 5, 10952 10.1038/srep10952. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Putta T.; Lu M.-C.; Anotai J. Photocatalytic activity of tungsten-doped TiO2 with hydrothermal treatment under blue light irradiation. J. Environ. Manage. 2011, 92, 2272–2276. 10.1016/j.jenvman.2011.04.016. [DOI] [PubMed] [Google Scholar]
  82. Sheppard L. R.; Hager S.; Holik J.; Liu R.; Macartney S.; Wuhrer R. Tantalum Segregation in Ta-Doped TiO2 and the Related Impact on Charge Separation during Illumination. J. Phys. Chem. C 2014, 119, 392–400. 10.1021/jp509806h. [DOI] [Google Scholar]
  83. Bhachu D. S.; Egdell R. G.; Sankar G.; Carmalt C. J.; Parkin I. P. Electronic properties of antimony-doped anatase TiO2 thin films prepared by aerosol assisted chemical vapour deposition. J. Mater. Chem. C 2017, 5, 9694–9701. 10.1039/C6TC04462J. [DOI] [Google Scholar]
  84. Yin W.-J.; Tang H.; Wei S.-H.; Al-Jassim M. M.; Turner J.; Yan Y. Band structure engineering of semiconductors for enhanced photoelectrochemical water splitting: The case TiO2. Phys. Rev. B 2010, 82, 045106 10.1103/PhysRevB.82.045106. [DOI] [Google Scholar]
  85. Zhang P.; Yin S.; Sekino T.; Lee S. W.; Sato T. Nb and N codoped TiO2 for a high-performance deNOx photocatalyst under visible LED light irradiation. Res. Chem. Intermed. 2012, 39, 1509–1515. 10.1007/s11164-012-0615-9. [DOI] [Google Scholar]
  86. Chadwick N. P.; Glover E. N. K.; Sathasivam S.; Basahel S. N.; Althabaiti S. A.; Alyoubi A. O.; Parkin I. P.; Carmalt C. J. Photo-activity and low resistivity in N/Nb Codoped TiO2 thin films by combinatorial AACVD. J. Mater. Chem. A 2016, 4, 407–415. 10.1039/C5TA07922E. [DOI] [Google Scholar]
  87. Lim J.; Murugan P.; Lakshminarasimhan N.; Kim J. Y.; Lee J. S.; Lee S.-H.; Choi W. Synergic photocatalytic effects of nitrogen and niobium codoping in TiO2 for the redox conversion of aquatic pollutants under visible light. J. Catal. 2014, 310, 91–99. 10.1016/j.jcat.2013.05.014. [DOI] [Google Scholar]
  88. Breault T. M.; Bartlett B. M. Lowering the Band Gap of Anatase-Structured TiO2 by Coalloying with Nb and N: Electronic Structure and Photocatalytic Degradation of Methylene Blue Dye. J. Phys. Chem. C 2012, 116, 5986–5994. 10.1021/jp2078456. [DOI] [Google Scholar]
  89. Breault T. M.; Bartlett B. M. Composition Dependence of TiO2:(Nb,N)-Compounds on the Rate of Photocatalytic Methylene Blue Dye Degradation. J. Phys. Chem. C 2013, 117, 8611–8618. 10.1021/jp312199t. [DOI] [Google Scholar]
  90. Zhao Y. F.; Li C.; Hu J. Y.; Gong Y. Y.; Niu L. Y.; Liu X. J. Ta and N modulated electronic, optical and photocatalytic properties of TiO2. Phys. Lett. A 2016, 380, 910–916. 10.1016/j.physleta.2015.12.027. [DOI] [Google Scholar]
  91. Obata K.; Irie H.; Hashimoto K. Enhanced photocatalytic activities of Ta, N codoped TiO2 thin films under visible light. Chem. Phys. 2007, 339, 124–132. 10.1016/j.chemphys.2007.07.044. [DOI] [Google Scholar]
  92. McDonnell K. A.; English N. J.; Rahman M.; Dowling D. P. Influence of doping on the photoactive properties of magnetron-sputtered titania coatings: Experimental and theoretical study. Phys. Rev. B 2012, 86, 115306 10.1103/PhysRevB.86.115306. [DOI] [Google Scholar]
