Next Article in Journal
Additive-Free Baeyer–Villiger Oxidation of Cyclic Ketone Catalyzed by Carboxylic-Functionalized Poly(Ionic Liquids) and Polyoxometalate Ionic Self-Assemblies
Next Article in Special Issue
Preparation of Quasi-MIL-101(Cr) Loaded Ceria Catalysts for the Selective Catalytic Reduction of NOx at Low Temperature
Previous Article in Journal
Selective Hydration of Nitriles to Corresponding Amides in Air with Rh(I)-N-Heterocyclic Complex Catalysts
Previous Article in Special Issue
Structural Evolution of Ni-Based Co-Catalysts on [Ca2Nb3O10] Nanosheets during Heating and Their Photocatalytic Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Niobium-Based Materials for Photocatalytic Solar Fuel Production

by
Barbara Nascimento Nunes
1,2,
Osmando Ferreira Lopes
2,
Antonio Otavio T. Patrocinio
2,* and
Detlef W. Bahnemann
1,3,*
1
Institute of Technical Chemistry, Leibniz University Hannover, 30167 Hannover, Germany
2
Laboratory of Photochemistry and Materials Science, Institute of Chemistry, Federal University of Uberlandia, Uberlandia 38400-902, Brazil
3
Laboratory ‘Photoactive Nanocomposite Materials’, Saint-Petersburg State University, Saint-Petersburg 199034, Russia
*
Authors to whom correspondence should be addressed.
Submission received: 15 December 2019 / Revised: 6 January 2020 / Accepted: 11 January 2020 / Published: 16 January 2020
(This article belongs to the Special Issue Photocatalytic Nanocomposite Materials)

Abstract

:
The search for renewable and clean energy sources is a key aspect for sustainable development as energy consumption has continuously increased over the years concomitantly with environmental concerns caused by the use of fossil fuels. Semiconductor materials have great potential for acting as photocatalysts for solar fuel production, a potential energy source able to solve both energy and environmental concerns. Among the studied semiconductor materials, those based on niobium pentacation are still shallowly explored, although the number of publications and patents on Nb(V)-based photocatalysts has increased in the last years. A large variety of Nb(V)-based materials exhibit suitable electronic/morphological properties for light-driving reactions. Not only the extensive group of Nb2O5 polymorphs is explored, but also many types of layered niobates, mixed oxides, and Nb(V)-doped semiconductors. Therefore, the aim of this manuscript is to provide a review of the latest developments of niobium based photocatalysts for energy conversion into fuels, more specifically, CO2 reduction to hydrocarbons or H2 evolution from water. Additionally, the main strategies for improving the photocatalytic performance of niobium-based materials are discussed.

1. Introduction

Increasing energy demand concomitantly with global warming is a major world concern. Currently, fuels and electricity supplies are still strongly dependent on nonrenewable sources. For instance, in 2018, the worldwide consumption of fossil energy was almost 36.5 billion barrels of oil, 3850 billion cubic meters of natural gas, and 3800 million tons oil equivalent of coal per year. As a consequence, more than 33,000 million tons of carbon dioxide was emitted, which represents an annual increasing of 2.0%, the fastest growth for seven years [1]. This scenario can trigger serious problems for the current and next generations such as the insufficiency of fossil fuel resources, air pollution, and also climate changes as a result of the greenhouse effect [2].
In the face of this issue, the development of methods to obtain energy in a clean and sustainable way becomes imperative. Therefore, solar light as a primary energy resource stands out for its advantages in terms of spread availability and accessibility in an inexhaustible and inexpensive way around the world. In addition, obtaining energy from the sun may cause less damage to the environment in relation to waste production, polluting gases emission, or impacts to different ecosystems. The possibility to convert light into valuable solar fuels from water or even CO2 is a promising approach to provide clean energy and, at the same time, contribute to diminishing the global warming effect [3]. Additionally, the total sum of renewable energy from all the reserves is around 1% of the solar energy supplied to the earth surface by the sun, that is, 173,000 TW. This amount of energy is more than 9000 times larger than the current total energy consumption of the world (17.91 TW in 2017). The challenge lies on the development of economically viable technologies for solar energy harvesting, storage, and utilization [4].
Thus, since Fujishima and Honda [5] reported the possibility of promoting the photo-oxidation of water on the surface of TiO2 photoanodes, semiconductor materials have been extensively investigated as photocatalysts to store solar energy in chemical bonds. Among different photocatalysts, it is worth pointing out Nb(V)-based materials that have been investigated owing to their suitable electronic and morphological properties for light-driving reactions. Nb(V) is characterized by its high affinity to oxygen and its oxides exhibit different properties depending on the preparation and desired application [6,7]. Niobium pentoxide is considered the most thermodynamically stable Nb-based compound with more than 15 different structural configurations. Depending on the crystalline phase, its band-gap energy can vary from 3.2 eV to larger values such as 5 eV [8,9,10].
Besides Nb2O5 polymorphs, Nb(V) can also combine with oxygen to form extended lamellar structures resulting in polyoxy anions, the niobates. They are basically constituted of stacked negative charged layers with intercalated cations [11]. This configuration provides a considerable surface area and allows easy modification either by intercalation, superficial modification, or formation of nanosheets and nanoscrolls [12,13,14]. Furthermore, Nb(V) is also frequently applied as a doping agent. It is commonly combined to TiO2 and considered one of the most promising dopants for modifying the matrix crystalline structure and minimize the electron–hole recombination [15,16,17,18].
In this paper, the recent developments of niobium-based materials for photocatalytic solar fuel production, especially those in the last three years, are systematically reviewed. The main strategies for improving the photocatalytic performance for H2 production from water splitting and CO2 reduction to hydrocarbons, focused on niobium species, will be presented. In Section 2, application to H2 evolution is discussed, followed by Section 3 related to CO2 reduction.

2. H2 Evolution and Water Splitting

2.1. Niobium Pentoxide (Nb2O5)

Nb2O5 is a typical n-type semiconductor that, along with other oxides, has been widely studied as a photocatalyst thanks to its electronic properties such as excellent chemical and thermal stability, large abundance, rich morphologies, and several polymorphs [19,20,21,22]. It is a white solid powder, insoluble in water, and can be dissolved just by fusion with strong basic or acid fluxes. In general, its structure is formed by NbO6 octahedral, which can result in an amorphous state or may crystallize in a wide range of polymorphs with different physical properties [6,8]. Some of those polymorphs are pseudohexagonal (TT-Nb2O5), orthorhombic (T-Nb2O5), and monoclinic (B, H, R, and N-Nb2O5) [19]. The most known nomenclature is based on Brauer’s system, in which the crystalline phases are identified by letters corresponding to the temperature at which they are obtained [8].
Despite these properties, bulk Nb2O5 is not able to achieve a considerable photoactivity owing to its low specific surface area and fast recombination rate of the photogenerated charge carriers [21]. In this sense, several approaches have been applied to increase its photocatalytic performance. One of them is to control the shape and morphology of particles in order to tune the Nb2O5 properties. This can be possible, for instance, by selecting a suitable precursor to obtain the desired configuration of Nb2O5. Wen and coauthors firstly synthesized orthorhombic Nb2O5 (O-Nb2O5) by a hydrothermal reaction using K4Nb6O17 4.5H2O as a precursor. This sample exhibited rectangle nanosheets with a dominantly exposed (010) plane (Figure 1a–d) [22]. After acid treatment of K4Nb6O17 4.5H2O to get H4Nb6O17 3H2O, the ion-exchanged material was treated hydrothermally for 24 h at different conditions of temperature (195–215 °C) and pH (1.5–11.5), as shown in Figure 1e. At first, by changing the temperature at pH 3.5, X ray diffraction (XRD) results showed that the O-Nb2O5 phase began to be formed at 200 °C, and at 210 °C, the peaks of H4Nb6O17 3H2O disappeared completely and all observed peaks could be indexed as the O-Nb2O5 phase. From the Figure 1e, it can be also observed that the O-Nb2O5 was obtained in all selected pHs; however, the authors reported that, at pH 3.5, the material showed higher crystallinity and more regular morphology. A commercial O-Nb2O5 sample with spherical morphology was used for comparison and exhibited a band gap value of 3.2 eV, in contrast to 3.0 eV for the as synthesized O-Nb2O5 rectangle nanosheets. Then, the photoelectrochemical activity of both samples was evaluated by measuring the photocurrent resulted from H2 evolution under 370 nm UV light irradiation (30 W·m−2). Both photoelectrodes show very low dark current, but under illumination, the O-Nb2O5 nanosheet-based photoelectrode exhibited photocurrents 4.3 times higher than the commercial sample. The authors could justify the improved performance to a larger number of charge carriers coming from the exposed (010) facet with lower band gap energy, fewer recombination centers, and also faster charge transfer by the regular rectangle nanosheets with high crystallinity.
Later, the same strategy was employed by the same group to obtain crystalline hexagonal Nb2O5 (H-Nb2O5) nanobelt clusters with an exposed (−110) plane [23]. In this case, a fiber-like K2Nb2O6·H2O material was used as precursor. In order to improve the photocatalytic performance of H-Nb2O5 under visible light, metallic Ag was deposited by in situ photoreduction. UV/vis diffuse reflectance spectra showed that the absorption edges were 375 nm for the H-Nb2O5 nanobelt clusters and 423 nm for the Ag-modified sample. Then, both samples were evaluated photoelectrochemically under simulated sunlight illumination (100 W m−2) at zero applied voltage. The Ag-modified sample achieved a photocurrent of 10.73 μA, 535 times higher than that observed for bare H-Nb2O5, which was attributed to the improved light-harvesting efficiency as well as the better charge transfer performance.
Zhou and coauthors reported monoclinic Nb2O5 nanorod superstructures by a similar route using Sn2Nb2O7 nanoparticles as the precursor [24]. SEM images showed that Nb2O5 nanorods were obtained with lengths around 2 µm and diameters of 200 nm. The light absorption properties of the Nb2O5 nanorods were compared to those of commercial Nb2O5 powders and a considerable blue shift was observed in the absorption edge, which could be related to the nanosize of the former. Photocatalytic H2 production was also evaluated with 0.5 wt% Pt as co-catalyst in 25% methanol aqueous solution, employing a 500 W high-pressure mercury lamp as light source. The Nb2O5 nanorod superstructures sample showed a H2 evolution rate of 91 µmol h−1, whereas the commercial one exhibited a rate of 0.6 µmol h−1, ca. 150 times lower than that for the Nb2O5 nanorod. Moreover, Nb2O5 nanorod superstructures maintained an almost constant photocatalytic H2 production rate after 24 h. The higher performances were explained by the authors based on the superstructures of the material (Figure 2), in which photogenerated electrons could be transferred along the 1D nanorods direction and also transported to the neighboring nanorods, leading to a more efficient separation of photogenerated charge carriers.
Clearly, such morphological and structural transformations lead to changes in the optical and electronic properties of niobium oxide. In this sense, Zhang et al. reported an enclosed porous structure of pseudo-hexagonal tief-tief (TT) Nb2O5 nanowires, which showed a significant decrease in the band gap from 3.22 to 2.95 eV in relation to the nonporous Nb2O5 nanowires, and then an enhancement of H2 evolution under visible light irradiation (λ > 420 nm) [25]. The material was prepared by a solvothermal synthesis using ammonium niobate(V) oxalate hydrate and oleic acid mixed with trioctylamine. The sample was calcined at different temperatures. Thermal treatment at 600 °C and 700 °C causes the formation of orthorhombic phase, while at 580 °C, the pristine TT-phase was formed, so this last condition was selected for further analysis. In TEM images, it was observed that the porous nanowires were shorter than 1 µm and exhibited well separated pores inside the structure, whereas nonporous ones were longer than 5 µm and showed ripple-like contrast owing to bending-induced strain. The authors could explain the formation of such porosity to the replacement of oxalate anion by oleic acid and trioctylamine in the Nb center, resulting in a structure with hydrophobic core and surrounding Nb species. Then, the subsequent heating treatment induced the formation of pores from the cores and the Nb2O5 crystalline nanowires from the coordination structure. However, the samples were not characterized by isotherms of N2 adsorption to be properly classified as porous material. The authors suggested that under-coordinated ions in the twisted lattice around the pores could generate mid-gap states, and hence lower band gap energies. As a result, the H2 production rate of the porous sample was 243.8 μmol g−1 h−1, considerable higher than the nonporous one, which exhibited H2 evolution rate of 113.1 μmol g−1 h−1 under similar conditions, and even higher than that for P25 TiO2 (129.6 μmol g−1 h−1).
Besides the morphology, other strategies can be applied to increase the photocatalytic activity of Nb2O5 species. The introduction of defects into the crystal lattice could be used to engineer the electronic transport properties along the oxide structure or even to create new trap states inside the forbidden band gap region [26]. Zhao and coauthors reported improved solar-light absorption via partial reduction of Nb2O5 nanorods with active exposed (001) surfaces, which result in the black niobium oxide [27]. In summary, the previous prepared Nb2O5 was thermally treated in an aluminum reduction device, where the sample and aluminum powder were positioned independently in a two-zone tube oven, and then the Nb2O5 and aluminum powders were heated to 500 °C and 800 °C, respectively. This procedure led to the formation of Nb4+ centers and hence substantial oxygen vacancies that resulted in a black powder, and thus an enhancement in visible and infrared light absorption compared with the pristine Nb2O5. The reduction of Nb2O5 could be confirmed by Raman and X-ray photoelectron spectroscopies. The Nb4+ centers were also characterized by electron paramagnetic resonance. High resolution transmission electron microscopy indicated predominant (001) crystal plane exposed on the nanorod surfaces, suggesting good crystallinity of samples before and after reduction. The samples were evaluated as photoanode using dip-coated FTO glasses in a 1 M NaOH electrolyte under 100 mW cm−2 illumination with a 150 W Xe lamp. At 1.23 V, the reduced Nb2O5 sample showed a photocurrent density 138 times higher than that for Nb2O5. The carrier density was calculated by Mott–Schottky plots and was 3 × 104 times higher for the reduced sample, evidencing an enhanced electrical conductivity, charge transport, and separation. Finally, the samples were evaluated to produce H2 with 0.5 wt% Pt and 20% methanol solution under full-sunlight irradiation from a 300 W Xe lamp, which, for the reduced Nb2O5, resulted in 13 times higher production rate of 274.8 µmol h−1 g−1 in comparison with Nb2O5 nanorods (21.1 µmol h−1 g−1).
An extended visible light activity of orthorhombic Nb2O5 was achieved by Kulkarni et al. as result of nitrogen doping using a solid state reaction of urea and niobium salt [20]. At 500 °C, ammonia is released and works as nitrogen source to be introduced into the Nb2O5 lattice, filling oxygen vacancies. Field emission scanning electron microscopy (FE-SEM) analysis showed that the urea concentration was crucial for the observed morphologic changes. Concerning the optical properties, the modified samples were analyzed by UV/vis diffuse reflectance and compared to the pristine Nb2O5, which had a band gap energy of 3.4 eV. In general, all the N-doped Nb2O5 samples showed significant absorption in the range of 400–600 nm and a higher red shift in the band gap absorption when the urea concentration was increased (2.6–2.4 eV). Density functional theory (DFT) calculations described that N 2p bands falls above the O 2p bands, and so the top of the valence band gives rise to the reduction in the band gap. At the highest Nb/urea ratio investigated (1:15), this trend was not observed and a higher band gap edge than the expected trend (2.5 eV) was obtained. This would have resulted from nitrogen saturation of the niobia matrix. Photoluminescence measurements indicate that the optimal Nb/urea ratio was 1:10. This sample exhibited the lowest luminescence intensity among the other samples, indicating the efficient electron/hole separation. These results were in agreement with the photocatalytic H2 evolution experiments.
Another strategy to achieve higher photocatalytic performances is coupling to different Nb2O5 compounds with special attributes to create an efficient heterostructured photocatalyst. Owing to its flexibility in morphological, electronic, and structural properties, Nb2O5 is a promising candidate to design an effective heterojunction. Huang and coauthors reported a synergic combination of highly ordered mesoporous Nb2O5 (MNb) nanostructure to two-dimensional N-doped graphene (NGR) through a good interfacial connection (Figure 3) [28]. UV/visible diffuse reflectance spectroscopy evidenced gradual shifts to longer wavelengths as the NGR quantity was increased, so the band gap energies were shift from 3.12 eV up to 2.67 eV. Photoluminescence analysis suggested that NGR supported an efficient separation of charge carriers owing to the fast electron transfer from the conduction band (CB) of MNb to NGR sheets.
Huang and coauthors reported an in situ growth of mesoporous Nb2O5 microspheres on the surface of g-C3N4 [21]. In this case, assuming the photocatalytic activity of g-C3N4 under visible light, the enhanced photoactivity was mainly achieved thanks to the effective interfacial charge transfer between g-C3N4 and Nb2O5. Therefore, it is mandatory to ensure sufficient interfacial contact, and this condition was succeeded by the authors via a so-called in situ self-assembly process of synthesis. The introduction of Nb2O5 microspheres on g-C3N4 caused an expressive enhancement on the specific surface area. UV/visible diffuse reflectance spectroscopy showed that g-C3N4 and the composites presented an absorption edge at ca. 450 nm, while pure Nb2O5 had slight a band edge of ca. 420 nm. The slightly visible light absorption of Nb2O5 was justified by traces of carbonate species from the thermal decomposition of the triblock copolymer used in the synthesis. Photoluminescence spectroscopy evidenced that pure g-C3N4 exhibits stronger emission than the composites. Photocatalytic activities for H2 evolution over the samples were evaluated and the composite with 38.1 wt% of Nb2O5 exhibited the highest H2 evolution rate. With higher amounts of the Nb2O5 (69.6 wt%), the photocatalytic performance was still higher than pure g-C3N4, but slightly decreased in relation to the composite with 38.1 wt% of Nb2O5. This fact could be related to a less effective interfacial contact between the two semiconductors and to a considerable reduction of light absorption by the composite. Other successful combinations between Nb2O5 and g-C3N4 are reported elsewhere [29,30].
Therefore, it can be observed that Nb2O5 pristine or modified has been employed successfully as a photocatalyst for H2 evolution. The role of the crystalline phase, morphology/shape, defects/oxygen vacancies, and heterojunction formation were investigated and showed great contributions for H2 evolution activity of Nb2O5.

