Abstract
Controlling the formation of electron polarons in TiO2 doped with transition metals is important for the design of transparent conducting oxides for high-efficiency photovoltaics and photocatalysts with tunable reaction selectivities. In this work, EPR spectroscopy is combined with Hubbard-corrected density functional theory (DFT+U), with refined atomic-like Hubbard projectors, to show the sensitivity of charge compensation in substitutionally doped Nb-TiO2 and W-TiO2 with respect to the TiO2 polymorph (i.e., anatase or rutile). Both EPR magnetic tensors and DFT+U-predicted Nb 4d and W 5d orbital occupancies show the formation of differing dopant charge states depending on the TiO2 polymorph, with nonmagnetic Nb5+ and W6+ in doped anatase and paramagnetic Nb4+ and W5+ in doped rutile. The results provide an example of how a coherent experimental and theory-validated framework can be used to understand and predict the reducibility of dopants and electron trapping energetics in TiO2 polymorphs. The outcome enables greater control over the electronic and magnetic properties of metal oxide semiconductors, which are crucial for the rational design of next-generation materials for energy conversion and catalytic applications.


1. Introduction
Transparent conducting oxides (TCOs) underpin modern consumer electronics, photovoltaics, and light-emitting devices (LED and OLED). For example, anatase Nb-doped TiO2 (NTO) has emerged − as a more sustainable alternative to the widely used indium tin oxide, − capable of a resistivity of 2 × 10–4 to 3 × 10–4 Ωcm and 97% internal transmittance under visible light at room temperature for a 40 nm-thick film of anatase NTO with 3%at Nb. In contrast, rutile NTO is more resistive, which allows for tailored applications based on the choice of TiO2 polymorph. Similarly, anatase W-doped TiO2 (WTO) exhibits an n-type metallic behavior , with a resistivity of 1.5 × 10–2 Ωcm at room temperature for films with a doping concentration of 6.3%at. WTO has potential for TCO applications in electron transport layers in perovskite solar cells, with a demonstrated efficiency 28× greater than undoped TiO2. NTO and WTO have also attracted attention as promising photocatalysts with a 65% increase in photocurrent demonstrated for NTO nanorod photoelectrodes doped with 0.25%at Nb when compared to undoped TiO2. NTO nanostructures, ranging from rutile nanorods to anatase nanosheets, also show improved adsorption of molecular O2 and the formation of superoxide radicals O2 under irradiation when compared to pristine TiO2 with the same morphology. Furthermore, both NTO and WTO nanostructures show enhanced dye photodegradation, , photooxidation of nitrogen oxides (NO x ) to nitrates, and ozone gas sensing when compared to undoped TiO2.
To understand and further optimize the performance of NTO and WTO, it is necessary to characterize and model the behavior of electrons and holes that contribute to material conductivity and chemistry. This includes the mobility of charge carriers within the bands (conduction and valence, respectively), their recombination, their trapping (forming electron polarons), and the charge carrier transfer mechanisms that drive redox processes when these materials are used as semiconductor photocatalysts. Given the electronic spin associated with electrons and holes, electron paramagnetic resonance (EPR) spectroscopy is a powerful tool for the precise identification and characterization of the dynamics, lifetimes, and spatial distribution of excitons and their trapping states within NTO and WTO, including the paramagnetic species following insertion of Nb and W dopants in the TiO2 lattice. For example, electron trapping states are mostly associated with Nb4+ and W5+ species, which are both paramagnetic d 1 species that can be detected and interrogated by EPR spectroscopy. The formation of Nb4+ and W5+ versus vs. Nb5+ and W6+ is particularly important as it directly influences conductivity and photocatalytic efficiency.
The literature is conflicted with respect to the lack of Nb4+ and W5+ EPR signals when substitutional Nb and W are introduced in anatase TiO2, in contrast to the presence of Nb4+ and W5+ EPR signals in doped rutile TiO2. There is currently no clear explanation of this observation, and so far, first-principles atomistic modeling methods like density functional theory (DFT) have not completely clarified these experimental observations. For example, geometry optimization with semilocal DFT followed by a single point calculation using a screen exchange hybrid functional (sX) predicts shallow donor states that are largely delocalized over Ti sites in anatase NTO, supporting resonant photoemission experiments, which confirm the absence of midgap states for anatase NTO. , However, the same DFT calculations also predict a deep localized state in rutile NTO that is 0.9 eV below the conduction band edge, involving Ti 3d xy orbitals with a small contribution from Nb 4d orbitals, thus contradicting the EPR observations. Hubbard-corrected density functional theory (DFT+U) calculations predict Ti 3d midgap states in rutile NTO, but reports vary, with some predicting shallow donor states in rutile NTO and deep Ti 3d states in both anatase NTO , and anatase WTO, i.e., contradicting results. These computational results, based on DFT+U using planewave basis sets, fail to provide an unambiguous interpretation/prediction of the experimental observations; however, hybrid-DFT has been demonstrated to successfully predict W 5d midgap states in rutile WTO when using all-electron atom-centered basis functions to describe the wave function near the nucleus (in the full-potential linear augmented plane wave approach), and this gives promise for application of alternative basis representations.