  93. Dong P.; Liu B.; Wang Y.; Pei H.; Yin S. Enhanced photocatalytic activity of (Mo, C)-codoped anatase TiO2 nanoparticles for degradation of methyl orange under simulated solar irradiation. J. Mater. Res. 2010, 25, 2392–2400. 10.1557/jmr.2010.0307. [DOI] [Google Scholar]
  94. Yang F.; Yang H.; Tian B.; Zhang J.; He D. Preparation of an Mo and C codoped TiO2 catalyst by a calcination–hydrothermal method, and degradation of rhodamine B in visible light. Res. Chem. Intermed. 2012, 39, 1685–1699. 10.1007/s11164-012-0902-5. [DOI] [Google Scholar]
  95. Zhang J.; Pan C.; Fang P.; Wei J.; Xiong R. Codoped TiO2 Using Thermal Oxidation for Enhancing Photocatalytic Activity. ACS Appl. Mater. Interfaces 2010, 2, 1173–1176. 10.1021/am100011c. [DOI] [PubMed] [Google Scholar]
  96. Zhe-Peng Z.; Biao Y.; Hai-Bo F.; Xin-Liang Z.; He-Bao Y. Visible-light photocatalytic properties of Mo–C codoped anatase TiO2 films prepared by magnetron sputtering. J. Phys. Chem. Solids 2015, 87, 53–57. 10.1016/j.jpcs.2015.08.001. [DOI] [Google Scholar]
  97. Li Z.; Wang X.; Xing X.; Wang Y. Tailoring the electronic structure of anatase TiO2 (001) surface through W and N codoping: a DFT calculation. J. Korean Phys. Soc. 2017, 70, 286–291. 10.3938/jkps.70.286. [DOI] [Google Scholar]
  98. Ren D.; Cheng J.; Cheng X. Ab initio studies of Nb–N–S tri-doped TiO2 with enhanced visible light photocatalytic activity. J. Solid State Chem. 2016, 238, 83–87. 10.1016/j.jssc.2016.03.020. [DOI] [Google Scholar]
  99. Choi H.; Shin D.; Yeo B. C.; Song T.; Han S. S.; Park N.; Kim S. Simultaneously Controllable Doping Sites and the Activity of a W–N Codoped TiO2 Photocatalyst. ACS Catal. 2016, 6, 2745–2753. 10.1021/acscatal.6b00104. [DOI] [Google Scholar]
  100. Rimoldi L.; Ambrosi C.; Liberto G. D.; Presti L. L.; Ceotto M.; Oliva C.; Meroni D.; Cappelli S.; Cappelletti G.; Soliveri G.; Ardizzone S. Impregnation versus Bulk Synthesis: How the Synthetic Route Affects the Photocatalytic Efficiency of Nb/Ta:N Codoped TiO2 Nanomaterials. J. Phys. Chem. C 2015, 119, 24104–24115. 10.1021/acs.jpcc.5b06827. [DOI] [Google Scholar]
  101. Gong J.; Yang C.; Zhang J.; Pu W. Origin of Photocatalytic Activity of W/N-Codoped TiO2 H2 Production and DFT Calculation with GGA+U. Appl. Catal., B 2014, 152–153, 73–81. 10.1016/j.apcatb.2014.01.028. [DOI] [Google Scholar]
  102. Patel N.; Jaiswal R.; Warang T.; Scarduelli G.; Dashora A.; Ahuja B.; Kothari D.; Miotello A. Efficient photocatalytic degradation of organic water pollutants using V–N-codoped TiO2 thin films. Appl. Catal., B 2014, 150–151, 74–81. 10.1016/j.apcatb.2013.11.033. [DOI] [Google Scholar]
  103. Fang Y.; Cheng D.; Niu M.; Yi Y.; Wu W. Tailoring the Electronic and Optical Properties of Rutile TiO2 by (Nb+Sb, C) Codoping from DFT+U Calculations. Chem. Phys. Lett. 2013, 567, 34–38. 10.1016/j.cplett.2013.02.070. [DOI] [Google Scholar]
  104. Huang J.; Wen S.; Liu J.; He G. Band gap narrowing of TiO2 by compensated codoping for enhanced photocatalytic activity. J. Nat. Gas Chem. 2012, 21, 302–307. 10.1016/S1003-9953(11)60368-X. [DOI] [Google Scholar]