2.2. Niobium Layered Compounds

Layered niobates are composed by a repetition of [NbO6] octahedral units connected through adjacent or opposite sharing edges or corners, resulting in an extended 2D layered arrangement. This configuration is built by the stacking of negative charged layers separated by cations in the interlayer space [11,31]. The first niobate phases were reported in 1955 by Reisman and Frederic [32]. In this work, they could investigate the products K3NbO4, KNbO3, K4Nb6O17, KNb3O8, and “K6Nb44O113” from the reaction of K2CO3 and Nb2O5 by differential thermal analysis [32,33]. Then, in 1969, the K4Nb6O17 and KNb3O8 crystals were further investigated by Nassau and coauthors [33]. In 1982, a new series of phases MCa2Nb3O10 (M = Li, Na, K, Rb, Cs, NH4, Tl) was reported by Dion et al. [34]. Similarly, some years later, Jacobson and coauthors were able to prepare and investigate the interlayer reactivity of the series of layered compounds K[Ca2Nan−3NbnO3n+1] [35]. Since then, a class of layered perovskite with the general formula Ax[Bm−1NbnO3n+1] is known as Dion–Jacobson compounds (where A represents an alkaline monocation; B an alkaline earth ion; m = 1, 2, and 2 ≤ n ≤ 7; n indicates the number of [NbO6] chains that form each perovskite-like slab) [11]. Moreover, plenty of different niobates are currently known such as alkali (M+NbO3), columbite (M2+Nb2O6), and rare-earths orthoniobates (YNbO4) [6].
These materials are known as promising photocatalysts not just because of their different structures and band gap energies, but also their extended 2D layered arrangement, which provides a considerable surface area and allows easy modification either by intercalation, superficial modification, or formation of nanosheets and nanoscrolls through exfoliation with bulky n-alkylammonium salts [11,31,36]. Recently, our group investigated the influence of the medium, co-catalyst precursor, and heat treatment in the photoactivity of Pt-modified hexaniobate (K4−xHxNb6O17) for H2 evolution [31]. The effect of the Pt(0) precursor and the deposition method (absorption or impregnation) was also investigated. These different conditions of the samples’ preparation mainly affected their configuration and morphology. At the end, the H2 evolution performance was dependent on those conditions, where a ‘soft’ photoreduction of platinum precursors was identified as the best method for the preparation of those photocatalysts.
Suspended hexaniobate layers (K4−xHxNb6O17) in tetrabutylammonium hydroxide (TBAOH) showed a good performance in designing the photocatalytic film through the layer-by-layer (LbL) deposition technique [36]. Thin films were assembled by alternative immersions of FTO substrate into exfoliated hexaniobate with pre-adsorbed [Pt(NH3)4]2+ suspension (pH = 8) and poly(allylamine hydrochloride) solution (pH = 4) up to 25 bilayers. Further thermal treatment at 500 °C removed the organic species, leading to a fuzzy assembly of hexaniobate nanoscrolls with Pt evenly distributed in the surface, as shown in the FE-SEM image (Figure 4). In this case, it could be concluded that the scrolled morphology was favored in relation to the opened sheets. The Nb-based films were photoactive to produce H2 from 20% (v/v) methanol/water solutions under UV-irradiation. The bare hexaniobate LbL films were able to photocatalyze the H2 evolution, with the apparent quantum yield being proportional to the number of deposited bilayers. This behavior evidence that no photocatalytic surface is lost owing to material deposition. Moreover, when the Pt nanoclusters were added to the film composition, the observed H2 evolution rates were approximately two times higher, reaching 4.0% ± 0.5%.
Oshima and coauthors reported the photocatalytic activity of Pt-deposited KCa2Nb3O10 for water splitting [37]. With an adsorption method, cationic precursor of Pt ([Pt(NH3)4]Cl2) was attached on niobate surface (1.3 wt%) by the TBA+/Ca2Nb3O10 suspension, followed by restacking with potassium hydroxide and reduction with H2 gas in various temperatures. The authors noted that the deposition of Pt occurred in the interlayer spaces of the restacked nanosheets. Additionally, the increase in the annealing temperature up to 973 K decreased the Pt content in the KCa2Nb3O10 interlayer space. Diffuse reflectance and X-ray photoelectron spectroscopies suggested that the adsorbed Pt was reduced from an oxidized to the metallic state according to the applied temperature, and that interlayer Pt species were less susceptible to reduction. The samples were evaluated for water splitting in NaI 10 mmol L−1 solution and using a 300 W xenon lamp (λ ≥ 300 nm). The maximum rate of both H2 and O2 production was achieved with the sample treated at 473 K, and higher reduction temperatures led to a decline in the photoactivity. This fact could be explained by an improved contact between Pt and the niobate and the formation of suitable active sites. At higher temperatures, the Pt species were located mostly at the external surface. The authors also investigated the impact of the Pt valence state on water splitting activity, comparing the performance of samples heated at the 573 K under H2 and air atmosphere. By TEM, both methodologies led to the formation of Pt with the same size. However, more electron-deficient Pt species were identified on the sample treated in air, which performed a better overall water splitting reaction than the sample treated at H2.
Distinct species to form the layered Dion–Jacobson materials AB2Nb3O10 also play an essential role in the physical properties of the material and especially in their photocatalytic performance. From this point of view, Kulischow et al. investigated a family of layered Dion–Jacobson perovskite-type materials AB2Nb3O10 (A = K, Rb, Cs and B = Ca, Sr, Ba) and their photocatalytic activity for hydrogen production [38]. Seven different niobate were prepared by the molten salt method mixing BCO3 or B(NO3)2, (B = Ca, Sr, Ba), Nb2O5, A2CO3 with ACl (A = K, Rb, Cs), and at specific reaction temperatures (750–1200 °C). UV/vis diffuse reflectance revealed that the band gap energies varied only with different B-cations, seeming to be independent of A-cations. In the case of Ca2+ as B-cation, the band gap value was 3.6 eV, whereas it was 3.3 eV for Sr2+ and 3.0 eV for Ba2+. This trend suggested a dependence on the ionic radii of the B-cations that filled the spaces among NbO6 octahedral units. With bigger B-cation, the overlap of the Nb 4d orbitals was greater, stabilizing the conduction band energy and, as a result, the band was shifted negatively. The prepared niobates were evaluated for H2 evolution in 10% methanol aqueous solution without a co-catalyst, and then Rh was in situ photodeposited (0.05 wt% Rh loading) on the catalyst. Without co-catalyst, the KCa2Nb3O10 sample exhibited the highest H2 evolution rate; however, after the Rh addition, the highest overall hydrogen production was achieved by CsCa2Nb3O10. In this case, photocatalytic behavior was influenced by the type of the A-cation at the interlayer, suggesting a dependence on the ionic radii in the order Cs > Rb > K, with the exception of RbSr2Nb3O10. The B-cation influenced the hydrogen generation rates owing to more cathodic conduction band edges in the order Ca > Sr > Ba.
Zhou and coauthors showed that the performance of KCa2Nb3O10 for H2 evolution could be enhanced through a nitrogen and tetravalent niobium doping to generate yellow and black [Ca2Nb3O10] nanosheets [39]. After the preparation of KCa2Nb3O10 by solid state reaction, the material was treated at 800 °C for 5 h under NH3 flow, resulting in the black niobate (N/Nb4+-codoped). The previous addition of a certain stoichiometric excess of K2CO3 prevented the formation of oxygen defect, and thus the reduction of Nb5+ ions, which resulted in the yellow niobate (N-doped). Then, the samples were acid-exchanged, followed by the exfoliation reaction with tetrabutylammonium hydroxide (TBAOH). X-ray absorption fine structure (XAFS) data on the Nb K-edge (18.986 keV) of all the samples indicate that doping did not change the perovskite-type crystal structure, as all three samples showed very similar overall profiles. However, an increase of the pre-edge peak (18.985 keV) of black niobate could be attributed to the partial reduction from Nb5+ to Nb4+ and the increased Nb–N bonding for yellow niobate, whereas the black one, showing a simultaneous decrease of both Nb–O and Nb–N, could point out the dominating generation of oxygen vacancies during the ammoniation process without K2CO3 additive. The samples were evaluated for photocatalytic H2 production using 20 vol% methanol aqueous solution under full range irradiation with an Xe lamp. Undoped nanosheets show very low photocatalytic activity (11.4 mmol h−1) compared with the activity of yellow and black niobate nanosheets (42.9 and 154.7 mmol h−1, respectively). In this sense, the self-doping with Nb4+ could significantly promote an enhanced separation efficiency of photogenerated carriers as well as the light absorption beyond single N3− doping. When Pt was loaded at 0.5 wt%, the H2 evolution rates were three times higher, 190.1 and 429.5 mmol h−1 for yellow and black niobate nanosheets, respectively. The last one could achieve energy conversion efficiency for solar hydrogen production of 2.7%.
Later, the same group could also investigate the combination of elemental doping, liquid exfoliation, and composition control for a series of typical Dion–Jacobson phases KCa2Nan−3NbnO3n+1 photocatalysts [40]. As seen in the scheme presented in Figure 5, KCa2Nan−3NbnO3n+1 were prepared with various values of n in order to tune the thickness of the perovskite layer. Theoretically, using n = 3–6, the desired thickness of the layer should increase from 2.4 to 4.0 nm. The materials were obtained from the molar ratio of K/Na/Ca/Nb = 1.05/2/1.05(n−3)/n through calcination at 1200 °C, followed by the N/Nb4+-doping via heat treatment under NH3 flow. This method resulted in samples with different colors, varying from white to black. Samples before and after NH3 treatment were analyzed by UV/vis diffuse reflectance, where red-shifts of the absorption edge were observed in all N/Nb4+ doped samples. Even after liquid exfoliation of bulk doped samples to obtain nanosheets, their suspension in water kept the colors from the bulk materials. TEM and atomic force microscopy (AFM) images revealed ultrathin sheets with several micrometers in size and thicknesses of 2.4, 2.8, 3.3, and 4.0 nm for the n = 3, 4, 5, and 6 homologues, respectively. The photocatalytic performance of the doped nanosheets was evaluated towards hydrogen evolution from 20 vol% methanol solution under full arc irradiation (Xe lamp, 300 W). Considering all samples with and without 0.5 wt% Pt, the n = 4 product exhibited the highest H2 production rate. This fact can be related to the migration length, which affects the electron–hole separation step. The higher photocurrent measurement was also found for n = 4 with a sequence of n = 4 > 5 > 3 > 6, which is consistent of H2 evolution experiments.
The photocatalytic activity of HCa2Nb3O10 for hydrogen production was also extended for visible irradiation region by combination with CdS [41]. In this composite, after light excitation of CdS, the photo-generated electrons were transferred to the calcium niobate nanosheets thanks to their matched conduction band position of −1.32 eV for CdS and −1.18 eV for niobate (V vs. Ag/AgCl, pH = 7.0). So, the observed increase of the electron paramagnetic resonance (EPR) signal after illumination is the result of the efficient electron transfer that further induced the growth of oxygen vacancies. The samples were evaluated for photocatalytic H2 evolution in (0.1 M) SO32−/(0.15 M) S2− solution under visible light, in which the exfoliated niobate was not able to be excited. The best performance was achieved by samples containing 53.6 wt% of CdS with a H2 production rate of 16.5 μmol h−1, about four times higher than pure CdS. Further increase of CdS resulted in lower activity, which was related to the aggregation of CdS nanoparticles. Later, a similar approach was applied by Hu and coauthors to design nanohybrid of CdS and HCa2Nb3O10 and get photocatalytic H2 production under visible light [42]. It has been shown that CdS has more negative conduction band energy than the employed niobate, indicating the spontaneous transfer of photoinduced electrons from CdS to niobate nanosheets.
Xiong and coauthors could improve the photocatalytic hydrogen production activity of HNb3O8 nanosheets by loading Cu with a facile photodeposition method [43]. Exfoliated HNb3O8 niobate was dispersed in Cu2+ solutions with corresponding loadings of 2%, 1%, 0.5%, and 0.25% which was photodeposited in situ in triethanolamine solution. Firstly, the authors evaluated the samples for H2 evolution activity with 10 vol% of triethanolamine aqueous solution under the simulated solar light using a 300 W Xenon lamp. The photocatalytic performance increased after addition of Cu2+ up to 0.5%, reaching the maximum rate of 59.1 µmol h−1. With higher amounts of Cu, the shielding effect was detrimental for the photoactivity. In addition, when methanol was used as the sacrificial agent, the photocatalytic hydrogen production activity achieved the value of 98.2 µmol h−1. The authors argued that in situ photodeposition induced selective deposition on the photocatalytic reduction sites of HNb3O8 nanosheets rather than arbitrary distribution.
The same type of niobate was recently applied to engineer a stable nanocomposite photocatalyst. Xia and coauthors reported a method to growth NiS in a highly dispersive way on the HNb3O8 surface [44]. In general, the random deposition and easy agglomeration in large particles of NiS are described as being problematic for the photocatalytic H2 evolution. Therefore, an electrostatic adsorption/self-assembly methodology was developed to obtain NiS on the HNb3O8 nanosheet as an efficient photocatalyst for the first time. The photocatalytic H2 evolution was performed with a 300 W Xenon lamp in a 10% triethanolamine (TEOA) aqueous solution, where the bare NiS and layered KNb3O8 showed no or very low H2 production. In this condition, for the composite with 1 wt% NiS, the H2 evolution rate was 1519.4 µmol g−1 h−1, higher than the physical mixture of the components (582.5 μmol g−1 h−1). Time-resolved fluorescence decay analysis was applied to calculate the lifetime of the photogenerated electrons and holes. The kinetic data showed a slower decay for the composite, which indicates that the applied methodology resulted in more efficient utilization of photogenerated electrons and holes.