In this work, experimental EPR spectra for powder anatase and rutile NTO and WTO are presented, including the magnetic tensors characterizing Nb4+ and W5+ polarons in doped rutile, alongside DFT+U simulations in an all-electron numerical atom-centered orbital (NAO) framework to successfully simulate the experimental EPR observations. The DFT+U simulations use a refined atomic-like Ti 3d Hubbard projector, as well as the “occupation matrix control” (OMC) method, for controlling the Ti 3d, Nb 4d, and W 5d orbital occupancies through the self-consistent optimization of the system electronic structure towards the ground state. The presented combination of experiment and theory improves our fundamental understanding of the nature and formation of reduced species in semiconductor metal oxides, enabling the rational design of superior TCOs and heterogeneous photocatalysts.
2. Methodology
2.1. Material Synthesis
The NTO and WTO materials studied in this work were synthesized via a sol–gel route. 10 mL of titanium isopropoxide (≥97%, Sigma-Aldrich) was dissolved in 10 mL of anhydrous ethanol. After thorough mixing, 5 mL of deionized water (18 MΩcm) was slowly added to the solution. The resulting white precipitate redissolved upon further stirring. In the next step, 20 mL of a pH 10 ammonia/ammonium chloride buffer (5% ammonia, Sigma-Aldrich) was added to the solution. Finally, the desired amount of ammonium tungstate (BDH Chemicals) or ammonium niobate (V) oxalate hydrate (BDH Chemicals) to achieve a nominal 0.1 or 1.0 atom % was dissolved in 10 mL of warm deionized water and subsequently added to the solution. After thorough stirring for at least 4 h, the solution was filtered, washed several times with deionized water, and then dried at 60 °C for 4 h. The dry powders were ground in an agate mortar and transferred into a crucible for calcination. The samples were calcined at 600 °C for 4 h and then ground again afterward.
2.2. Powder X-ray Diffraction
To confirm mineralogy and crystallinity, X-ray diffraction (XRD) patterns were obtained using a Bruker D8 Advance diffractometer equipped to deliver CuKα1 X-ray radiation (1.54 Å) at room temperature. Refinement of the powder XRD patterns was carried out using the Profex suite for XRD.
2.3. EPR Spectroscopy
X-band continuous-wave (CW) EPR spectra were recorded on a Bruker Elexsys E500 spectrometer equipped with an Oxford Instruments liquid-helium cryostat and a Bruker ER4122 SHQE-W1 superhigh Q resonator operating at 50 K. Before each measurement, the samples were evacuated for at least 12 h at 393 K and under a dynamic vacuum at ca. 1 × 10–4 bar. Experimental spectra were simulated using the EasySpin toolbox for Mathworks Matlab.
2.4. Electronic Structure Calculations
2.4.1. DFT
All electronic structure calculations were performed using the Fritz-Haber Institute ab initio materials simulation (FHI-aims) software package, which uses an all-electron NAO basis set, interfaced with the Python-based atomic simulation environment (ASE). The standard light basis set (2020) was used, with equivalent accuracy to the TZVP Gaussian-type orbital basis set, as decided after benchmarking the TiO2 formation energy (see the Supplementary Information, SI, Section S1.1). Relativistic effects were accounted for using the zeroth-order regular approximation as a scalar correction, while the system charge and spin were set to zero. Periodic boundary conditions were applied using converged k-point sampling for the optimized anatase and rutile unit cell, separately (see SI Section S1.1). The mBEEF meta-GGA exchange correlation functional was used, , as defined in Libxc, providing the best balance of accuracy and cost compared to other local, semilocal, and hybrid functionals (see SI Section S1.1). Self-consistent field (SCF) optimization of the electronic structure was achieved using a convergence criterion of 1 × 10–6 eV for the change in total energy, 1 × 10–4 eV for the change in the sum of eigenvalues, and 1 × 10–6 e a0 for the change in charge density. Geometry optimization used the quasi-Newton BFGS algorithm − with a force convergence criterion of 0.01 eV/Å. Point defect calculations were performed using a 3 × 3 × 3 TiO2 supercell containing 324 and 162 atoms for anatase and rutile, respectively. The supercell size avoids spurious long-range defect–defect interactions between periodic images while simulating at low defect concentrations of 0.308% (anatase) and 0.617% (rutile) following substitution of a Ti atom with a Nb or W atom. Defect energies were calculated as
| 1 |
For these systems, the chemical potential μ was calculated using the energy of bulk Ti (hexagonal close-packed structure), Nb (body-centered cubic structure), and W (body-centered cubic structure). All defect energies are listed in SI Section S1.3.3.
2.4.2. DFT+U
All DFT+U calculations were performed using the on-site definition of the occupation matrix and the fully localized limit double counting correction. A Hubbard correction was applied to correct for the Coulomb self-interaction of Ti 3d orbital electrons only. Both the Hubbard U value and the occupation matrix (n I ), which contains orbital occupancies for every spin channel (σ) for every atom (I), are used to calculate the corrective Hubbard term (E U ) as
| 2 |
Once E U is calculated, the total energy of the system (E DFT+U ) is calculated using the DFT+U Hamiltonian, where ρ(r) is the electron density, E DFT is the total energy of the system accounting for localized and delocalized orbitals, and E U prevents the double counting of localized states in both E DFT and E U :
| 3 |
The occupation matrix is calculated by projecting all DFT-predicted Kohn–Sham states onto atomic-like orbitals defined by a Hubbard projector. However, using the default atomic Ti 3d Hubbard projector presented challenges in identifying the ground-state electronic structures of anatase and rutile NTO and WTO due to numerical instability. Therefore, constrained and self-consistent DFT+U calculations were performed with an atomic and modified Ti 3d Hubbard projector, respectively. Constrained DFT+U calculations were performed using the default atomic Ti 3d Hubbard projector and a Ti 3d Hubbard U value of 3 eV in anatase TiO2 and 4 eV in rutile TiO2. These Hubbard U values were chosen to minimize the average error in the DFT+U-predicted band gap and unit cell equilibrium volume, calculated by fitting to the Birch–Murnaghan equation of state using ASE, versus experimental references (see SI Section S1.2). , Here, the OMC method was used to fix the polaron(s) at specific atom(s) by modifying the corresponding atomic orbital occupation matrix (see SI Section S1.3.1).