  105. Kurtoglu M. E.; Longenbach T.; Sohlberg K.; Gogotsi Y. Strong Coupling of Cr and N in Cr–N-doped TiO2 and Its Effect on Photocatalytic Activity. J. Phys. Chem. C 2011, 115, 17392–17399. 10.1021/jp2026972. [DOI] [Google Scholar]
  106. Liu L.; Chen S.; Sun W.; Xin J. Enhancing the visible light absorption via combinational doping of TiO2 with nitrogen (N) and chromium (Cr). J. Mol. Struct. 2011, 1001, 23–28. 10.1016/j.molstruc.2011.06.009. [DOI] [Google Scholar]
  107. Di Valentin C.; Finazzi E.; Pacchioni G.; Selloni A.; Livraghi S.; Czoska A. M.; Paganini M. C.; Giamello E. Density Functional Theory and Electron Paramagnetic Resonance Study on the Effect of N-F Codoping of TiO2. Chem. Mater. 2008, 20, 3706–3714. 10.1021/cm703636s. [DOI] [Google Scholar]
  108. Long R.; English N. J. Band gap engineering of (N,Ta)-codoped TiO2: A first-principles calculation. Chem. Phys. Lett. 2009, 478, 175–179. 10.1016/j.cplett.2009.07.084. [DOI] [Google Scholar]
  109. Long R.; English N. J. Synergistic Effects on Band Gap-Narrowing in Titania by Codoping from First-Principles Calculations. Chem. Mater. 2010, 22, 1616–1623. 10.1021/cm903688z. [DOI] [Google Scholar]
  110. Kresse G.; Hafner J. Ab initiomolecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561. 10.1103/PhysRevB.47.558. [DOI] [PubMed] [Google Scholar]
  111. Kresse G.; Furthmüller J. Efficient iterative schemes forab initiototal-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. 10.1103/PhysRevB.54.11169. [DOI] [PubMed] [Google Scholar]
  112. Kresse G.; Furthmüller J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. 10.1016/0927-0256(96)00008-0. [DOI] [PubMed] [Google Scholar]
  113. Kresse G.; Hafner J. Ab initiomolecular-dynamics simulation of the liquid-metal–amorphous-semiconductor transition in germanium. Phys. Rev. B 1994, 49, 14251–14269. 10.1103/PhysRevB.49.14251. [DOI] [PubMed] [Google Scholar]
  114. Krukau A. V.; Vydrov O. A.; Izmaylov A. F.; Scuseria G. E. Influence of the exchange screening parameter on the performance of screened hybrid functionals. J. Chem. Phys. 2006, 125, 224106 10.1063/1.2404663. [DOI] [PubMed] [Google Scholar]
  115. Heyd J.; Scuseria G. E.; Ernzerhof M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. 10.1063/1.1564060. [DOI] [Google Scholar]
  116. Paier J.; Marsman M.; Hummer K.; Kresse G.; Gerber I. C.; Ángyán J. G. Screened hybrid density functionals applied to solids. J. Chem. Phys. 2006, 124, 154709 10.1063/1.2187006. [DOI] [PubMed] [Google Scholar]
  117. Regoutz A.; Mascheck M.; Wiell T.; Eriksson S. K.; Liljenberg C.; Tetzner K.; Williamson B. A. D.; Scanlon D. O.; Palmgren P. A novel laboratory-based hard X-ray photoelectron spectroscopy system. Rev. Sci. Instrum. 2018, 89, 073105 10.1063/1.5039829. [DOI] [PubMed] [Google Scholar]
  118. Boonchun A.; Reunchan P.; Umezawa N. Energetics of native defects in anatase TiO2: a hybrid density functional study. Phys. Chem. Chem. Phys. 2016, 18, 30040–30046. 10.1039/C6CP05798E. [DOI] [PubMed] [Google Scholar]
  119. Matsubara M.; Saniz R.; Partoens B.; Lamoen D. Doping anatase TiO2 with group V-b and VI-b transition metal atoms: a hybrid functional first-principles study. Phys. Chem. Chem. Phys. 2017, 19, 1945–1952. 10.1039/C6CP06882K. [DOI] [PubMed] [Google Scholar]