The potential of niobates structures for water splitting has also been demonstrated by some theoretical works. The perovskite-type niobate NaNbO3 was selected by Wang and coauthors to investigate the effect of anionic monodoping with N, C, P, and S dopants, as well as with (N + N), (C + S), and (N + P) codoping pairs by hybrid density functional theory calculations [45]. At first, the direct band gap of pure cubic NaNbO3 was predicted to be 3.30 eV, with the conduction band mainly formed by Nb-4d orbitals and valence band by the O-2p orbitals, and almost no Na-related states around the band edge were found, indicating that the Na atoms have negligible effects on the electronic structures near the Fermi level. In this sense, one of the O atoms in the cell was firstly substituted by one of four different species (N, C, P, and S), corresponding to a doping concentration of 2.5%. In the case of N-doped NaNbO3, a hole would be generated as the N dopant has less valence electrons than the O atom. The higher energy of N-2p states (around 2.0 eV) induced a localized unoccupied impurity state above the Fermi level as well as several occupied impurity states around the valence band of NaNbO3. For C doping, the substitution would result in two holes in the system, however, no unoccupied localized impurity states were found above the Fermi level. As a result, the total magnetic moment was higher and there were several impurity states above the valence band, which shifted up by about 1.58 eV. The doping with P resulted in a spin-polarized ground state as well, and, owing to the P-3p higher energy than the O-2p, several empty impurity states were found near the conductive band. When S was applied as a dopant, no empty states were found within the band gap, but several localized impurity states were localized below the Fermi level, right above the valence band of the niobate, which resulted in a narrowed band gap. This filled impurity states originated from the hybridization of S-3p and O-2p orbitals. Thus, the band gap values for N-, C-, P-, and S-doped NaNbO3 were calculated to be 1.88, 1.61, 1.04, and 2.34 eV, respectively. Despite the narrower band gap energies, some disadvantages were observed as the undesirable empty gap states from N and P monodoping, which may act as trapping and recombination centers, and the elevated valence band of NaNbO3 after C doping, precluding the oxygen evolution reaction. In this sense, the codoping possibility using pairs of (N + N), (C + S), and (N + P) was also investigated. It resulted in the band values as shown in Figure 6. In comparison with the pristine niobate, for (N + N) codoping, two fully filled states were generated within the band gap and its effective energy was 1.24 eV lower. The (C + S) dopping led to mid-gap states above the valence band of host system, and then a narrow band gap of 1.42 eV. For both systems, the band edges matched properly with the redox potentials of water, indicating that the photocatalyst would be suitable for overall photocatalytic water splitting. However, in the case of (C + S) codoping, the conductive band was shown to no be longer good for a spontaneous water splitting process even with a band gap value of about 1.33 eV. This study suggested that codoping can be a successful approach to improve the visible light photocatalytic performance of perovskite NaNbO3.
Kaneko and coauthors deeply investigated the origin of the visible light absorption and the optical band gap of 1.9 eV in the d1 metallic strontium niobate (SrNbO3) [46]. By computational methods, firstly, they could verify at which electronic states the photoexcitation takes place, from CB → B1 transition or B−1 → CB (B−1 denotes the band below the conductive band (CB), and B1 denotes the band above the conductive band). By the band structures and projected density of states calculations, it was found that B−1 state lies at around −4 eV and consists of O(p). Thus, from the atomic orbitals constituting each band, it was inferred that the optical gap corresponds to the Nb(d) → Sr(d)/Nb(d) transition, and so the optical gap was attributed to the CB → B1 transition. The shifts of Fermi level as well as the changes in the optical gap concerning the Sr defects and O vacancies were investigated. The authors could conclude that the optical gap hardly depends on the amount of Sr defects. Sr and O are divalent cation and divalent anion, respectively, and the valence of the Nb ion constituting CB did not change even if the Sr defect and the O vacancy were produced in the same amount. Thus, if the valence of the Nb ions constituting CB does not change, the Fermi level will not shift either. Consequently, the optical gap, which is mainly caused by the excitation from the Fermi level to B1, does not change by Sr1−xNbO3−x. In addition, in the case of Sr0.875NbO3 and Sr0.875NbO2.875, it was found that the latter evidenced the dependence of the optical gap on the amount of Sr defects and its light absorption intensity was larger than that of the former.
Such doped niobates have been prepared and evaluated as photocatalysts. Yu et al. reported at first time the performance of carbon-doped KNbO3 as well as the effect of MoS2 addition in the material for photocatalytic H2 evolution [47]. Both pristine and C-doped KNbO3 presented the orthorhombic phase, however, the latter resulted in peaks of 30–32° shifted to a lower angle, indicating that the carbon atom was doped into the lattice of KNbO3. In addition, the C species were identified as C4+ by X ray photoelectron spectroscopy (XPS) analysis, which meant that it did not replace the position of lattice oxygen. In addition, it was also confirmed that the substitution of Nb5+ by C4+ occurred and also generated some oxygen vacancies to compensate the charge balance. The pristine niobate showed only UV light absorption with band gap of 3.07 eV. C-doped niobate presented very similar behavior, however, it showed a difference in the peak tailing in the range of 400–600 nm. The samples after the addition of MoS2 in 0.2 wt% also presented the peak tailing in the visible-light region. H2 evolution was performed in methanol solution (20% v/v) under 300 W Xe lamp irradiation and in situ photodeposition of metallic Pt (0.37 wt%). The C doping of KNbO3 could enhance the H2 evolution from 5 µmol g−1 h−1 to 142 µmol g−1 h−1 in comparison with the pristine sample. The decoration of MoS2 on C-KNbO3 could further increase the rate of H2 evolution to 1300 µmol g−1 h−1. The samples were also evaluated under visible light, where pure niobate did not show activity, while C-KNbO3 and MoS2/C-KNbO3 resulted in rates of 4.2 µmol g−1 h−1 and 9.3 µmol g−1 h−1, respectively. These activities were much lower compared with values under UV/vis light, indicating that the composite is more suitable for working in the UV region of irradiation.
Columbites are also founded in the literature with further applications in photocatalytic research. The effect of structural distortion of SnNbxOy in comparison with the crystalline SnNb2O6 regarding the photocatalytic activity was investigated by Huang et al. [48]. The crystalline sample exhibits a band gap energy of 2.10 eV, while the band gap energies for disordered and amorphous samples were 2.33 eV and 2.43 eV, respectively. Furthermore, the electronic band gap could be tuned by the structural disorder engineering, as an increase of structural lattice disorder led to an increase of band gap energies. By XPS and UV/vis absorption analysis, the valence band edge of crystalline SnNb2O6 was about 1.42 V with the conduction band of −0.68 V, whereas for the disordered sample, the values were 1.98 V and −0.35 V, respectively. As a consequence, the latter exhibited H2 evolution rates 11 times and 1.7 times higher than that of crystalline and amorphous samples, respectively, under visible light irradiation (λ ≥ 420 nm). The photoinduced charge carrier behaviors under pulse laser at 532 nm were investigated by transient absorption spectroscopy (TAS). The three samples presented broad and continuous absorption in the visible range at a delay time of 100 ns, indicating the separation of charge carries at different trap states. This behavior was especially pronounced in disordered SbNbxOy, which indicated a better charge separation. The transient absorption decay kinetics also expressed the same, with effective lifetimes for the amorphous, disordered, and crystalline samples of 0.52 ms, 0.56 ms, and 0.26 ms, respectively.
The influence of type of polymorph on the photocatalytic activity of CuNb2O6 columbite was further explored by Kamimura et al. [49]. Two different polymorphs, monoclinic and orthorhombic, revealed that the local crystal structure of the CuO6 octahedral was strongly correlated to the optical absorption property. Both samples exhibited the same band gap, 2.7 eV. However, two additional broad bands in the near-infrared (NIR) region were observed for the orthorhombic sample, being ascribed to Cu2+ d–d transitions in a distorted octahedral ligand field. In principle, this transition is forbidden by the Laporte selection rule owing to the symmetric CuO6, but in the case of the orthorhombic phase, it was partially allowed because of the Jahn–Teller effects of orthorhombic CuNb2O6. The photocatalytic H2 evolution was performed using a 10 vol% methanol solution and an Air Mass (AM) 1.5G solar light irradiation. The two polymorphs of CuNb2O6 produced H2 under these conditions, but the monoclinic sample was more active than the orthorhombic one. Thus, to elucidate if the d–d absorption affects the photocatalytic activity, the previous results were compared to those using a near-infrared (NIR) cutoff filter. When the monoclinic sample was tested, the rate of H2 evolution rate did not change expressively when the NIR filter was applied. In this case, the loss of 20% of activity could be explained by the low transmittance of the employed cutoff filter in the visible region. In contrast, the active for orthorhombic CuNb2O6 decreased 90% in relation to that observed without an NIR filter. On contrary to the monoclinic phase, which was insensitive to the excitation of the d–d transition in Cu2+ ions, for the orthorhombic one, the electron 3d9 configuration of Cu2+ readily induces structural distortion by Jahn–Teller effects in an octahedral crystal field. The empty 3d orbitals of Cu2+ could work as deep traps for the photo-generated electrons, and thus increase the recombination rate, resulting in a significant decrease in photocatalytic activity in the presence of an NIR cut-off filter. With NIR photons, the photo-generated electrons filled 3d orbitals of Cu2+ via d–d absorption, a state that inhibited the recombination of photogenerated electrons in the conduction band. These distinct mechanisms were schematically represented on Figure 7. So, in the case of orthorhombic structure, the photocatalytic performance was dependent on fully filled Cu 3d-orbitals.
Chun and coauthors employed a two-step hydrothermal process to form nanosized ZnNb2O6 columbite [50]. The samples were applied to photocatalytic H2 evolution in 20 vol% methanol under irradiation of 500 W high-pressure mercury lamp. The sample prepared by the new method produced an optimal activity of 23.6 µmol H2 h−1 g−1, while a bulk ZnNb2O6 prepared by solid state exhibited a H2 evolution rate of 9 µmol h−1 g−1. This fact was attributed to its appropriate crystallinity and high specific surface area (61 m2 g−1). Further, the authors evaluated the effect co-catalyst deposition on nano-ZnNb2O6. In this case, both pristine water and methanol aqueous solution media were applied. In pure water, the bare photocatalyst showed an undetectable amount of H2. The optimal co-catalyst was found to be Pt, as the H2 generation rate achieved up to 680 and 3200 µmol h−1 g−1 in pure water and methanol aqueous solution, respectively. Moreover, with the modified sample and at optimal conditions, a maximal apparent quantum yield (AQY) value of 4.54% was obtained in pure water. The value increases to 9.25% with the assistance of methanol as the sacrificial agent.
The manipulation of composition, morphological, and surface properties of three-dimensional hierarchical Nb3O7(OH) superstructures was achieved through Ti(IV) incorporation, as reported by Betzler and coauthors [51]. Firstly, the undoped Nb3O7(OH) consisted of blocks of corner-sharing NbO6 octahedra, forming hollow cubic superstructures of nanowire networks (Figure 8a). When an average amount of 5.5% titanium was incorporated into the crystal lattice, it formed a spherical morphology, hollow and built up from nanowires as well (Figure 8b). With a higher amount of Ti(IV) (10.8%–31.2%), a cubic morphology with smaller nanowires arranged to form the walls of hollow cubes was observed (Figure 8c). All the samples presented the same nanowire arrangement; however, it became smaller, shorter, and flatter as the quantity of Ti was increased. Both XRD and energy-dispersive X ray (EDX) measurements indicated a homogeneous distribution of titanium, however, with more than 12% Ti, the crystal lattice of Nb3O7(OH) was not able to accommodate it, and then the formation of niobium-doped anatase TiO2 plates was favored for further titanium excess. The specific surface area was also affected and it changed from 79 m2 g−1 for Nb3O7(OH) (morphology a) to 132 m2 g−1 and 173 m2 g−1 for morphologies b and c, respectively. The three samples showed the same band gap energy of 3.2 eV, but an additional absorbance in the 550−850 nm region was observed, which came from oxygen vacancies in the TiO2 plates. Further, the samples were evaluated to H2 evolution under an Xe lamp irradiation (600 mW cm−2), with 10 vol% methanol and in situ photodeposited Pt at 8 wt%. The Nb3O7(OH) resulted a rate of 870 μmol g−1 h−1, and then, as the Ti concentration increases, the H2 rates zre raised to 1773 μmol g−1 h−1 for morphology b and 1988 μmol g−1 h−1 for morphology c. The same trend was also expressed by transient absorption measurements, where morphology a showed an averaged lifetime of 54 ± 2 ps, while morphologies B and C presented lifetimes of 64 ± 2 ps and 73 ± 4 ps, respectively. The authors explained this behavior by the fewer amounts of hydroxyl groups required for charge neutrality with the replacement of Nb5+ by Ti4+, as they are known for acting as non-radiative recombination sites. The doubled H2 production rate of morphology b, compared with morphology a, was most likely the result of a combination of its fewer recombination sites and larger surface area. The further, but less pronounced increase of the H2 production rate for morphology c came from the presence of TiO2 plates, which did not have such a strong effect on the photocatalytic reaction.
It is clearly noted that niobates offer the possibility to be combined with innumerous co-catalysts. Previous cited works and other examples from the literature are summarized in Table 1.
Besides the previous cited materials, niobium is also flexible in terms of designing different layered compounds with distinct elements. Fujito and coauthors reported a visible light response of layered perovskite oxychloride Bi4NbO8Cl for water splitting [58]. This material consists of single-layer NbO4 perovskite units separated by (Bi2O2)2Cl blocks. It presented a band gap of 2.39 eV with valence and conductive bands of 2.11 eV and −0.28 eV, respectively. Experimental and theoretical results suggested that the valence band was formed by highly dispersive O-2p orbitals coming from the interactions within and between Bi−O and Nb−O slabs. As expected from the conductive band value, Bi4NbO8Cl showed activity for H2 evolution from an aqueous methanol solution under UV irradiation (300 W Xe lamp) and Pt (0.5 wt%) loading. However, the obtained rate of ~0.1 μmol h−1 was considered low and a higher performance for O2 was found. Apparent quantum efficiency for O2 evolution was determined as ~0.4% with FeCl3 as an electron acceptor, under monochromatic light at 420 nm (~25 mW/cm2). So, through a Z-scheme mechanism, the system was coupled to a H2-evolving photocatalyst of Rh-doped SrTiO3 and Fe3+/Fe2+. In this way, as in the schematic shown in Figure 9, a simultaneous evolution of H2 and O2 under visible light was successfully observed.
Wakayama and coauthors reported a novel compound obtained from a mixture of layered RbNdNb2O7 and Rb2CO3 [59]. The new layered niobium oxynitride, Rb2NdNb2O6N, resulted in a structure as Rb2NdNb2O6N H2O composed by double-layer [NdNb2O6N]2− perovskite slabs separated by two Rb cations and one H2O molecule. The material showed a band gap of 2.5 eV with visible light absorption, whereas the precursor RbNdNb2O7 essentially absorbs UV light with a band gap energy of 3.7 eV. In the former case, additional N 2p orbitals in the valence-band formation caused a valence-band maximum that was much more negative. In contrast, as the conduction band consists mainly of Nb 4d orbitals, its value remained basically unaffected. Thus, the H2 evolution was evaluated in the presence of TEOA as an electron donor in dimethyl sulfoxide (DMSO) containing 1 mL of water. The sample was irradiated with visible light (λ > 400 nm) by 300 W xenon lamp, and Pt was previously deposited at 0.5 wt%. As shown in Figure 10, a stable H2 evolution was observed for 20 h from the Rb2NdNb2O6N sample, in contrast to the precursor, which did not show a response under visible light irradiation.
Therefore, it can be observed that niobium layered compounds form a versatile Nb-based materials class that exhibits great activity for H2 evolution reaction. Additionally, several procedures to tuning the photocatalytic performance niobium layered compounds were discussed, such as changes in the structural and morphology, doping and co-doping, and combination with other semiconductors.