Next, self-consistent DFT+U calculations were performed with a refined atomic-like Ti 3d Hubbard projector. Here, the “occupation matrix release” method was used to initialize the polaron(s) at specific atom(s) before a two-step optimization of the system electronic structure until full geometry optimization was achieved (see SI Section S1.3.1). A refined atomic-like Ti 3d Hubbard projector was defined as a linear combination of the atomic Ti 3d and hydrogenic auxiliary basis function in the light basis set, where the auxiliary function is subject to a Gram–Schmidt orthogonalization with respect to the atomic function, with the corresponding linear combination expansion coefficients c 1 = 0.828 and c 2 = −0.561. The values of c 1 and c 2 were chosen based on the work of Jakob and Oberhofer, who computed a Ti 3d Hubbard projector for bulk rutile TiO2 in FHI-aims from first-principles. These coefficients enabled successful convergence to the ground state using a Ti 3d Hubbard U value of 3 eV for both anatase and rutile NTO and WTO. In general, tuning the Hubbard projector in FHI-aims to navigate numerical instability in the simulation of defective transition metal oxides has not yet been systematically investigated and will be the focus of a future study.
3. Results and Discussion
3.1. Polymorphism and Dopant Loadings
The refined X-ray powder diffraction patterns, presented in SI Section S2, show that 0.1%at. Nb or W in TiO2, synthesized via a sol–gel route and calcined at 600 K, allows for the formation of mixed anatase and rutile polymorphs. The same result occurs for undoped TiO2, which exhibits a rutilization temperature higher than 500 K (calcination of sol–gel TiO2 precursors with 0.1%at. Nb or W at T < 500 K generates anatase-only polymorphs ,, ). Refinement of the XRD patterns (SI Section S2) reveals 92% anatase and 8% rutile for NTO and 72% anatase and 28% rutile for WTO. These samples are referred to as NTO-AR and WTO-AR herein. On the contrary, 1.0%at. of the same dopants in conjunction with calcination at 600 K allows for the formation of anatase-only NTO and WTO (SI Section S2). These samples are referred to as NTO-A and WTO-A herein.
3.2. Experimentally Detected Electron Trapping States in Anatase and Rutile NTO and WTO
Electron and hole trapping are normally single electron transfer events occurring within the material. As such, they can be followed experimentally by detecting the formation or disappearance of paramagnetic species using EPR spectroscopy. Nb5+ ([Kr]4d 0) and W6+ ([Xe]5d 0) replace Ti4+ ([Ar]3d 0) in the TiO2 lattice aliovalently and isomorphically, as Nb5+, W6+, and Ti4+ have almost identical octahedral coordination environments with ionic radii of 78 pm, 74 pm, and 74.5 pm, respectively. The dopant incorporation reactions for extrinsic defects using the Kröger–Vink notation can be written as follows:
| 4 |
| 5 |
| 6 |
| 7 |
Valence-induced electron formation, as highlighted by eqs and , increases the n-type character of NTO and WTO compared to that of undoped TiO2. Our EPR attempts at identifying substitutional and isolated Nb4+ and W5+ in NTO-A and WTO-A failed, which corroborates with previous experiments by De Trizio et al., which could not detect Nb4+ in Nb-doped colloidal anatase nanocrystals even at liquid-helium temperature. Giamello and co-workers and ourselves independently showed that, in the case of Nb doping in anatase-only TiO2, Ti3+ is mostly formed as a result of valence induction when Nb5+ aliovalently replaces Ti4+ in the anatase lattice:
| 8 |
The resulting Ti3+ exhibits an anisotropic EPR spectrum characterized by a g tensor with axial symmetry and principal values equal to g ⊥ = 1.988 and g // = 1.957, , consistent with a highly delocalized bulk species that is responsible for causing an increased conductivity of the doped anatase TiO2. This Ti3+ is structurally and magnetically different from a surface-localized Ti3+ that forms following chemical or chemo/thermal reduction of undoped TiO2. The amount of delocalized bulk Ti3+ can also be augmented by photoinjection of extra conduction electrons , followed by trapping:
| 9 |
The situation appears completely different in the cases of NTO-AR and WTO-AR, as demonstrated by the respective X-band CW EPR spectra reported in Figure a (NTO-AR) and Figure b (WTO-AR).
1.

X-band CW EPR spectra at 50 K for (a) Nb-doped and (b) W-doped mixed-polymorph TiO2 nanoparticles with a doping concentration of 0.1%at. (NTO-AR and WTO-AR, respectively).
The clear axial signal in the spectrum of NTO-AR in Figure a can be attributed to bulk Ti3+, as previously described for the case of solely the anatase polymorph. A very broad and low-intensity signal is also visible at 50 K as shown by the arrows in Figure a. We propose that this broad signal is associated with Nb4+ in rutile. This assignment can be explained as follows. Figure b,c show the simulated angular dependency of the single-crystal EPR spectra at 4.2 K of Nb4+ centers in pure rutile NTO at the X- and Q-bands, respectively, computed using experimental values derived from a rutile single crystal.