  120. Huy H. A.; Aradi B.; Frauenheim T.; Deák P. Calculation of carrier-concentration-dependent effective mass in Nb-doped anatase crystals of TiO2. Phys. Rev. B 2011, 83, 155201 10.1103/PhysRevB.83.155201. [DOI] [Google Scholar]
  121. Çelik V.; Mete E. Range-separated hybrid exchange-correlation functional analyses of anatase TiO2 doped with W, N, S, W/N, or W/S. Phys. Rev. B 2012, 86, 205112 10.1103/PhysRevB.86.205112. [DOI] [Google Scholar]
  122. Alotaibi A. M.; Sathasivam S.; Williamson B. A. D.; Kafizas A.; Sotelo-Vazquez C.; Taylor A.; Scanlon D. O.; Parkin I. P. Chemical Vapor Deposition of Photocatalytically Active Pure Brookite TiO2 Thin Films. Chem. Mater. 2018, 30, 1353–1361. 10.1021/acs.chemmater.7b04944. [DOI] [Google Scholar]
  123. Quesada-Gonzalez M.; Williamson B. A. D.; Sotelo-Vazquez C.; Kafizas A.; Boscher N. D.; Quesada-Cabrera R.; Scanlon D. O.; Carmalt C. J.; Parkin I. P. Deeper Understanding of Interstitial Boron-Doped Anatase Thin Films as A Multifunctional Layer Through Theory and Experiment. J. Phys. Chem. C 2017, 122, 714–726. 10.1021/acs.jpcc.7b11142. [DOI] [Google Scholar]
  124. Mori-Sánchez P.; Cohen A. J.; Yang W. Localization and Delocalization Errors in Density Functional Theory and Implications for Band-Gap Prediction. Phys. Rev. Lett. 2008, 100, 146401 10.1103/PhysRevLett.100.146401. [DOI] [PubMed] [Google Scholar]
  125. Blöchl P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. 10.1103/PhysRevB.50.17953. [DOI] [PubMed] [Google Scholar]
  126. Allison T JANAF Thermochemical Tables. NIST Standard Reference Database 13 1996, 10.18434/T42S31. [DOI] [Google Scholar]
  127. Landmann M.; Rauls E.; Schmidt W. G. The electronic structure and optical response of rutile, anatase and brookite TiO2. J. Phys.: Condens. Matter 2012, 24, 195503 10.1088/0953-8984/24/19/195503. [DOI] [PubMed] [Google Scholar]
  128. Dou M.; Persson C. Comparative study of rutile and anatase SnO2 and TiO2: Band-edge structures, dielectric functions, and polaron effects. J. Appl. Phys. 2013, 113, 083703 10.1063/1.4793273. [DOI] [Google Scholar]
  129. Patrick C. E.; Giustino F. GW Quasiparticle Bandgaps of Anatase TiO2 Starting from DFT+U. J. Phys.: Condens. Matter 2012, 24, 202201 10.1088/0953-8984/24/20/202201. [DOI] [PubMed] [Google Scholar]
  130. Kang W.; Hybertsen M. S. Quasiparticle and optical properties of rutile and anatase TiO2. Phys. Rev. B 2010, 82, 085203 10.1103/PhysRevB.82.085203. [DOI] [Google Scholar]
  131. Gong S.; Liu B.-G. Electronic structures and optical properties of TiO2: Improved density-functional-theory investigation. Chin. Phys. B 2012, 21, 057104 10.1088/1674-1056/21/5/057104. [DOI] [Google Scholar]
  132. Chiodo L.; Salazar M.; Romero A. H.; Laricchia S.; Sala F. D.; Rubio A. Structure, electronic, and optical properties of TiO2 atomic clusters: An ab initio study. J. Chem. Phys. 2011, 135, 244704 10.1063/1.3668085. [DOI] [PubMed] [Google Scholar]
  133. Nieminen R. M. Issues in first-principles calculations for defects in semiconductors and oxides. Modell. Simul. Mater. Sci. Eng. 2009, 17, 084001 10.1088/0965-0393/17/8/084001. [DOI] [Google Scholar]