2.3. Nb-Doped Materials

Two of the main drawbacks in the photocatalysis heterogeneous are the rapid recombination of electron/hole pair and semiconductor activation only under UV irradiation. A strategy to increase the photocatalytic performance of semiconductors by overcoming these challenges is the introduction of impurities able to change the electronic behavior of the material, in one word, the doping. Niobium is well known in the literature as a dopant agent, especially for titanium-based semiconductors, as its higher charged ion can substitute the Ti4+ site in the oxide structure, and simultaneously play the role of donor, improving carrier concentration and conductivity [60]. For TiO2, the defect state provided by the doping is located upon the conduction band minimum and contributes electrons to the unoccupied Ti 3d orbital without introducing additional states in the bandgap. However, in order to satisfy the charge compensation, the substitution can be accompanied by the incorporation of other defects as oxygen vacancies or even Ti3+ defects below the conductive band, which lead to a visible range absorption [61]. In Table 2, examples of Nb-doped semiconductors for H2 evolution are listed.
The performance of the Nb-TiO2 photocatalyst was evaluated using various sacrificial agents such as methanol, ethanol, and 2-propanol by Fontelles-Carceller [68]. The photocatalyst was prepared with an Nb/Ti ratio of 0.025:0.975, which was confirmed by XPS measurements. In principle, the presence of Nb does not significantly change the band gap (3.15 eV) or the specific surface area (80.5 m2 g−1) of the photocatalyst in comparison with pure TiO2 (3.18 eV of band gap and 82.1 m2 g−1 of surface area). The H2 evolution was investigated by the photocatalysts loaded with 0.5 wt% of Pt in presence of a 3:7 (v/v) alcohol/water mixture under irradiation of an Hg-Xe lamp (500 W). For both TiO2 and Nb–TiO2, the maximum rates were observed for methanol, followed by ethanol and 2-propanol. In all tested conditions, as shown in Figure 11, the doped sample showed a significantly better result than the non-doped sample. Both samples presented maximum rates for methanol, followed by ethanol and 2-propanol. In addition, the products formed during the photocatalytic reaction were investigated and, for methanol and ethanol, basically the same compounds were identified. Both samples generated formic acid and metyl formate by methanol oxidation with the addition of formaldehyde for the Nb-doped sample under UV illumination. When ethanol was applied, in all conditions of irradiation and applied samples, acetaldehyde and ethyl acetate were found. Nevertheless, when 2-propanol was applied as a sacrificial reagent, the Nb-doped sample generated acetone and propanone–diisopropyl–acetal, whereas bare TiO2 produced only acetone. So, it was possible to affirm that the presence of Niobium at the surface even in a small proportion was able to generate a new compound not observed for the pristine sample. In this sense, in situ infrared studies were also carried out under dark and illumination conditions and in the presence of the alcohol in water vapor. In principle, under the dark, it was possible to verify the adsorption of the alcohols on the surface of both samples. Under UV irradiation, the infrared spectra showed that alcohol and alcoxy species were consumed and/or desorbed from the surface of the samples. For methanol and ethanol, carboxylate/carbonate species were found as the main surface product. This could indicate the attack from hole to the alcohol to form oxidized species related to the corresponding alcohol/alcoxy adsorbed species. In the case of 2-propanol, carboxylate species were less present at the surface owing to the low coverage of products from the relatively easy desorption of acetone. Regarding the H2 evolution, the authors suggested that the differences in activity are correlated quantitatively with the average oxidation potential of the alcohol. When comparing methanol with ethanol using infrared spectroscopy, they showed essentially the same behavior of the titania surfaces against the alcohol and evolving molecules coming from the hole-attack.
Different materials that were doped by niobium ions to make them able to split water have been reported in the literature. The improvement of WO3 for water splitting by photoelectrochemical process was achieved by Nb-doping [69]. The morphology of the material turned from nanorod to nanotriangle, which also increased the porosity of the WO3 films. Also, the band gap increased to 2.74 eV by Nb-doping in comparison with WO3 (2.63 eV) and additional oxygen vacancies were induced. Besides that, the carrier densities significantly increased from ~3.44 × 1019 cm−3 to 1.27 × 1022 cm−3, indicating that Nb5+ ions act as donor dopants in the WO3 lattice. It could improve the charge separation and electron transport, and then the doped sample exhibited higher photocurrent, incident photon to charge carrier efficiency (IPCE) (39% to 52% at 300 nm and 1.23 V vs. RHE), and photoconversion efficiencies.
Also, with the objective to promote photoelectrochemical water splitting, niobium could support the activity of BiVO4 films [67]. In this case, the authors concluded that Nb5+ ions were not placed in the crystalline structure of BiVO4, rather as niobium oxide or oxochloride. Firstly, the addition of 10% of Nb(V) resulted a higher IPCE, which reached 75% in comparison with pristine BiVO4 of 55% at 2.08 V versus RHE (380–550 nm). The H2 evolution experiments were run by an H-shaped cell with pristine or modified BiVO4 film as a photoanode coupled with a cathode of Pt and carbon paste. After 16 h of irradiation at 2.08 V, the pristine sample reached a maximum H2 evolution rate of 0.09 mmol h−1, whereas Nb-modified BiVO4 increased about two times to 0.18 mmol h−1. The authors argued that the higher performance of the modified sample would be related to the modification of the nanostructure of the film, but also owing to other effects not obvious by the obtained data. A high efficiency for the photoelectrochemical water splitting reaction was achieved by doping of SnO2 nanotubes (NTs) with Nb and N co-doping [70]. Niobium could be ascribe as an effective dopant owing to its abundant electronic states and lower ionic radii of Nb4+ (69 pm) compared with Sn4+ (71 pm), resulting in minimum lattice strain. As shown in Figure 12a, the addition of Nb could improve the absorption of light at lower wavelengths, suggesting a decrease in the band gap of the doped samples, with an increase in Nb content. Then, a further doping with N could result in a significant decrease of band gap to 1.99 eV from 3.1 eV of (Sn0.95Nb0.05)O2. The lowering of the optical band gap for this sample could provide an IPCE of 10% at 500 nm and potential of ~0 V (vs. RHE), which was described by the author as the highest IPCE value obtained for semiconductor materials explored so far for photoelectrochemical (PEC) water splitting. The generated H2 and O2 gases at the cathode (Pt wire) and the photoanode are shown in Figure 12b. The theoretical concentration of H2 from Faraday’s Law was also calculated, which was very similar to the experimental one, suggesting a Faradaic efficiency close to 100%. Besides that, the ratio of produced gases was around 2, showing a stoichiometric decomposition of water into H2 and O2. From these as well as the theoretical results, it was possible to conclude that the Nb and N co-doping in SnO2 afford better light absorption properties, increased carrier density, and facile electrochemical charge transfer.