2.
(a) Tetragonal crystal structure of rutile TiO2 with two inequivalent Ti atoms (I and II). (b)–(e) Simulated angular dependency of the rutile single-crystal EPR spectra at 4.2 K for Nb4+ centers at (b) X-band and (c) Q-band; and W5+ centers at (d) X-band and (e) Q-band.
The crystal structure of rutile TiO2 in Figure a is tetragonal (D 4h , space group P42/mnm) with two Ti per unit cell, which are equivalent except for a π/2 rotation about the c axis, as also evidenced by the orientation of the eigenframes of the magnetic tensors reported for the two inequivalent sites I and II in Figure a. The two Ti sites possess orthorhombic point symmetry D 2h . Details of the spin Hamiltonian used for the simulations can be found in Table .
1. Spin Hamiltonian Parameters of Reduced Dopant Metal Centers Detected in Nb5+ and W6+ Doped TiO2.
| Reduced metal center | TiO2 polymorph | g x | g y | g z | A x (MHz) | A y (MHz) | A z (MHz) | Reference |
|---|---|---|---|---|---|---|---|---|
| Nb4+ | rutile | 1.970 | 1.985 | 1.941 | n.d. | n.d. | n.d. | This work |
| Nb4+ | rutile | 1.973 | 1.981 | 1.948 | 5.0 | 23.8 | 7.0 | |
| W5+ | rutile | 1.594 | 1.473 | 1.443 | 277.5 | 122.4 | 191.1 | This work |
| W5+ | rutile | 1.5944 | 1.4725 | 1.4431 | 277.3 | 122.3 | 191.0 |
x, y, and z are here along [11̅0], [110], and [001], respectively. [001] corresponds to the rutile crystal c-axis (see also Figure ).
The g x and g y completely overlap at X-band frequencies, but they can be resolved at Q-band frequencies. Hyperfine interaction of the 4d 1 unpaired electron with the 93Nb nucleus is expected (I(93Nb) = 9/2, 100% natural abundance), giving rise to 10 hyperfine lines. The A y value is noticeably larger than A x and A z , with some resolved structure due to δm I = ± 1 transitions that are allowed via the electric quadrupole interaction when B is not aligned to the principal axes of the crystal. The hyperfine structure has been shown to vanish above 25 K on the single crystal, with particularly the B//[001] and B//[11̅0] sets of lines coalescing into a single line already at 25 K. The increasingly dispersive character of the signal above 25 K and the line narrowing observed upon coalescing of the hyperfine structure were interpreted by Zimmermann as a combination of two contributions, i.e., thermal excitation of 4d 1 donor electrons to the conduction band plus exchange scattering of 4d 1 donor electrons with the conduction band. Both of these contributions give rise to electron hopping from different donor sites (i.e., Nb4+) via the conduction band (Anderson’s model of random frequency modulation). The activation energy for the hopping between the donor level and the conduction band was found to be E a/k = 72 K at temperatures below 40 K, while the same activation energy was much larger at temperatures up to 300 K, in agreement with more recent evidence suggesting that rutile NTO is resistive at room temperature. Moving from single crystal to powder samples, Kiwi et al. showed that, in a mixed anatase and rutile powder sample, a broad signal could be found at 4.2 K matching the g tensor reported by Zimmermann. Our broad signal in Figure a matches the signal reported by Kiwi et al., although it is much broader due to the much higher temperature of our measurement, i.e., 50 K (according to Zimmermann, the signal completely vanishes above 77 K).
| 10 |
In the case of WTO-AR, the situation is very similar to that described above for NTO-AR. Figure b shows a clear anisotropic EPR spectrum characterized by a rhombic g tensor with principal values reported in Table . Small intensity doublets (m I = ± 1/2 lines) are visible on each side of the three principal resonances due to hyperfine interaction of the 5d 1 unpaired electron with the 183W nucleus (I(183W) = 1/2, 14.3% natural abundance). The other naturally occurring isotopes of W are 180W, 182W, 184W, and 186W, all with nuclear spin quantum number I = 0, and these account for the three principal m I = 0 resonance lines. W5+ in TiO2 is not affected by the same fast relaxation issues as Nb4+ and therefore well-defined EPR spectra can be easily obtained at 50 K, as visible in Figure b. The values of the magnetic tensors are in good agreement with those of Chang for W5+ centers in rutile single crystals. The angular dependency has been simulated in Figure d,e at the X- and Q-bands, respectively.
This combined evidence of NTO-AR and WTO-AR indicates that the reduction of Nb5+ and W6+ dopants in the TiO2 host lattice occurs only in the rutile polymorph.
3.3. Computed Electron Trapping States in Anatase and Rutile NTO and WTO
To rationalize the magnetic resonance observations in terms of the energetics of intraband gap states, we relied on self-consistent DFT+U calculations in an NAO framework, given that constrained DFT+U calculations could not rationalize our experimental observations (see SI Section S1.3.2). The total density of states (TDOS) reported in Figure a,c shows defect states pinned to the bottom of the TiO2 conduction band for both anatase NTO and WTO, respectively, in perfect agreement with our EPR observations for NTO-A and WTO-A. From the corresponding projected density of states (PDOS) in Figure e,g, there are small Nb 4d and W 5d signatures at the Fermi level indicating these states are partially delocalized over Ti sites, contributing to metallic-type behavior. On the other hand, self-consistent DFT+U calculations show that the Nb 4d signature in rutile NTO is an order of magnitude greater at the Fermi level than in anatase NTO, when normalizing with respect to the different defect concentrations in our simulation supercells, as shown in the PDOS in Figure e,f.