  134. Hine N. D. M.; Frensch K.; Foulkes W. M. C.; Finnis M. W. Supercell size scaling of density functional theory formation energies of charged defects. Phys. Rev. B 2009, 79, 024112 10.1103/PhysRevB.79.024112. [DOI] [Google Scholar]
  135. Murphy S. T.; Hine N. D. M. Anisotropic charge screening and supercell size convergence of defect formation energies. Phys. Rev. B 2013, 87, 094111 10.1103/PhysRevB.87.094111. [DOI] [Google Scholar]
  136. Lany S.; Zunger A. Assessment of correction methods for the band-gap problem and for finite-size effects in supercell defect calculations: Case studies for ZnO and GaAs. Phys. Rev. B 2008, 78, 235104 10.1103/PhysRevB.78.235104. [DOI] [Google Scholar]
  137. Persson C.; Zhao Y.-J.; Lany S.; Zunger A. n-type doping ofCuInSe2andCuGaSe2. Phys. Rev. B 2005, 72, 054112 10.1103/PhysRevB.72.035211. [DOI] [Google Scholar]
  138. Ágoston P.; Körber C.; Klein A.; Puska M. J.; Nieminen R. M.; Albe K. Limits for n-type doping in In2O3 and SnO2: A theoretical approach by first-principles calculations using hybrid-functional methodology. J. Appl. Phys. 2010, 108, 053511 10.1063/1.3467780. [DOI] [Google Scholar]
  139. Stull D. R.; Prophet H.. Ultrastructure Processing of Advanced Materials; NSRDS-NBS Publication No. 37; National Bureau of Standards: U.S. Washington, DC, 1971. [Google Scholar]
  140. Alkauskas A.; McCluskey M. D.; de Walle C. G. V. Tutorial: Defects in semiconductors-Combining experiment and theory. J. Appl. Phys. 2016, 119, 181101 10.1063/1.4948245. [DOI] [Google Scholar]
  141. Taylor F. H.; Buckeridge J.; Catlow C. R. A. Defects and Oxide Ion Migration in the Solid Oxide Fuel Cell Cathode Material LaFeO3. Chem. Mater. 2016, 28, 8210–8220. 10.1021/acs.chemmater.6b03048. [DOI] [Google Scholar]
  142. Buckeridge J.; Jevdokimovs D.; Catlow C. R. A.; Sokol A. A. Nonstoichiometry and Weyl fermionic behavior in TaAs. Phys. Rev. B 2016, 94, 180101 10.1103/PhysRevB.94.180101. [DOI] [Google Scholar]
  143. Janotti A.; Varley J. B.; Rinke P.; Umezawa N.; Kresse G.; de Walle C. G. V. Hybrid functional studies of the oxygen vacancy in TiO2. Phys. Rev. B 2010, 81, 085212 10.1103/PhysRevB.81.085212. [DOI] [Google Scholar]
  144. Morgan B. J.; Watson G. W. Intrinsic n-type Defect Formation in TiO2: A Comparison of Rutile and Anatase from GGA+U calculations. J. Phys. Chem. C 2010, 114, 2321–2328. 10.1021/jp9088047. [DOI] [Google Scholar]
  145. Mattioli G.; Alippi P.; Filippone F.; Caminiti R.; Bonapasta A. A. Deep versus Shallow Behavior of Intrinsic Defects in Rutile and Anatase TiO2 Polymorphs. J. Phys. Chem. C 2010, 114, 21694–21704. 10.1021/jp1041316. [DOI] [Google Scholar]
  146. Deák P.; Aradi B.; Frauenheim T. Polaronic effects in TiO2 calculated by the HSE06 hybrid functional: Dopant passivation by carrier self-trapping. Phys. Rev. B 2011, 83, 155207 10.1103/PhysRevB.83.155207. [DOI] [Google Scholar]
  147. Morgan B. J.; Watson G. W. Polaronic trapping of electrons and holes by native defects in anatase TiO2. Phys. Rev. B 2009, 80, 233102 10.1103/PhysRevB.80.233102. [DOI] [Google Scholar]
  148. Oba F.; Nishitani S. R.; Isotani S.; Adachi H.; Tanaka I. Energetics of native defects in ZnO. J. Appl. Phys. 2001, 90, 824–828. 10.1063/1.1380994. [DOI] [Google Scholar]