3. Photoreduction of CO2

Artificial Photosynthesis

Carbon dioxide (CO2) emission from fossil fuels is the main concern when discussing global warming and related climate changes. Because the global energy demand will continue to increase, the consumption of fossil fuels will continue to be high for the next few decades, which undoubtedly brings the necessity of large-scale CO2 mitigation from the atmosphere [71]. The observation of natural photosynthesis in green plants motivated the development of photocatalytic CO2 reduction (artificial photosynthesis), which is one promising solution to transform the current fossil fuel-based economy into CO2-neutral energy systems using the sustainable “photon” economy and mitigate the greenhouse gas effect [3,72,73,74]. Another advantage of the photocatalytic CO2 reduction is the indirect solar energy storage on carbon-based molecules. Therefore, the artificial photosynthesis acts as killing two birds with one stone in view of protecting the environment and simultaneously supplying renewable energy. However, CO2 is very stable/inert regarding chemical reactions, meaning that CO2 conversion demands a considerable amount of energy; therefore, it is quite difficult to convert it to reusable chemicals [75].
In the beginning of the 1900s, Giacomo Ciamician published a seminal paper with the first idea about the artificial photosynthesis, with the suggestion that society should shift from consuming fossil fuels to generating sustainable energy from the sun [76]. Later, Inoue et al. reported the photoelectrocatalytic reduction of CO2 to form organic compounds such as formic acid, formaldehyde, methyl alcohol, and methane. In the presence of various semiconductors, the dependence between the formed products and the semiconductor conduction band energy was demonstrated [77].
The CO2 photoreduction process is a multielectron process (2, 4, 6, or 8 electrons) that can yield diverse products and occurs through different steps that involve adsorption, activation, and dissociation of the C=O bond. The adsorption and activation step is especially challenging because CO2 is highly stable and inert [78]. In Table 3, some of the possible CO2 reduction reactions are summarized along with their respective redox potentials. Selected potentials are compared with Nb-based photocatalysts at Figure 13. The formation of first intermediate (CO2•−) state imposes a significant energy to the reaction and is frequently identified as the rate determining step. Once CO2•− is formed, it may be subsequently reduced via the protonation of its oxygen atom, resulting in the formation of COOH. This intermediate can be reduced to CO and released from the semiconductor surface. Alternatively, CO2•− may also be reduced via the protonation of its carbon atom to form HCOO, which is further reduced to formate (HCOO) [79,80,81]. Therefore, the CO2 reduction mechanism is very complex and imposes kinetic and thermodynamic limitations.
The semiconductors that have been investigated for CO2 conversion present low activity, selectivity, stability, and high band gap values; therefore, challenges remain on how to make this process feasible. The CO2 reduction reaction competes with H2 evolution in protic media, thus efficient catalysts must exhibit high activity for CO2 conversion with low or negligible activity for H2 evolution. This requires a very high activation energy barrier for H2 evolution or a change in the reaction pathway to slow down the kinetics of H2 formation [82].
Few semiconductors have been identified with high surface affinity to CO2 and, simultaneously, with suitable electronic properties for CO2 reduction via a photo-activated mechanism. In this sense, Nb-based materials are promising for application in CO2 photoreduction owing to their properties, such as high acidity, excellent textural properties, and interesting electronic properties. Additionally, it is well-known that the product selectivity is significantly affected by the nanostructure design and the architecture of the photoactive components [83].
Efficient CO2 reduction involves the optimization of synthetic procedures and the development of different architectures in order to achieve the structural, electronic, morphological, and surface properties required for this application. Despite that the application of Nb-based semiconductors for CO2 photocatalytic reduction has only recently been explored, some reports have revealed the potential of these structures for artificial photosynthesis. It can be highlighted that some materials, such as alkali niobates (NaNbO3 and KNbO3), columbite niobates (CuNb2O6 and ZnNb2O6), niobium oxide (Nb2O5), and niobic acid (HNb3O8). Additionally, Nb-based materials have been involved in some strategies to improve the CO2 photocatalytic reduction activity and selectivity, such as doping and formation of composites
Shi et al. [84] evaluated the CO2 photocatalytic performance of NaNbO3 with different morphologies for the first time. NaNbO3 in orthorhombic phase were prepared by a conventional solid-state reaction and hydrothermal method, with micron-sized particles and homogeneous nanowires, respectively. It was observed that NaNbO3 decorated with Pt was efficient to convert CO2 in CH4, while there was almost no CH4 to be detected over pristine NaNbO3. Additionally, the CO2 reduction performance of Pt–NaNbO3 nanowires was higher than the micron-sized particles, owing to the higher specific surface area of Pt–NaNbO3 nanowires.
It is well-known that the photocatalytic reactions (oxidative or reductive processes) are particularly affected by the semiconductor crystal structure [85]. Therefore, the CO2 photocatalytic reduction performance of the cubic and orthorhombic NaNbO3 was evaluated. The samples were obtained using inorganic and organic Nb-precursors by a furfural alcohol-derived polymerization–oxidation [85]. Both samples exhibited the same morphology features, that is, cuboids particles with a size between 30 and 50 nm, likely owing to the use of poloxamers with surfactant behavior (Figure 14). The cubic Pt–NaNbO3 presented lower band gap energy than the orthorhombic Pt–NaNbO3. Cubic NaNbO3 exhibited a better photocatalytic performance for H2 evolution and CO2 reduction for CH4 than orthorhombic NaNbO3; the authors attributed this behavior to its unique electronic structure, which benefits the generation and migration of photo-generated electrons and holes. This behavior was confirmed for the same group in another publication [86]. The authors also investigated the effect of annealing temperature in the CO2 reduction performance, and a deleterious effect with the increase in the temperature was observed.
On the basis of previous reports, Fresno et al. evaluated the CO2 photoreduction performance of NaNbO3 obtained by solid state reaction at 900 °C and compared it with NaTaO3 photoactivity [87]. Owing to the high annealing temperature, orthorhombic NaNbO3 with low surface area was obtained (~1 m2·g−1). The main CO2 reduction products were CO, CH4, and CH3OH, together with H2 coming from the concomitant reduction of water competing with that of CO2. It was observed that NaNbO3 and NaTaO3 give rise to similar conversions in the CO2 reduction reaction. Additionally, they demonstrated that NaNbO3 was photoactive even without a co-catalyst, although the main product was CO; the authors attributed this difference in the selectivity to the absence of a co-catalyst (Pt) [87].
Inspired in the Degussa P25, which consists of anatase and rutile TiO2 (i.e., heterostructure) and shows higher activities than either pure anatase or rutile in water splitting and gaseous pollutant photodegradation [88], Ye’s group proposed the investigation of heterojunction formation between NaNbO3 in cubic and orthorhombic phases for the CO2 photocatalytic reduction [89]. The NaNbO3 samples were synthesized at 400 °C to 600 °C based on the polymerized complex method. Pure cubic and orthorhombic NaNbO3 were obtained at 400 and 600 °C, respectively, while the NaNbO3 samples with mixed phases were formed at the temperature ranging from 400 to 600 °C (Figure 15b,c). Along with the phase transition from the cubic phase to the orthorhombic phase, the absorption edge of NaNbO3 shows a slight and gradual blue shift, which is consistent with the previous report that cubic NaNbO3 has a narrower band gap than orthorhombic NaNbO3 (Figure 14c). Additionally, the presence of heterojunctions between the two phases was verified; the smaller ones less than 10 nm (cubic phase) and the bigger ones about 30 nm (orthorhombic). For the CO2 photocatalytic reduction experiment, Pt was used as a co-catalyst and CH4 was the main product. All the mixed-phase samples were more efficient in CH4 formation than the pure phases. The authors state that the photo-excited electrons migrate from the conduction band of the orthorhombic phase to the trapping sites on the cubic surface, thus avoiding the electron–hole recombination in orthorhombic NaNbO3 and improving the charge separation efficiency in the mixed-phase NaNbO3 samples.
Niobates materials are also efficient photocatalysts for CO2 photoreduction, because a layered materials configuration is favorable for the separation and transportation of photo-excited carriers (i.e., electrons and holes) [90,91]. In this sense, niobic acid (HNb3O8) is a promising material, as it exhibited the valence band and the conduction band edges at desirable potential levels. The layered structure facilitates electron transfer, and the protonic acidity is favorable for water adsorption through hydrogen bonding [91]. It is expected that these properties should exert a certain positive impact on CO2 photoreduction by water [91].
Therefore, Li et al. [92] evaluated the photocatalytic CO2 reduction performance of HNb3O8 obtained through two different methods, hydrothermal (HT) and solid state reaction (SSR), and compared it with that of KNb3O8. HNb3O8 is isostructural with KNb3O8, and both samples were crystallized in an orthorhombic symmetry. The samples obtained through the HT method exhibited thin nanobelts with several micrometers in length, while the samples obtained by the SSR method showed containing irregular particles in micron size. The surface area value increased from 2.7 m2 g−1 of KNb3O8 (SSR) to 28.8 m2 g−1 of KNb3O8 (HT), and from 6.5 m2 g−1 of HNb3O8 (SSR) to 39.4 m2 g−1 of HNb3O8 (HT). CH4 was the only one hydrocarbon product detected in the gas phase during CO2 photoreduction for all samples. The rate of CH4 formation over KNb3O8 (HT) was seventeen times higher than that over KNb3O8 (SSR), and the rate of CH4 formation over HNb3O8 (HT) was about twenty times higher than that over HNb3O8 (SSR). Additionally, the solid acid sample (HNb3O8) exhibited higher activity than the corresponding potassium salt (KNb3O8). Also, both samples exhibited higher activity than degussa TiO2–P25. Therefore, the sample morphology, specific surface area, and protonic acidity of HNb3O8 play important roles in the photocatalytic reduction of CO2 to CH4.
In order to improve the CO2 photocatalytic reduction performance of HNb3O8, the intercalation of guest components into the interlayer space was evaluated, and lamellar HNb3O8 was purposely pillared with silica [93]. The pillarization of HNb3O8 with SiO2 was verified by interlayer distance expansion, from 1.13 nm to 2.85 nm. Owing to the notably expanded interlayer distance, the surface area value of SiO2-HNb3O8 was as large as 197.3 m2 g−1, in contrast to 6.5 m2 g−1 of the nonpillared HNb3O8. All of the samples were loaded with 0.4 wt% Pt by photodeposition to evaluate the CO2 photoreduction. The yield of CH4 over the HNb3O8 sample was almost three times higher than that achieved over Nb2O5. The improved activity for the solid acid material might be ascribed to its layered structure and the stronger adsorption ability to H2O through hydrogen bonding. The rate of CH4 formation was further enhanced with the SiO2-HNb3O8 sample. Additionally, the activity of the current SiO2-HNb3O8 photocatalyst was higher than that of previously reported HNb3O8 nanobelts [92]. Compared with nonpillared HNb3O8, the SiO2 pillared HNb3O8 sample has a notably expanded interlayer distance and much greater surface area. Thus, the reactive active sites at the interlayer space of SiO2-HNb3O8 should be more accessible to the substrate. To investigate the role of protonic acidity for CO2 reduction, the activity of SiO2-HNb3O8 was compared with that of SiO2-KNb3O8 and anatase TiO2 in different molar ratios of H2O/CO2 (0.06 or 0.25). The SiO2-HNb3O8 sample showed higher activity than SiO2-KNb3O8 and anatase TiO2; it is noteworthy that the activity of the SiO2-HNb3O8 sample was more significantly enhanced than the other two samples at an elevated water content. This confirms that water molecules could react easily with the protons at the interlayer space of lamellar solid acids through hydrogen bonding.
Despite columbite niobates exhibiting interesting electronic properties and suitable conduction and valence band energies for CO2 photoreduction [6], this class of material has been explored only in a limited manner. Kormányos et al. evaluated the CO2 photoelectrochemical reduction activity of monoclinic type-p CuNb2O6 and orthorhombic type-n ZnNb2O6 semiconductors [94]. The CuNb2O6 and ZnNb2O6 samples showed an indirect bandgap of 1.77 eV and 3.55 eV, respectively. The photoelectrochemical (PEC) performance of the p-CuNb2O6 sample was tested in CO2- and N2-saturated 0.1 M NaHCO3 solution. The photocurrents are two times higher (at E = −0.4 V) when the solution was saturated with CO2. According to the authors, the dramatic increase in the presence of CO2 is clearly attributable to the ability of CuNb2O6 to photoelectrochemically reduce CO2. However, the CuNb2O6 sample’s efficiency and selectivity was not evaluated, as the products formed from CO2 photoelectrochemical reduction were not monitored.
Recently, da Silva et al. investigated the role of Nb2O5 surface acidity for CO2 reduction performance [95]. Nb2O5 catalysts were prepared through a modified peroxide sol–gel method using different annealing temperatures, and it was verified that the increase in the annealing temperature decreases the surface acidity. The authors related the activity and selectivity of the Nb2O5 samples to their surface acidity; high surface acidity prompted the conversion of CO2 to CO, HCOOH, and CH3COOH, whereas low surface acidity led to the conversion of CO2 to CH4. It was verified that CO is the main intermediary during the CO2 photoreduction in all conditions. This work unveiled the importance of surface acidity for CO2 photoreduction.
Despite the remarkable properties of NaNbO3 and interesting performance for CO2 photocatalytic reduction, the relative wide band gap (3.4 eV) limits its photocatalytic activity only to be active in the ultraviolet light region (4% of the solar spectrum). Therefore, there is a growing necessity to develop photocatalysts with an efficient response under visible-light irradiation. In this sense, Shi et al. design a heterostructure between NaNbO3 and C3N4 in order to extended the absorption spectral range for visible region [96]. The NaNbO3 sample was obtained through hydrothermal synthesis, and the heterostructure was tailored by annealing NaNbO3 with melamine (C3N4 precursor) at 520 °C for 4 h. The XRD patterns of g-C3N4/NaNbO3 revealed coexistence crystalline phase of NaNbO3 and g-C3N4 in the composite. NaNbO3 nanowires are found to randomly deposit and distribute on the surface of g-C3N4 sheets, which results in forming a heterostructured g-C3N4/NaNbO3 material. The CO2 photoreduction under visible irradiation showed that g-C3N4/NaNbO3 improved over eight-fold the activity than that of bare C3N4 in the CH4 formation. Such a remarkable enhancement of photocatalytic activity was mainly attributed to the improved separation and transfer of photogenerated electron-hole pairs at the intimate interface of g-C3N4/NaNbO3 heterojunctions. The same group evaluated the CO2 photoreduction performance of g-C3N4/KNbO3, and the photocatalytic activity of g-C3N4/KNbO3 for CO2 reduction was almost four times higher than that of individual g-C3N4 under visible light irradiation [97].
Another Nb-based material that is worth mentioning is Sr2Bi2Nb2TiO12, an Aurivillius-type perovskite. Yu et al. investigated, for the first time, the Sr2Bi2Nb2TiO12 as photocatalyst for CO2 photoreduction [98]. The introduction of oxygen vacancies on the surface of Sr2Bi2Nb2TiO12 extends photoresponse to cover the whole visible region and also tremendously promotes the separation of photoinduced charge carriers. Compared with the bulk material prepared by traditional solid-state reaction, the Sr2Bi2Nb2TiO12 nanosheets with the optimal oxygen vacancies concentration yields a substantially high CO evolution rate of 17.11 μmol g−1 h−1.
Niobium has been used also as dopant, as it can improve the photocatalytic activities of anatase TiO2 [61,99,100,101], by changing the crystal structure, electrical properties, and absorption characteristics of TiO2. It was proposed that Nb would dope into TiO2 lattice and decrease the band gap energy of the photocatalyst, and thereby Nb substitution on the Ti site creates a Nb5+ defect state located upon the conduction band minimum. Therefore, Nb doped TiO2 would work as a visible light-driven photocatalyst [61,100,101].
Nogueira et al. investigated the role of Nb-doped TiO2 for CO2 photocatalytic reduction to methanol [102]. The samples were obtained through the Pechini method, which promoted the substitution of Ti4+ for Nb4+ at different concentrations. It was observed that the surface area tends to increase with the concentration of Nb. The concentration of formed methanol was directly proportional to the Nb concentration in the TiO2, the authors attributed the formation of methanol to the presence of Nb in the lattice of TiO2.
The evaluation of CO2 photoreduction performance of Z-scheme formed between Nb–TiO2 and g-C3N4 was also evaluated [103]. It was confirmed that the band gap energy of the Nb–TiO2 (2.91 eV) was lower than that of the TiO2 (3.2 eV). The formation of Z-scheme Nb–TiO2/g-C3N4 was evidenced through XRD and TEM analysis. CO, CH4, and O2 were formed from CO2 photoreduction when the Nb–TiO2 sample was used. Nb–TiO2/g-C3N4 not only converted CO2 to CO, CH4, and O2, but also generated HCOOH. Additionally, 50Nb–TiO2/50g-C3N4 was the best material for the CO2 reduction.
It is well known that, while metals and semiconductors usually exhibit great stability, but poor selectivity for CO2 reduction, enzymes and metal complexes can selectively convert CO2 into CO, but they lack long-term stability. Therefore, Faustino et al. evaluated the role of hexaniobate nanoscrolls (KxH(4−x)Nb6O17) as suitable substrates for the immobilization of two kinds of Re(I)-based molecular catalysts (namely I and II) for CO2 photoreduction [104]. Hexaniobate nanoscrolls were obtained through solid state reaction with subsequent proton exchange and exfoliation with tert-butylammonium hydroxide. The Re(I) complexes were immobilized by an adsorption process that was verified for UV/vis absorption and Fourier-transform infrared (FTIR) spectroscopies. As a result of the hexaniobate exfoliation, the specific surface area was increased from 2 m2 g−1 to 110 m2 g−1. The sensitized oxides were employed as visible-light photocatalysts for CO2 reduction in 5:1 dimethylformamide (DMF)/TEOA exposed to a 300 W Xe lamp (λ = 420 nm). The possible gaseous products were analyzed by mass spectrometry with CO being the only detected photoproduct. The complexes immobilized on the surfaces of the oxides exhibited much higher turnover number (TONCO) than those in solution. For comparative purposes, the complexes were immobilized on TiO2 as well. While for complex I, the immobilization on the niobate nanoscrolls led to 20% higher TONCO in comparison with that observed on TiO2, for complex II, the opposite behavior was observed; that is, the immobilization on TiO2 led to better photoactivities. Such behavior was rationalized based on the recombination kinetics of each composite and on the reductive potential of the injected electrons on the niobate or TiO2 conduction bands. For comparison, a summary of the photoactivity of Nb-based photocatalysts towards CO2 reduction is presented in Table 4.