3.
Self-consistent DFT+U-predicted TDOS and PDOS for anatase NTO ((a) and (e), respectively), anatase WTO ((b) and (f), respectively), rutile NTO ((c) and (g), respectively), and rutile WTO ((d) and (h), respectively). All TDOS and PDOS are plotted relative to the Fermi level indicated by the red dashed line.
These results can be attributed to differences in the filling of the five Nb 4d orbitals, notably, the greater occupancy of the three t 2g orbitals in rutile NTO, which correspond to orbital magnetic quantum numbers m l = −2, −1, and 1 in Figure . There is a negligible difference in the trace of the Nb 4d occupation matrix (i.e., the total Nb 4d subshell occupancy) in anatase NTO (1.48) compared with that of rutile NTO (1.49). Figure d,h show the TDOS and W 5d PDOS for the rutile WTO, respectively. Here, a localized W 5d midgap state is predicted at ca. 0.7 eV below the TiO2 conduction band. The character of the midgap state is W 5d z 2 , corresponding to a large occupation number of 0.93 for the m l = 0 orbital in the W 5d occupation matrix (Figure ). The other diagonal terms of the W 5d occupation matrix are of similar magnitude in anatase WTO and rutile WTO, which suggests the formation of W5+ in rutile but not in anatase.
4.

Ground-state orbital occupation numbers for Nb 4d and W 5d in doped anatase and rutile TiO2 calculated using self-consistent DFT+U.
The formation of W5+ in rutile WTO is also suggested based on the local lattice distortion surrounding the W dopant, which is associated with localized polaronic states in defective TiO2, , as shown in Figure , which plots the change in the bond distance between the dopant atom and six neighboring Ti atoms relative to the average Ti–Ti bond distance in bulk anatase and rutile TiO2. Figure c shows symmetric geometric relaxation around the substitutional defect in anatase NTO and WTO, where the change in bond length between the dopant atom and Ti atoms A–D is almost constant for both materials, as is the change in bond length with Ti atoms E and F. In Figure d, there is a stronger asymmetric local lattice distortion around the dopant atom in rutile WTO compared to that in rutile NTO, as shown by the differences in the change in bond lengths between the dopant atom and Ti atoms G-L.
5.
Change in the self-consistent DFT+U-calculated bond distances between the dopant atom and surrounding Ti atoms in doped anatase (atoms A–F in (a)) and doped rutile (atoms G–L in (b)) calculated relative to the average Ti–Ti bond distance in bulk anatase (c) and rutile (d) TiO2.
4. Conclusions
NTO and WTO are promising TCOs with applications in heterogeneous photocatalysis and high-efficiency photovoltaics and electronics. However, there remains uncertainty in the atomistic mechanisms that govern charge compensation in these materials, which prevents the development of accurate structure–property models for material optimization. Using EPR spectroscopy, charge compensation is shown as highly sensitive to the TiO2 polymorph, with Nb4+ and W5+ signals present in substitutionally doped rutile but not in doped anatase. The observations are complemented by contemporary DFT+U calculations in an all-electron NAO framework, for which self-consistent resolution of the Ti 3d, Nb 4d, and W 5d orbital occupancies is crucial. Self-consistent DFT+U predicts the favorability of Nb4+ in rutile NTO through greater filling of the Nb 4d t 2g orbitals and reduced filling of the e g orbitals compared to anatase NTO. Self-consistent DFT+U also predicts W5+ in rutile WTO through the formation of a localized midgap state of 5d z 2 character that is not formed in anatase WTO. Our approach and findings provide a coherent view on the reducibility of metal centers in semiconducting TiO2, matching experiments and theory without apparent disagreement, while also providing a clear understanding of how the reducibility of metal centers and electron trapping energetics in TiO2 are polymorph-dependent. The improved fundamental understanding enables precise control over the electronic and magnetic properties of transition-metal-doped TiO2, advancing the rational design of high-performance materials for energy conversion and catalytic applications.
Supplementary Material
Acknowledgments
The authors thank Harald Oberhofer, Matthias Kick, and Maximilian Brand for valuable scientific discussions regarding the implementation of DFT+U in FHI-aims. The authors acknowledge funding by the Prosperity Partnership project Sustainable Catalysis for Clean Growth, funded by the UK Engineering and Physical Sciences Research Council (EPSRC), bp, through the bp International Centre for Advanced Materials (bp-ICAM) and Johnson Matthey plc in collaboration with Cardiff University and The University of Manchester (EPSRC grant number EP/V056565/1). A.F. acknowledges the Cardiff University Net Zero Innovation Institute for funding a University Research Fellowship at the Cardiff Catalysis Institute. A.J.L. acknowledges funding by the UKRI Future Leaders Fellowship program (MR/T018372/1 and MR/Y034279/1). The authors acknowledge computational resources and support from the Supercomputing Wales project, which is part-funded by the European Regional Development Fund (ERDF) via the Welsh Government; and the UK National Supercomputing Services ARCHER and ARCHER2, accessed via membership of the Materials Chemistry Consortium, which is funded by the Engineering and Physical Sciences Research Council (EP/L000202/1, EP/R029431/1, and EP/T022213/1).
The input and output files of all electronic structure calculations have been uploaded as a data set to the NOMAD repository at doi: 10.17172/NOMAD/2024.09.04-1.
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcc.5c04364.