  149. Oba F.; Togo A.; Tanaka I.; Paier J.; Kresse G. Defect energetics in ZnO: A hybrid Hartree-Fock density functional study. Phys. Rev. B 2008, 77, 245202 10.1103/PhysRevB.77.245202. [DOI] [Google Scholar]
  150. Janotti A.; de Walle C. G. V. Native point defects in ZnO. Phys. Rev. B 2007, 76, 165202 10.1103/PhysRevB.76.165202. [DOI] [Google Scholar]
  151. Ágoston P.; Albe K.; Nieminen R. M.; Puska M. J. Intrinsicn-Type Behavior in Transparent Conducting Oxides: A Comparative Hybrid-Functional Study of In2O3, SnO2 and ZnO. Phys. Rev. Lett. 2009, 103, 245501 10.1103/PhysRevLett.103.245501. [DOI] [PubMed] [Google Scholar]
  152. Kılıç Ç.; Zunger A. Origins of Coexistence of Conductivity and Transparency in SnO2. Phys. Rev. Lett. 2002, 88, 095501 10.1103/PhysRevLett.88.095501. [DOI] [PubMed] [Google Scholar]
  153. Singh A. K.; Janotti A.; Scheffler M.; de Walle C. G. V. Sources of Electrical Conductivity in SnO2. Phys. Rev. Lett. 2008, 101, 055502 10.1103/PhysRevLett.101.055502. [DOI] [PubMed] [Google Scholar]
  154. Scanlon D. O. Defect engineering of BaSnO3 for high-performance transparent conducting oxide applications. Phys. Rev. B 2013, 87, 161201 10.1103/PhysRevB.87.161201. [DOI] [Google Scholar]
  155. Sallis S.; Scanlon D. O.; Chae S. C.; Quackenbush N. F.; Fischer D. A.; Woicik J. C.; Guo J.-H.; Cheong S. W.; Piper L. F. J. La-doped BaSnO3—Degenerate perovskite transparent conducting oxide: Evidence from synchrotron X-ray spectroscopy. Appl. Phys. Lett. 2013, 103, 042105 10.1063/1.4816511. [DOI] [Google Scholar]
  156. Wang W.; Lu C.; Ni Y.; Su M.; Huang W.; Xu Z. Preparation and characterization of visible-light-driven N-F-Ta tri-doped TiO2 photocatalysts. Appl. Surf. Sci. 2012, 258, 8696–8703. 10.1016/j.apsusc.2012.05.077. [DOI] [Google Scholar]
  157. Li H.; Yin S.; Wang Y.; Sato T. Efficient persistent photocatalytic decomposition of nitrogen monoxide over a fluorescence-assisted CaAl2O4:(Eu, Nd)/(Ta, N)-codoped TiO2/Fe2O3. Appl. Catal., B 2013, 132-133, 487–492. 10.1016/j.apcatb.2012.12.026. [DOI] [Google Scholar]
  158. Furubayashi Y.; Hitosugi T.; Yamamoto Y.; Inaba K.; Kinoda G.; Hirose Y.; Shimada T.; Hasegawa T. A transparent metal: Nb-doped anatase TiO2. Appl. Phys. Lett. 2005, 86, 252101 10.1063/1.1949728. [DOI] [Google Scholar]
  159. Konstantinova E. A.; Kokorin A. I.; Lips K.; Sakthivel S.; Kisch H. EPR study of the illumination effect on properties of paramagnetic centers in nitrogen-doped TiO2 active in visible light photocatalysis. Appl. Magn. Res. 2009, 35, 421–427. 10.1007/s00723-009-0173-5. [DOI] [Google Scholar]
  160. Morgan B. J.; Scanlon D. O.; Watson G. W. Small polarons in Nb- and Ta-doped rutile and anatase TiO2. J. Mater. Chem. 2009, 19, 5175–5178. 10.1039/b905028k. [DOI] [Google Scholar]
  161. Osorio-Guillén J.; Lany S.; Zunger A. Atomic Control of Conductivity Versus Ferromagnetism in Wide-Gap Oxides Via Selective Doping: V, Nb, Ta in Anatase TiO2. Phys. Rev. Lett. 2008, 100, 036601 10.1103/PhysRevLett.100.036601. [DOI] [PubMed] [Google Scholar]
  162. Walsh A.; Kehoe A. B.; Temple D. J.; Watson G. W.; Scanlon D. O. PbO2: from semi-metal to transparent conducting oxide by defect chemistry control. Chem. Commun. 2013, 49, 448–450. 10.1039/C2CC35928F. [DOI] [PubMed] [Google Scholar]