4. Concluding Remarks and Perspectives

Over the past few years, great efforts have been spent by the scientific community to develop clean renewable energy sources. Photocatalysis can be described as an efficient method to produce solar fuels from available sources as H2O and CO2. Nevertheless, it is always a challenge to develop efficient materials able to promote solar-to-fuel conversion. In this sense, niobium-based materials offer diverse possibilities to engineer compounds showing important characteristics for photocatalytic processes such as efficient light absorption, charge separation, and charge transfer reaction. Herein, we could provide a glimpse from the last years on the extensive flexibility offered by niobium-based compounds for CO2 reduction and notably for H2 evolution. Niobium oxides as Nb2O5 and the wide range of niobates offer untold possibilities of structure, morphologies, and electronic configuration, which can supply desired properties for photocatalytic reactions. Moreover, niobium cations as a dopant have provided the opportunity to tune band energy positions and increase the efficient of other materials. Thus, niobium-based compounds are promising materials that still offer opportunities of new investigations and insights regarding photocatalytic applications.

Funding

This research was funded by Doutorado CAPES/DAAD/CNPQ (15/2017) grant number 88887.161403/2017-00. The present study was also financially supported by the Saint-Petersburg State University (Grant No. 39054581) Osmando F. Lopes also acknowledges to CNPq (Conselho Nacional de Desenvolvimento Científico e Tecnológico) for the financial support (grants #407497/2018-8).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. BP Statistical Review of World Energy. Available online: https://www.bp.com/content/dam/bp/business-sites/en/global/corporate/pdfs/energy-economics/statistical-review/bp-stats-review-2019-full-report.pdf (accessed on 13 November 2019).
  2. Christoforidis, K.C.; Fornasiero, P. Photocatalysis for Hydrogen Production and CO2 Reduction: The Case of Copper-Catalysts. ChemCatChem 2019, 11, 368–382. [Google Scholar] [CrossRef]
  3. Li, K.; Peng, B.; Peng, T.Y. Recent Advances in Heterogeneous Photocatalytic CO2 Conversion to Solar Fuels. ACS Catal. 2016, 6, 7485–7527. [Google Scholar] [CrossRef]
  4. Kim, J.H.; Hansora, D.; Sharma, P.; Jang, J.W.; Lee, J.S. Toward practical solar hydrogen production—An artificial photosynthetic leaf-to-farm challenge. Chem. Soc. Rev. 2019, 48, 1908–1971. [Google Scholar] [CrossRef] [PubMed]
  5. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 38–40. [Google Scholar] [CrossRef] [PubMed]
  6. Nico, C.; Monteiro, T.; Graça, M.P.F. Niobium oxides and niobates physical properties: Review and prospects. Prog. Mater. Sci. 2016, 80, 1–37. [Google Scholar] [CrossRef]
  7. Lopes, O.F.; de Mendonça, V.R.; Silva, F.B.F.; Paris, E.C.; Ribeiro, C. Niobium oxides: An overview of the synthesis of Nb2O5 and its application in heterogeneous photocatalysis. Quim. Nova 2014, 38, 106–117. [Google Scholar] [CrossRef]
  8. Nunes, B.N.; Faustino, L.A.; Muller, A.V.; Polo, A.S.; Patrocinio, A.O.T. Nb2O5 dye-sensitized solar cells. In Nanomaterials for Solar Cell Application, 2nd ed.; Thomas, S., Sakho, E.H.M., Kalarikkal, N., Oluwafemi, S.O., Wu, J., Eds.; Elsevier: Oxford, UK, 2019; pp. 287–322. [Google Scholar]
  9. Nowak, I.; Ziolek, M. Niobium Compounds: Preparation, Characterization, and Application in Heterogeneous Catalysis. Chem. Rev. 1999, 99, 3603–3624. [Google Scholar] [CrossRef]
  10. Ziolek, M. Niobium-containing catalysts—The state of the art. Catal. Today 2003, 78, 47–64. [Google Scholar] [CrossRef]
  11. Bizeto, M.A.; Shiguihara, A.L.; Constantino, V.R.L. Layered niobate nanosheets: Building blocks for advanced materials assembly. J. Mater. Chem. 2009, 19, 2512–2525. [Google Scholar] [CrossRef]
  12. Sarahan, M.C.; Carroll, E.C.; Allen, M.; Larsen, D.S.; Browning, N.D.; Osterloh, F.E. K4Nb6O17-derived photocatalysts for hydrogen evolution from water: Nanoscrolls versus nanosheets. J. Solid State Chem. 2008, 181, 1678–1683. [Google Scholar] [CrossRef]
  13. Ebina, Y.; Sasaki, T.; Watanabe, M. Study on exfoliation of layered perovskite-type niobates. Solid State Ion. 2002, 151, 177–182. [Google Scholar] [CrossRef]
  14. Shiguihara, A.L.; Bizeto, M.A.; Constantino, V.R.L. Exfoliation of layered hexaniobate in tetra(n-butyl)ammonium hydroxide aqueous solution. Colloids Surf. A Physicochem. Eng. Asp. 2007, 295, 123–129. [Google Scholar] [CrossRef]
  15. Ravishankar, T.N.; Vaz, M.d.O.; Ramakrishnappa, T.; Teixeira, S.R.; Dupont, J. Ionic liquid–assisted hydrothermal synthesis of Nb/TiO2 nanocomposites for efficient photocatalytic hydrogen production and photodecolorization of Rhodamine B under UV-visible and visible light illuminations. Mater. Today Chem. 2019, 12, 373–385. [Google Scholar] [CrossRef]
  16. Yang, M.; Kim, D.; Jha, H.; Lee, K.; Paul, J.; Schmuki, P. Nb doping of TiO2 nanotubes for an enhanced efficiency of dye-sensitized solar cells. Chem. Commun. 2011, 47, 2032–2034. [Google Scholar] [CrossRef]
  17. Emeline, A.V.; Furubayashi, Y.; Zhang, X.; Jin, M.; Murakami, T.; Fujishima, A. Photoelectrochemical behavior of Nb-doped TiO2 electrodes. J. Phys. Chem. B 2005, 109, 24441–24444. [Google Scholar] [CrossRef]
  18. Singh, R.; Ryu, I.; Yadav, H.; Park, J.; Jo, J.W.; Yim, S.; Lee, J.J. Non-hydrolytic sol-gel route to synthesize TiO2 nanoparticles under ambient condition for highly efficient and stable perovskite solar cells. Sol. Energy 2019, 185, 307–314. [Google Scholar] [CrossRef]
  19. El-Shazly, T.S.; Hassan, W.M.I.; Abdel Rahim, S.T.; Allam, N.K. Unravelling the interplay of dopant concentration and band structure engineering of monoclinic niobium pentoxide: A model photoanode for water splitting. Int. J. Hydrogen Energy 2015, 40, 13867–13875. [Google Scholar] [CrossRef]
  20. Kulkarni, A.K.; Praveen, C.S.; Sethi, Y.A.; Panmand, R.P.; Arbuj, S.S.; Naik, S.D.; Ghule, A.V.; Kale, B.B. Nanostructured N-doped orthorhombic Nb2O5 as an efficient stable photocatalyst for hydrogen generation under visible light. Dalton Trans. 2017, 46, 14859–14868. [Google Scholar] [CrossRef]
  21. Huang, Q.Z.; Wang, J.C.; Wang, P.P.; Yao, H.C.; Li, Z.J. In-situ growth of mesoporous Nb2O5 microspheres on g-C3N4 nanosheets for enhanced photocatalytic H2 evolution under visible light irradiation. Int. J. Hydrogen Energy 2017, 42, 6683–6694. [Google Scholar] [CrossRef]
  22. Wen, P.H.; Ai, L.L.; Liu, T.T.; Hu, D.W.; Yao, F.Y. Hydrothermal topological synthesis and photocatalyst performance of orthorhombic Nb2O5 rectangle nanosheet crystals with dominantly exposed (010) facet. Mater. Des. 2017, 117, 346–352. [Google Scholar] [CrossRef]
  23. Wen, P.H.; Ai, L.L.; Wei, F.Y.; Hu, D.W.; Guo, J.J.; Yao, F.Y.; Liu, T.T. In-situ synthesis of crystalline Ag-Nb2O5 nanobelt clusters with enhanced solar photo-electrochemical performance for splitting water. Mater. Des. 2017, 131, 219–225. [Google Scholar] [CrossRef]
  24. Zhou, C.; Shi, R.; Yang, G.; Meng, X.; Wu, L.Z.; Tung, C.H.; Zhang, T. Spatial separation of charge carriers in Nb2O5 nanorod superstructures for enhanced photocatalytic H2 production activity. Mater. Today Chem. 2018, 10, 259–263. [Google Scholar] [CrossRef]
  25. Zhang, Y.; Zhao, H.; Zhao, X.F.; Lin, J.N.; Li, N.; Huo, Z.Y.; Yan, Z.F.; Zhang, M.; Hu, S. Narrow-bandgap Nb2O5 nanowires with enclosed pores as high-performance photocatalyst. Sci. China Mater. 2019, 62, 203–210. [Google Scholar] [CrossRef] [Green Version]
  26. Kumar, A.; Islam, R.; Pramanik, D.; Saraswat, K. On the limit of defect doping in transition metal oxides. J. Vac. Sci. Technol. A 2019, 37. [Google Scholar] [CrossRef]
  27. Zhao, W.L.; Zhao, W.; Zhu, G.L.; Lin, T.Q.; Xu, F.F.; Huang, F.Q. Black Nb2O5 nanorods with improved solar absorption and enhanced photocatalytic activity. Dalton Trans. 2016, 45, 3888–3894. [Google Scholar] [CrossRef] [PubMed]
  28. Huang, H.; Zhou, J.; Zhou, J.; Zhu, M.S. Structure-retentive synthesis of a highly ordered mesoporous Nb2O5/N-doped graphene nanocomposite with superior interfacial contacts and improved visible-light photocatalysis. Catal. Sci. Technol. 2019, 9, 3373–3379. [Google Scholar] [CrossRef]
  29. Khan, I.; Baig, N.; Qurashi, A. Graphitic carbon nitride impregnated niobium oxide (g-C3N4/Nb2O5) type (II) heterojunctions and its synergetic solar-driven hydrogen generation. ACS Appl. Energy Mater. 2019, 2, 607–615. [Google Scholar] [CrossRef]
  30. Idrees, F.; Dillert, R.; Bahnemann, D.; Butt, F.; Tahir, M. In-Situ Synthesis of Nb2O5/g-C3N4 Heterostructures as Highly Efficient Photocatalysts for Molecular H2 Evolution under Solar Illumination. Catalysts 2019, 9, 169. [Google Scholar] [CrossRef] [Green Version]
  31. Nunes, B.N.; Patrocinio, A.O.T.; Bahnemann, D.W. Influence of the preparation conditions on the morphology and photocatalytic performance Pt-modified hexaniobate composites. J. Phys. Condens. Matter 2019, 31, 394001. [Google Scholar] [CrossRef]
  32. Reisman, A.; Holtzberg, F. Phase Equilibria in the System K2CO3-Nb2O5 by the Method of Differential Thermal Analysis. J. Am. Chem. Soc. 1955, 77, 2115–2119. [Google Scholar] [CrossRef]
  33. Nassau, K.; Shiever, J.W.; Bernstein, J.L. Crystal Growth and Properties of Nica-Like Potassium Niobates. J. Electrochem. Soc. 1969, 116, 348–353. [Google Scholar] [CrossRef]
  34. Dion, M.; Ganne, M.; Tournoux, M. Nouvelles familles de phases MIMII2Nb3O10 a feuillets “perovskites”. Mater. Res. Bull. 1981, 16, 1429–1435. [Google Scholar] [CrossRef]
  35. Jacobson, A.J.; Johnson, J.W.; Lewandowski, J.T. Interlayer Chemistry between Thick Transition-Metal Oxide Layers: Synthesis and Intercalation Reactions of K[Ca2Nan-3NbnO3n+1] (3 ≤ n ≤ 7). Inorg. Chem. 1985, 24, 3727–3729. [Google Scholar] [CrossRef]
  36. Nunes, B.N.; Haisch, C.; Emeline, A.V.; Bahnemann, D.W.; Patrocinio, A.O.T. Photocatalytic properties of layer-by-layer thin films of hexaniobate nanoscrolls. Catal. Today 2019, 326, 60–67. [Google Scholar] [CrossRef]