Section S1, electronic structure calculations; and Section S2, powder XRD characterization (PDF)
A.F. and A.J.L. contributed to the conceptualization of the project and funding. A.J.L. and A.C. contributed to software choices and computational method development in this work. A.C. performed the electronic structure calculations presented in this work. A.F. performed the EPR spectroscopy and related simulations presented in this work. All authors contributed to the analysis of results and preparation of the manuscript.
The authors declare no competing financial interest.
References
- 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. doi: 10.1063/1.1949728. [DOI] [Google Scholar]
- Zhang S. X., Kundaliya D. C., Yu W., Dhar S., Young S. Y., Salamanca-Riba L. G., Ogale S. B., Vispute R. D., Venkatesan T.. Niobium doped TiO2: Intrinsic transparent metallic anatase versus highly resistive rutile phase. J. Appl. Phys. 2007;102:013701. doi: 10.1063/1.2750407. [DOI] [Google Scholar]
- Lee S., Noh J. H., Han H. S., Yim D. K., Kim D. H., Lee J. kun, Kim J. Y., Jung H. S., Hong K. S.. Nb-Doped TiO2: A New Compact Layer Material for TiO2 Dye-Sensitized Solar Cells. J. Phys. Chem. C. 2009;113:6878–6882. doi: 10.1021/jp9002017. [DOI] [Google Scholar]
- Tasaki C., Oka N., Yagi T., Taketoshi N., Baba T., Kamiyama T., Nakamura S. ichi, Shigesato Y.. Thermophysical Properties of Transparent Conductive Nb-Doped TiO2 Films. Jpn. J. Appl. Phys. 2012;51:035802. doi: 10.7567/JJAP.51.035802. [DOI] [Google Scholar]
- Lee H. Y., Robertson J.. Doping and compensation in Nb-doped anatase and rutile TiO2 . J. Appl. Phys. 2013;113:213706. doi: 10.1063/1.4808475. [DOI] [Google Scholar]
- De Trizio L., Buonsanti R., Schimpf A. M., Llordes A., Gamelin D. R., Simonutti R., Milliron D. J.. Nb-Doped Colloidal TiO2 Nanocrystals with Tunable Infrared Absorption. Chem. Mater. 2013;25:3383–3390. doi: 10.1021/cm402396c. [DOI] [Google Scholar]
- Park J. T., Chi W. S., Jeon H., Kim J. H.. Improved electron transfer and plasmonic effect in dye-sensitized solar cells with bi-functional Nb-doped TiO2/Ag ternary nanostructures. Nanoscale. 2014;6:2718–2729. doi: 10.1039/C3NR05652J. [DOI] [PubMed] [Google Scholar]
- Biedrzycki J., Livraghi S., Giamello E., Agnoli S., Granozzi G.. Fluorine- and Niobium-Doped TiO2: Chemical and Spectroscopic Properties of Polycrystalline n-Type-Doped Anatase. J. Phys. Chem. C. 2014;118:8462–8473. doi: 10.1021/jp501203h. [DOI] [Google Scholar]
- Folli A., Bloh J. Z., Walker R., Lecaplain A., Macphee D. E.. Properties and Photochemistry of Valence-Induced-Ti3+ Enriched (Nb,N)-Codoped Anatase TiO2 Semiconductors. Phys. Chem. Chem. Phys. 2015;17:4849–4853. doi: 10.1039/C4CP05521G. [DOI] [PubMed] [Google Scholar]
- Yue J., Suchomski C., Voepel P., Ellinghaus R., Rohnke M., Leichtweiss T., Elm M. T., Smarsly B. M.. Mesoporous niobium-doped titanium dioxide films from the assembly of crystalline nanoparticles: study on the relationship between the band structure, conductivity and charge storage mechanism. J. Mater. Chem. A. 2017;5:1978–1988. doi: 10.1039/C6TA06840E. [DOI] [Google Scholar]
- Phipps G., Mikolajczak C., Guckes T.. Indium and Gallium: long-term supply. Renew. Energy Focus. 2008;9:56–59. doi: 10.1016/S1471-0846(08)70140-9. [DOI] [Google Scholar]
- Green M. A.. The Path to 25% Silicon Solar Cell Efficiency: History of Silicon Cell Evolution. Prog. Photovolt: Res. Appl. 2009;17:183–189. doi: 10.1002/pip.892. [DOI] [Google Scholar]
- Vesborg P. C. K., Jaramillo T. F.. Addressing the terawatt challenge: scalability in the supply of chemical elements for renewable energy. RSC Adv. 2012;2:7933. doi: 10.1039/c2ra20839c. [DOI] [Google Scholar]
- Lokanc, M. ; Eggert, R. ; Redlinger, M. . The Availability of Indium: The Present, Medium Term, and Long Term, 2015, www.nrel.gov/publications.\%0Awww.nrel.gov/publications.\%0Ahttps://www.nrel.gov/docs/fy16osti/62409.pdf. [Google Scholar]
- Chen D.-m., Xu G., Miao L., Chen L.-h., Nakao S., Jin P.. W-doped anatase TiO2 transparent conductive oxide films: Theory and experiment. J. Appl. Phys. 2010;107:063707. doi: 10.1063/1.3326940. [DOI] [Google Scholar]