  163. Catlow C. R. A.; Sokol A. A.; Walsh A. Microscopic origins of electron and hole stability in ZnO. Chem. Commun. 2011, 47, 3386–3388. 10.1039/c1cc10314h. [DOI] [PubMed] [Google Scholar]
  164. Scanlon D. O.; Dunnill C. W.; Buckeridge J.; Shevlin S. A.; Logsdail A. J.; Woodley S. M.; Catlow C. R. A.; Powell M. J.; Palgrave R. G.; Parkin I. P.; Watson G. W.; Keal T. W.; Sherwood P.; Walsh A.; Sokol A. A. Band alignment of rutile and anatase TiO2. Nat. Mater. 2013, 12, 798–801. 10.1038/nmat3697. [DOI] [PubMed] [Google Scholar]
  165. Swank R. K. Surface Properties of II-VI Compounds. Phys. Rev. 1967, 153, 844–849. 10.1103/PhysRev.153.844. [DOI] [Google Scholar]
  166. Butler M. A. Prediction of Flatband Potentials at Semiconductor-Electrolyte Interfaces from Atomic Electronegativities. J. Electrochem. Soc. 1978, 125, 228–232. 10.1149/1.2131419. [DOI] [Google Scholar]
  167. Gassenbauer Y.; Klein A. Electronic and Chemical Properties of Tin-Doped Indium Oxide (ITO) Surfaces and ITO/ZnPc Interfaces Studied In-situ by Photoelectron Spectroscopy. J. Phys. Chem. B 2006, 110, 4793–4801. 10.1021/jp056640b. [DOI] [PubMed] [Google Scholar]
  168. Kramm B.; Laufer A.; Reppin D.; Kronenberger A.; Hering P.; Polity A.; Meyer B. K. The band alignment of Cu2O/ZnO and Cu2O/GaN heterostructures. Appl. Phys. Lett. 2012, 100, 094102 10.1063/1.3685719. [DOI] [Google Scholar]
  169. Miao M.-S.; Yarbro S.; Barton P. T.; Seshadri R. Electron affinities and ionization energies of Cu and Ag delafossite compounds: A hybrid functional study. Phys. Rev. B 2014, 89, 045306 10.1103/PhysRevB.89.045306. [DOI] [Google Scholar]
  170. Quackenbush N. F.; Allen J. P.; Scanlon D. O.; Sallis S.; Hewlett J. A.; Nandur A. S.; Chen B.; Smith K. E.; Weiland C.; Fischer D. A.; Woicik J. C.; White B. E.; Watson G. W.; Piper L. F. J. Origin of the Bipolar Doping Behavior of SnO from X-ray Spectroscopy and Density Functional Theory. Chem. Mater. 2013, 25, 3114–3123. 10.1021/cm401343a. [DOI] [Google Scholar]
  171. Scanlon D. O.; Watson G. W. Uncovering the Complex Behavior of Hydrogen in Cu2O. Phys. Rev. Lett. 2011, 106, 186403 10.1103/PhysRevLett.106.186403. [DOI] [PubMed] [Google Scholar]
  172. Scanlon D. O.; Morgan B. J.; Watson G. W. Modeling the polaronic nature of p-type defects in Cu2O: The failure of GGA and GGA + U. J. Chem. Phys. 2009, 131, 124703 10.1063/1.3231869. [DOI] [PubMed] [Google Scholar]
  173. Scanlon D. O.; Watson G. W. Undoped n-Type Cu2O: Fact or Fiction?. J. Phys. Chem. Lett. 2010, 1, 2582–2585. 10.1021/jz100962n. [DOI] [Google Scholar]
  174. Scanlon D. O.; Watson G. W. Conductivity Limits in CuAlO2 from Screened-Hybrid Density Functional Theory. J. Phys. Chem. Lett. 2010, 1, 3195–3199. 10.1021/jz1011725. [DOI] [Google Scholar]
  175. Allen J. P.; Scanlon D. O.; Piper L. F. J.; Watson G. W. Understanding the defect chemistry of tin monoxide. J. Mater. Chem. C 2013, 1, 8194–8208. 10.1039/c3tc31863j. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

cm9b00257_si_001.pdf (81.5KB, pdf)

Articles from Chemistry of Materials are provided here courtesy of American Chemical Society

RESOURCES