  37. Oshima, T.; Lu, D.; Maeda, K. Preparation of Pt-Intercalated KCa2Nb3O10 Nanosheets and Their Photocatalytic Activity for Overall Water Splitting. ChemNanoMat 2016, 2, 748–755. [Google Scholar] [CrossRef]
  38. Kulischow, N.; Ladasiu, C.; Marschall, R. Layered Dion-Jacobson type niobium oxides for photocatalytic hydrogen production prepared via molten salt synthesis. Catal. Today 2017, 287, 65–69. [Google Scholar] [CrossRef]
  39. Zhou, Y.N.; Wen, T.; Guo, Y.Z.; Yang, B.C.; Wang, Y.G. Controllable doping of nitrogen and tetravalent niobium affords yellow and black calcium niobate nanosheets for enhanced photocatalytic hydrogen evolution. RSC Adv. 2016, 6, 64930–64936. [Google Scholar] [CrossRef]
  40. Zhou, Y.N.; Wen, T.; Zhang, X.F.; Chang, B.B.; Kong, W.Q.; Guo, Y.Z.; Yang, B.C.; Wang, Y.G. A Multiple Structure-Design Strategy towards Ultrathin Niobate Perovskite Nanosheets with Thickness-Dependent Photocatalytic Hydrogen-Evolution Performance. Chem. Asian J. 2017, 12, 2727–2733. [Google Scholar] [CrossRef] [PubMed]
  41. Xiong, J.H.; Jing, K.Q.; Zou, J.H.; Liang, S.J.; Wu, L. A hybrid of CdS/HCa2Nb3O10 ultrathin nanosheets for promoting photocatalytic hydrogen evolution. Dalton Trans. 2017, 46, 13935–13942. [Google Scholar] [CrossRef] [PubMed]
  42. Hu, Y.C.; Li, G.S.; Zong, S.C.; Shi, J.W.; Guo, L.J. Self-assembled nanohybrid of cadmium sulfide and calcium niobate: Photocatalyst with enhanced charge separation for efficient visible light induced hydrogen generation. Catal. Today 2018, 315, 117–125. [Google Scholar] [CrossRef]
  43. Xiong, J.H.; Liu, Y.H.; Liang, S.J.; Zhang, S.Y.; Li, Y.H.; Wu, L. Insights into the role of Cu in promoting photocatalytic hydrogen production over ultrathin HNb3O8 nanosheets. J. Catal. 2016, 342, 98–104. [Google Scholar] [CrossRef]
  44. Xia, Y.Z.; Liang, S.J.; Wu, L.; Wang, X.X. Ultrasmall NiS decorated HNb3O8 nanosheeets as highly efficient photocatalyst for H2 evolution reaction. Catal. Today 2019, 330, 195–202. [Google Scholar] [CrossRef]
  45. Wang, J.J.; Teng, J.; Pu, L.Z.; Huang, J.; Wang, Y.; Li, Q.X. Double-hole-mediated coupling of anionic dopants in perovskite NaNbO3 for efficient solar water splitting. Int. J. Quantum Chem. 2019, 119, e25930. [Google Scholar] [CrossRef]
  46. Kaneko, M.; Mishima, K.; Yamashita, K. First-principles study on visible light absorption of defected SrNbO3. J. Photochem. Photobiol. A Chem. 2019, 375, 175–180. [Google Scholar] [CrossRef]
  47. Yu, J.X.; Chen, Z.Q.; Chen, Q.Q.; Wang, Y.; Lin, H.J.; Hu, X.; Zhao, L.H.; He, Y.M. Giant enhancement of photocatalytic H2 production over KNbO3 photocatalyst obtained via carbon doping and MoS2 decoration. Int. J. Hydrogen Energy 2018, 43, 4347–4354. [Google Scholar] [CrossRef]
  48. Huang, S.S.; Lang, J.Y.; Du, C.F.; Bian, F.G.; Su, Y.G.; Wang, X.J. Enhanced driving force and charge separation efficiency in disordered SnNbxOy: Boosting photocatalytic activity toward water reduction. Chem. Eng. J. 2017, 309, 313–320. [Google Scholar] [CrossRef]
  49. Kamimura, S.; Abe, S.; Tsubota, T.; Ohno, T. Solar-driven H2 evolution over CuNb2O6: Effect of two polymorphs (monoclinic and orthorhombic) on optical property and photocatalytic activity. J. Photochem. Photobiol. A Chem. 2018, 356, 263–271. [Google Scholar] [CrossRef]
  50. Chun, Y.T.; Yue, M.F.; Jiang, P.F.; Chen, S.J.; Gao, W.L.; Cong, R.H.; Yang, T. Optimizing the performance of photocatalytic H2 generation for ZnNb2O6 synthesized by a two-step hydrothermal method. RSC Adv. 2018, 8, 13857–13864. [Google Scholar] [CrossRef] [Green Version]
  51. Betzler, S.B.; Podjaski, F.; Beetz, M.; Handloser, K.; Wisnet, A.; Handloser, M.; Hartschuh, A.; Lotsch, B.V.; Scheu, C. Titanium Doping and Its Effect on the Morphology of Three-Dimensional Hierarchical Nb3O7(OH) Nanostructures for Enhanced Light-Induced Water Splitting. Chem. Mater. 2016, 28, 7666–7672. [Google Scholar] [CrossRef] [Green Version]
  52. Zhang, B.; Hui, D.P.; Li, Y.X.; Zhao, H.; Wang, C.Y. Synthesis and photocatalytic hydrogen production activity of the Ni-CH3CH2NH2/H1.78Sr0.78Bi0.22Nb2O7 hybrid layered perovskite. Chin. J. Catal. 2017, 38, 2039–2047. [Google Scholar] [CrossRef]
  53. Grodziuk, G.Y.; Shcherban, N.D.; Shvalagin, V.V.; Korzhak, A.V.; Andryushina, N.S.; Skoryk, M.A.; Kuchmiy, S.Y. Photocatalytic activity of nanostructured composites based on layered niobates and C3N4 in the hydrogen evolution reaction from electron donor solutions under visible light. Int. J. Hydrogen Energy 2017, 42, 24108–24116. [Google Scholar] [CrossRef]
  54. Thaweesak, S.; Lyu, M.; Peerakiatkhajohn, P.; Butburee, T.; Luo, B.; Chen, H.J.; Wang, L.Z. Two-dimensional g-C3N4/Ca2Nb2TaO10 nanosheet composites for efficient visible light photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2017, 202, 184–190. [Google Scholar] [CrossRef]
  55. Wang, K.; Li, Y.; Li, J.; Zhang, G. Boosting interfacial charge separation of Ba5Nb4O15/g-C3N4 photocatalysts by 2D/2D nanojunction towards efficient visible-light driven H2 generation. Appl. Catal. B Environ. 2019, 263, 117730. [Google Scholar] [CrossRef]
  56. Chen, P.F.; Xing, P.X.; Chen, Z.Q.; Hu, X.; Lin, H.J.; Zhao, L.H.; He, Y.M. In-situ synthesis of AgNbO3/g-C3N4 photocatalyst via microwave heating method for efficiently photocatalytic H2 generation. J. Colloid Interface Sci. 2019, 534, 163–171. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, Z.Q.; Chen, P.F.; Xing, P.X.; Hu, X.; Lin, H.J.; Wu, Y.; Zhao, L.H.; He, Y.M. Rapid fabrication of KTa0.75Nb0.25/g-C3N4 composite via microwave heating for efficient photocatalytic H2 evolution. Fuel 2019, 241, 1–11. [Google Scholar] [CrossRef]
  58. Fujito, H.; Kunioku, H.; Kato, D.; Suzuki, H.; Higashi, M.; Kageyama, H.; Abe, R. Layered Perovskite Oxychloride Bi4NbO8Cl: A Stable Visible Light Responsive Photocatalyst for Water Splitting. J. Am. Chem. Soc. 2016, 138, 2082–2085. [Google Scholar] [CrossRef]
  59. Wakayama, H.; Hibino, K.; Fujii, K.; Oshima, T.; Yanagisawa, K.; Kobayashi, Y.; Kimoto, K.; Yashima, M.; Maeda, K. Synthesis of a Layered Niobium Oxynitride, Rb2NdNb2O6NH2O, Showing Visible-Light Photocatalytic Activity for H2 Evolution. Inorg. Chem. 2019, 58, 6161–6166. [Google Scholar] [CrossRef]
  60. Tian, Z.L.; Cui, H.L.; Xu, J.J.; Zhu, G.L.; Shao, F.; He, J.Q.; Huang, F.Q. Efficient Charge Separation of In-Situ Nb-Doped TiO2 Nanowires for Photoelectrochemical Water–splitting. ChemistrySelect 2017, 2, 2822–2827. [Google Scholar] [CrossRef]
  61. Kong, L.; Wang, C.H.; Zheng, H.; Zhang, X.T.; Liu, Y.C. Defect-Induced Yellow Color in Nb-Doped TiO2 and Its Impact on Visible-Light Photocatalysis. J. Phys. Chem. C 2015, 119, 16623–16632. [Google Scholar] [CrossRef]
  62. Caudillo-Flores, U.; Muñoz-Batista, M.J.; Cortés, J.A.; Fernández-García, M.; Kubacka, A. UV and visible light driven H2 photo-production using Nb-doped TiO2: Comparing Pt and Pd co-catalysts. Mol. Catal. 2017, 437, 1–10. [Google Scholar] [CrossRef]
  63. Flores, E.; Ares, J.R.; Sánchez, C.; Ferrer, I.J. Ternary transition titanium-niobium trisulfide as photoanode for assisted water splitting. Catal. Today 2019, 321–322, 107–112. [Google Scholar] [CrossRef]
  64. Zoellner, B.; O’Donnell, S.; Wu, Z.; Itanze, D.; Carbone, A.; Osterloh, F.E.; Geyer, S.; Maggard, P.A. Impact of Nb(V) Substitution on the Structure and Optical and Photoelectrochemical Properties of the Cu5(Ta1−xNbx)11O30 Solid Solution. Inorg. Chem. 2019, 58, 6845–6857. [Google Scholar] [CrossRef] [PubMed]
  65. Thanh Truc, N.T.; Tran, D.T.; Hanh, N.T.; Pham, T.D. Novel visible light-driven Nb-doped Ta3N5 sensitized/protected by PPy for efficient overall water splitting. Int. J. Hydrogen Energy 2018, 43, 15898–15906. [Google Scholar] [CrossRef]
  66. Chen, Z.Q.; Chen, P.F.; Xing, P.X.; Hu, X.; Lin, H.J.; Wu, Y.; Zhao, L.H.; He, Y.M. Novel carbon modified KTa0.75Nb0.25O3 nanocubes with excellent efficiency in photocatalytic H2 evolution. Fuel 2018, 233, 486–496. [Google Scholar] [CrossRef]
  67. Monfort, O.; Sfaelou, S.; Satrapinskyy, L.; Plecenik, T.; Roch, T.; Plesch, G.; Lianos, P. Comparative study between pristine and Nb-modified BiVO4 films employed for photoelectrocatalytic production of H2 by water splitting and for photocatalytic degradation of organic pollutants under simulated solar light. Catal. Today 2017, 280, 51–57. [Google Scholar] [CrossRef]
  68. Fontelles-Carceller, O.; Muñoz-Batista, M.J.; Conesa, J.C.; Kubacka, A.; Fernández-García, M. H2 photo-production from methanol, ethanol and 2-propanol: Pt-(Nb)TiO2 performance under UV and visible light. Mol. Catal. 2018, 446, 88–97. [Google Scholar] [CrossRef]
  69. Kalanur, S.S.; Yoo, I.H.; Cho, I.S.; Seo, H. Niobium incorporated WO3 nanotriangles: Band edge insights and improved photoelectrochemical water splitting activity. Ceram. Int. 2019, 45, 8157–8165. [Google Scholar] [CrossRef]
  70. Patel, P.P.; Hanumantha, P.J.; Velikokhatnyi, O.I.; Datta, M.K.; Gattu, B.; Poston, J.A.; Manivannan, A.; Kumta, P.N. Vertically aligned nitrogen doped (Sn,Nb)O2 nanotubes—Robust photoanodes for hydrogen generation by photoelectrochemical water splitting. Mater. Sci. Eng. B Solid State Mater. Adv. Technol. 2016, 208, 1–14. [Google Scholar] [CrossRef] [Green Version]
  71. IPCC The Intergovernmental Panel on Climate Change. Available online: http://www.ipcc.ch/ (accessed on 20 November 2019).
  72. Lim, C.H.; Holder, A.M.; Hynes, J.T.; Musgrave, C.B. Catalytic Reduction of CO2 by Renewable Organohydrides. J. Phys. Chem. Lett. 2015, 6, 5078–5092. [Google Scholar] [CrossRef]
  73. Mondal, B.; Song, J.; Neese, F.; Ye, S. Bio-inspired mechanistic insights into CO2 reduction. Curr. Opin. Chem. Biol. 2015, 25, 103–109. [Google Scholar] [CrossRef]
  74. Mao, Z.Y.; Chen, J.J.; Yang, Y.F.; Wang, D.J.; Bie, L.J.; Fahlman, B.D. Novel g-C3N4/CoO Nanocomposites with Significantly Enhanced Visible-Light Photocatalytic Activity for H2 Evolution. ACS Appl. Mater. Interfaces 2017, 9, 12427–12435. [Google Scholar] [CrossRef]
  75. Gutierrez-Guerra, N.; Moreno-Lopez, L.; Serrano-Ruiz, J.C.; Valverde, J.L.; de Lucas-Consuegra, A. Gas phase electrocatalytic conversion of CO2 to syn-fuels on Cu based catalysts-electrodes. Appl. Catal. B Environ. 2016, 188, 272–282. [Google Scholar] [CrossRef]