- Hou Q., Zhao C., Guo S., Mao F., Zhang Y.. Effect on electron structure and magneto-optic property of heavy W-doped anatase TiO2 . PLoS One. 2015;10:e0122620. doi: 10.1371/journal.pone.0122620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang H., Zhao C., Yin L., Li X., Tu X., Lim G., Liu Y., Zhao Z.. W-doped TiO2 as electron transport layer for high performance solution-processed perovskite solar cells. Appl. Surf. Sci. 2021;563:150298. doi: 10.1016/j.apsusc.2021.150298. [DOI] [Google Scholar]
- Wang H. Y., Yang H., Zhang L., Chen J., Liu B.. Niobium Doping Enhances Charge Transport in TiO2 Nanorods. ChemNanoMat. 2016;2:660–664. doi: 10.1002/cnma.201600035. [DOI] [Google Scholar]
- Wang H. Y., Chen J., Xiao F. X., Zheng J., Liu B.. Doping-induced structural evolution from rutile to anatase: Formation of Nb-doped anatase TiO2 nanosheets with high photocatalytic activity. J. Mater. Chem. A. 2016;4:6926–6932. doi: 10.1039/C5TA08202A. [DOI] [Google Scholar]
- Santos E., Catto A. C., Peterline A. F., Avansi W.. Transition metal (Nb and W) doped TiO2 nanostructures: The role of metal doping in their photocatalytic activity and ozone gas-sensing performance. Appl. Surf. Sci. 2022;579:152146. doi: 10.1016/j.apsusc.2021.152146. [DOI] [Google Scholar]
- 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. doi: 10.1038/srep10952. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Folli A., Bloh J. Z., Armstrong K., Richards E., Murphy D. M., Lu L., Kiely C. J., Morgan D. J., Smith R. I., Mclaughlin A. C., Macphee D. E.. Improving the Selectivity of Photocatalytic NOx Abatement through Improved O2 Reduction Pathways Using Ti0.909W0.091O2Nbx Semiconductor Nanoparticles: From Characterization to Photocatalytic Performance. ACS Catal. 2018;8:6927–6938. doi: 10.1021/acscatal.8b00521. [DOI] [Google Scholar]
- Hitosugi T., Kamisaka H., Yamashita K., Nogawa H., Furubayashi Y., Nakao S., Yamada N., Chikamatsu A., Kumigashira H., Oshima M., Hirose Y., Shimada T., Hasegawa T.. Electronic band structure of transparent conductor: Nb-doped anatase TiO2 . Appl. Phys. Express. 2008;1:111203. doi: 10.1143/APEX.1.111203. [DOI] [Google Scholar]
- Ghuman K. K., Singh C. V.. A DFT + U study of (Rh, Nb)-codoped rutile TiO2 . J. Phys.: Condens. Matter. 2013;25:085501. doi: 10.1088/0953-8984/25/8/085501. [DOI] [PubMed] [Google Scholar]
- 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. doi: 10.1039/b905028k. [DOI] [Google Scholar]
- Raghav A., Hongo K., Maezono R., Panda E.. Electronic structure and effective mass analysis of doped TiO2 (anatase) systems using DFT + U . Comput. Mater. Sci. 2022;214:111714. doi: 10.1016/j.commatsci.2022.111714. [DOI] [Google Scholar]
- Belošević-Čavor J., Koteski V., Umićević A., Ivanovski V.. Effect of 5d transition metals doping on the photocatalytic properties of rutile TiO2 . Comput. Mater. Sci. 2018;151:328–337. doi: 10.1016/j.commatsci.2018.05.022. [DOI] [Google Scholar]
- Kick M., Reuter K., Oberhofer H.. Intricacies of DFT+U, Not Only in a Numeric Atom Centered Orbital Framework. J. Chem. Theory Comput. 2019;15:1705–1718. doi: 10.1021/acs.jctc.8b01211. [DOI] [PubMed] [Google Scholar]
- Allen J. P., Watson G. W.. Occupation matrix control of d- and f-electron localisations using DFT+U . Phys. Chem. Chem. Phys. 2014;16:21016–21031. doi: 10.1039/C4CP01083C. [DOI] [PubMed] [Google Scholar]
- Doebelin N., Kleeberg R.. Profex: A graphical user interface for the Rietveld refinement program BGMN. J. Appl. Crystallogr. 2015;48:1573–1580. doi: 10.1107/S1600576715014685. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stoll S., Schweiger A.. EasySpin, a Comprehensive Software Package for Spectral Simulation and Analysis in EPR. J. Magn. Reson. 2006;178:42–55. doi: 10.1016/j.jmr.2005.08.013. [DOI] [PubMed] [Google Scholar]
- Blum V., Gehrke R., Hanke F., Havu P., Havu V., Ren X., Reuter K., Scheffler M.. Ab initio molecular simulations with numeric atom-centered orbitals. Comput. Phys. Commun. 2009;180:2175–2196. doi: 10.1016/j.cpc.2009.06.022. [DOI] [Google Scholar]
- Larsen A. H., Mortensen J. J., Blomqvist J., Castelli I. E., Christensen R., Dułak M., Friis J., Groves M. N., Hammer B., Hargus C., Hermes E. D., Jennings P. C., Jensen P. B., Kermode J., Kitchin J. R., Kolsbjerg E. L., Kubal J., Kaasbjerg K., Lysgaard S., Maronsson J. B., Maxson T., Olsen T., Pastewka L., Peterson A., Rostgaard C., Schio̷tz J., Schütt O., Strange M., Thygesen K. S., Vegge T., Vilhelmsen L., Walter M., Zeng Z., Jacobsen K. W.. The atomic simulation environmenta Python library for working with atoms. J. Phys.: Condens. Matter. 2017;29:273002. doi: 10.1088/1361-648X/aa680e. [DOI] [PubMed] [Google Scholar]
- Lamiel-Garcia O., Ko K. C., Lee J. Y., Bromley S. T., Illas F.. When Anatase Nanoparticles Become Bulklike: Properties of Realistic TiO2 Nanoparticles in the 1–6 nm Size Range from All Electron Relativistic Density Functional Theory Based Calculations. J. Chem. Theory Comput. 2017;13:1785–1793. doi: 10.1021/acs.jctc.7b00085. [DOI] [PubMed] [Google Scholar]
- Wellendorff J., Lundgaard K. T., Jacobsen K. W., Bligaard T.. mBEEF: An accurate semi-local Bayesian error estimation density functional. J. Chem. Phys. 2014;140:144107. doi: 10.1063/1.4870397. [DOI] [PubMed] [Google Scholar]
- Perdew J. P., Ruzsinszky A., Csonka G. I., Vydrov O. A., Scuseria G. E., Constantin L. A., Zhou X., Burke K.. Restoring the Density-Gradient Expansion for Exchange in Solids and Surfaces. Phys. Rev. Lett. 2008;100:136406. doi: 10.1103/PhysRevLett.100.136406. [DOI] [PubMed] [Google Scholar]
- Lehtola S., Steigemann C., Oliveira M. J., Marques M. A.. Recent developments in libxc A comprehensive library of functionals for density functional theory. SoftwareX. 2018;7:1–5. doi: 10.1016/j.softx.2017.11.002. [DOI] [Google Scholar]
- Broyden C. G.. The Convergence of a Class of Double-rank Minimization Algorithms 1 General Considerations, IMA. J. Appl. 1970;6:76–90. doi: 10.1093/imamat/6.1.76. [DOI] [Google Scholar]
- Fletcher R.. A new approach to variable metric algorithms. Comput. J. 1970;13:317–322. doi: 10.1093/comjnl/13.3.317. [DOI] [Google Scholar]
- Shanno D. F.. Conditioning of Quasi-Newton Methods for Function Minimization. Math. Comput. 1970;24:647–656. doi: 10.1090/S0025-5718-1970-0274029-X. [DOI] [Google Scholar]
- Goldfarb D.. A Family of Variable-Metric Methods Derived by Variational Means. Math. Comput. 1970;24:23–26. doi: 10.1090/S0025-5718-1970-0258249-6. [DOI] [Google Scholar]
- Birch F.. Finite Elastic Strain of Cubic Crystals. Phys. Rev. 1947;71:809–824. doi: 10.1103/PhysRev.71.809. [DOI] [Google Scholar]
- Kavan L., Grätzel M., Gilbert S. E., Klemenz C., Scheel H. J.. Electrochemical and Photoelectrochemical Investigation of Single-Crystal Anatase. J. Am. Chem. Soc. 1996;118:6716–6723. doi: 10.1021/ja954172l. [DOI] [Google Scholar]
- Arlt T., Bermejo M., Blanco M. A., Gerward L., Jiang J. Z., Olsen J. S., Recio J. M.. High-pressure polymorphs of anatase TiO 2 . Phys. Rev. B. 2000;61:14414. doi: 10.1103/PhysRevB.61.14414. [DOI] [Google Scholar]
- Jakob, K. ; Oberhofer, H. . Master’s thesis: Self-Consistency in the Hubbard-Corrected DFT+U Method; Technical University of Munich: Faculty of Chemistry, 2021. [Google Scholar]
- Folli A., Bloh J., Beukes E.-P., Howe R., MacPhee D.. Photogenerated charge carriers and paramagnetic species in (W,N)-codoped TiO2 photocatalysts under visible-light irradiation: An EPR study. J. Phys. Chem. C. 2013;117:22149. doi: 10.1021/jp408181r. [DOI] [Google Scholar]
- Bloh J. Z., Folli A., Macphee D. E.. Adjusting Nitrogen Doping Level in Titanium Dioxide by Codoping with Tungsten: Properties and Band Structure of the Resulting Materials. J. Phys. Chem. C. 2014;118:21281–21292. doi: 10.1021/jp507264g. [DOI] [Google Scholar]
- Zimmermann P. H.. Temperature Dependence of the EPR Spectra of Niobium-Doped TiO2 . Phys. Rev. B. 1973;8:3917–3927. doi: 10.1103/PhysRevB.8.3917. [DOI] [Google Scholar]
- Chang T. T.. Paramagnetic-Resonance Spectrum of W5+ in Rutile (TiO2) Phys. Rev. 1966;147:264–267. doi: 10.1103/PhysRev.147.264. [DOI] [Google Scholar]
- Kiwi J., Suss J. T., Szapiro S.. EPR spectra of niobium-doped TiO2 and implications for water photocleavage processes. Chem. Phys. Lett. 1984;106:135–138. doi: 10.1016/0009-2614(84)87027-X. [DOI] [Google Scholar]
- Lin C., Shin D., Demkov A. A.. Localized states induced by an oxygen vacancy in rutile TiO2 . J. Appl. Phys. 2015;117:225703. doi: 10.1063/1.4922184. [DOI] [Google Scholar]
- Yim C. M., Watkins M. B., Wolf M. J., Pang C. L., Hermansson K., Thornton G.. Engineering Polarons at a Metal Oxide Surface. Phys. Rev. Lett. 2016;117:116402. doi: 10.1103/PhysRevLett.117.116402. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The input and output files of all electronic structure calculations have been uploaded as a data set to the NOMAD repository at doi: 10.17172/NOMAD/2024.09.04-1.