  76. Ciamician, G. The Photochemistry of the Future. Science 1912, 36, 385–394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photoelectrocatalytic reduction of carbon dioxide in aqueous suspensions of semiconductor powders. Nature 1979, 277, 637–638. [Google Scholar] [CrossRef]
  78. D’Alessandro, D.M.; Smit, B.; Long, J.R. Carbon dioxide capture: Prospects for new materials. Angew. Chem. Int. Ed. 2010, 49, 6058–6082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Kauffman, D.R.; Thakkar, J.; Siva, R.; Matranga, C.; Ohodnicki, P.R.; Zeng, C.; Jin, R. Efficient Electrochemical CO2 Conversion Powered by Renewable Energy. ACS Appl. Mater. Interfaces 2015, 7, 15626–15632. [Google Scholar] [CrossRef] [PubMed]
  80. Hsieh, Y.C.; Senanayake, S.D.; Zhang, Y.; Xu, W.Q.; Polyansky, D.E. Effect of Chloride Anions on the Synthesis and Enhanced Catalytic Activity of Silver Nanocoral Electrodes for CO2 Electroreduction. ACS Catal. 2015, 5349–5356. [Google Scholar] [CrossRef]
  81. Kortlever, R.; Shen, J.; Schouten, K.J.P.; Calle-Vallejo, F.; Koper, M.T.M. Catalysts and Reaction Pathways for the Electrochemical Reduction of Carbon Dioxide. J. Phys. Chem. Lett. 2015, 6, 4073–4082. [Google Scholar] [CrossRef]
  82. Ma, Y.; Wang, X.L.; Jia, Y.S.; Chen, X.B.; Han, H.X.; Li, C. Titanium dioxide-based nanomaterials for photocatalytic fuel generations. Chem. Rev. 2014, 114, 9987–10043. [Google Scholar] [CrossRef]
  83. Ong, W.J.; Tan, L.L.; Ng, Y.H.; Yong, S.T.; Chai, S.P. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer To Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  84. Shi, H.F.; Wang, T.Z.; Chen, J.; Zhu, C.; Ye, J.H.; Zou, Z.G. Photoreduction of carbon dioxide over NaNbO3 nanostructured photocatalysts. Catal. Lett. 2011, 141, 525–530. [Google Scholar] [CrossRef]
  85. Li, P.; Ouyang, S.; Zhang, Y.J.; Kako, T.; Ye, J.H. Surface-coordination-induced selective synthesis of cubic and orthorhombic NaNbO3 and their photocatalytic properties. J. Mater. Chem. A 2013, 1, 1185–1191. [Google Scholar] [CrossRef]
  86. Li, P.; Ouyang, S.; Xi, G.C.; Kako, T.; Ye, J.H. The Effects of Crystal Structure and Electronic Structure on Photocatalytic H2 Evolution and CO2 Reduction over Two Phases of Perovskite-Structured NaNbO3. J. Phys. Chem. C 2012, 116, 7621–7628. [Google Scholar] [CrossRef] [Green Version]
  87. Fresno, F.; Jana, P.; Reñones, P.; Coronado, J.M.; Serrano, D.P.; De La Peña O’Shea, V.A. CO2 reduction over NaNbO3 and NaTaO3 perovskite photocatalysts. Photochem. Photobiol. Sci. 2017, 16, 17–23. [Google Scholar] [CrossRef] [PubMed]
  88. Ohno, T.; Sarukawa, K.; Tokieda, K.; Matsumura, M. Morphology of a TiO2 Photocatalyst (Degussa, P-25) Consisting of Anatase and Rutile Crystalline Phases. J. Catal. 2001, 86, 82–86. [Google Scholar] [CrossRef]
  89. Li, P.; Xu, H.; Liu, L.Q.; Kako, T.; Umezawa, N.; Abe, H.; Ye, J.H. Constructing cubic-orthorhombic surface-phase junctions of NaNbO3 towards significant enhancement of CO2 photoreduction. J. Mater. Chem. A 2014, 2, 5606–5609. [Google Scholar] [CrossRef]
  90. Low, J.X.; Cao, S.W.; Yu, J.G.; Wageh, S. Two-dimensional layered composite photocatalysts. Chem. Commun. 2014, 50, 10768–10777. [Google Scholar] [CrossRef]
  91. Li, X.K.; Kikugawa, N.; Ye, J.H. Nitrogen-doped lamellar niobic acid with visible light-responsive photocatalytic activity. Adv. Mater. 2008, 20, 3816–3819. [Google Scholar] [CrossRef]
  92. Li, X.K.; Pan, H.Q.; Li, W.; Zhuang, Z.J. Photocatalytic reduction of CO2 to methane over HNb3O8 nanobelts. Appl. Catal. A Gen. 2012, 413–414, 103–108. [Google Scholar] [CrossRef]
  93. Li, X.K.; Li, W.; Zhuang, Z.J.; Zhong, Y.S.; Li, Q.; Wang, L.Y. Photocatalytic Reduction of Carbon Dioxide to Methane over SiO2-Pillared HNb3O8. J. Phys. Chem. C 2012, 116, 16047–16053. [Google Scholar] [CrossRef]
  94. Kormányos, A.; Thomas, A.; Huda, M.N.; Sarker, P.; Liu, J.P.; Poudyal, N.; Janáky, C.; Rajeshwar, K. Solution Combustion Synthesis, Characterization, and Photoelectrochemistry of CuNb2O6 and ZnNb2O6 Nanoparticles. J. Phys. Chem. C 2016, 120, 16024–16034. [Google Scholar] [CrossRef] [Green Version]
  95. da Silva, G.T.S.T.; Nogueira, A.E.; Oliveira, J.A.; Torres, J.A.; Lopes, O.F.; Ribeiro, C. Acidic surface niobium pentoxide is catalytic active for CO2 photoreduction. Appl. Catal. B Environ. 2019, 242, 349–357. [Google Scholar] [CrossRef]
  96. Shi, H.F.; Chen, G.Q.; Zhang, C.L.; Zou, Z.G. Polymeric g-C3N4 coupled with NaNbO3 nanowires toward enhanced photocatalytic reduction of CO2 into renewable fuel. ACS Catal. 2014, 4, 3637–3643. [Google Scholar] [CrossRef]
  97. Shi, H.F.; Zhang, C.L.; Zhou, C.P.; Chen, G.Q. Conversion of CO2 into renewable fuel over Pt-g-C3N4/KNbO3 composite photocatalyst. RSC Adv. 2015, 5, 93615–93622. [Google Scholar] [CrossRef]
  98. Yu, H.J.; Li, J.Y.; Zhang, Y.H.; Yang, S.Q.; Han, K.; Dong, F.; Ma, T.Y.; Huang, H.W. Three-in-One Oxygen Vacancies: Whole Visible-Spectrum Absorption, Efficient Charge Separation, and Surface Site Activation for Robust CO2 Photoreduction. Angew. Chem. Int. Ed. 2019, 58, 3880–3884. [Google Scholar] [CrossRef]
  99. Yang, X.Y.; Min, Y.X.; Li, S.B.; Wang, D.W.; Mei, Z.W.; Liang, J.; Pan, F. Conductive Nb-doped TiO2 thin films with whole visible absorption to degrade pollutants. Catal. Sci. Technol. 2018, 8, 1357–1365. [Google Scholar] [CrossRef]
  100. Mattsson, A.; Leideborg, M.; Larsson, K.; Westing, G.; Österlund, L. Adsorption and solar light decomposition of acetone on anatase TiO2 and niobium doped TiO2 thin films. J. Phys. Chem. B 2006, 110, 1210–1220. [Google Scholar] [CrossRef]
  101. Yang, J.K.; Zhang, X.T.; Wang, C.H.; Sun, P.P.; Wang, L.L.; Xia, B.; Liu, Y.C. Solar photocatalytic activities of porous Nb-doped TiO2 microspheres prepared by ultrasonic spray pyrolysis. Solid State Sci. 2012, 14, 139–144. [Google Scholar] [CrossRef]
  102. Nogueira, M.V.; Lustosa, G.M.M.M.; Kobayakawa, Y.; Kogler, W.; Ruiz, M.; Monteiro Filho, E.S.; Zaghete, M.A.; Perazolli, L.A. Nb-Doped TiO2 Photocatalysts Used to Reduction of CO2 to Methanol. Adv. Mater. Sci. Eng. 2018, 2018, 7326240. [Google Scholar] [CrossRef] [Green Version]
  103. Thanh Truc, N.T.; Giang Bach, L.; Thi Hanh, N.; Pham, T.D.; Thi Phuong Le Chi, N.; Tran, D.T.; Nguyen, M.V.; Nguyen, V.N. The superior photocatalytic activity of Nb doped TiO2/g-C3N4 direct Z-scheme system for efficient conversion of CO2 into valuable fuels. J. Colloid Interface Sci. 2019, 540, 1–8. [Google Scholar] [CrossRef]
  104. Faustino, L.A.; Souza, B.L.; Nunes, B.N.; Duong, A.T.; Sieland, F.; Bahnemann, D.W.; Patrocinio, A.O.T. Photocatalytic CO2 Reduction by Re(I) Polypyridyl Complexes Immobilized on Niobates Nanoscrolls. ACS Sustain. Chem. Eng. 2018, 6, 6073–6083. [Google Scholar] [CrossRef]
Figure 1. O-Nb2O5 microscopies images of (a) field emission scanning electron microscopy (FE-SEM), (b) high-resolution transmission electron microscopy (HRTEM), (c) transmission electron microscopy (TEM), and (d) selected area electron diffraction (SAED) pattern and (e) diagram for the transformation of O-Nb2O5 nanosheets from H4Nb6O17 3H2O (HNbO) nanosheets under hydrothermal conditions. Reproduced with permission from the authors of [22]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Figure 1. O-Nb2O5 microscopies images of (a) field emission scanning electron microscopy (FE-SEM), (b) high-resolution transmission electron microscopy (HRTEM), (c) transmission electron microscopy (TEM), and (d) selected area electron diffraction (SAED) pattern and (e) diagram for the transformation of O-Nb2O5 nanosheets from H4Nb6O17 3H2O (HNbO) nanosheets under hydrothermal conditions. Reproduced with permission from the authors of [22]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Catalysts 10 00126 g001
Figure 2. Scheme of Nb2O5 nanorod superstructures synthesis and the improved photogenerated charge carries separation. Reproduced with permission from the authors of [24]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Figure 2. Scheme of Nb2O5 nanorod superstructures synthesis and the improved photogenerated charge carries separation. Reproduced with permission from the authors of [24]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Catalysts 10 00126 g002
Figure 3. Schematic illustration of the transfer of photoexcited electrons and production of H2 photocatalyzed by MNb–N-doped graphene (NGR) under visible light. Reproduced with permission from the authors of [28]. Copyright (2019), The Royal Society of Chemistry, London, United Kingdom.
Figure 3. Schematic illustration of the transfer of photoexcited electrons and production of H2 photocatalyzed by MNb–N-doped graphene (NGR) under visible light. Reproduced with permission from the authors of [28]. Copyright (2019), The Royal Society of Chemistry, London, United Kingdom.
Catalysts 10 00126 g003
Figure 4. FE-SEM image of the surface of Pt-modified hexaniobate layer-by-layer (LbL) films with 25 bilayers (the bright small spots are Pt nanoclusters). Reproduced with permission from the authors of [36]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Figure 4. FE-SEM image of the surface of Pt-modified hexaniobate layer-by-layer (LbL) films with 25 bilayers (the bright small spots are Pt nanoclusters). Reproduced with permission from the authors of [36]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Catalysts 10 00126 g004
Figure 5. Schematic representation of (a) the general structure of KCa2Nb3O10 (KCNNO), (b) KCa2Nan−3NbnO3n+1 with thickness control of the perovskite layer by varying the n value, and (c) preparation of N/Nb4+ codoped nanosheets (CNNO). Reproduced with permission from the authors of [40]. Copyright (2019), John Wiley & Sons, Inc., Hoboken, NJ, USA.
Figure 5. Schematic representation of (a) the general structure of KCa2Nb3O10 (KCNNO), (b) KCa2Nan−3NbnO3n+1 with thickness control of the perovskite layer by varying the n value, and (c) preparation of N/Nb4+ codoped nanosheets (CNNO). Reproduced with permission from the authors of [40]. Copyright (2019), John Wiley & Sons, Inc., Hoboken, NJ, USA.
Catalysts 10 00126 g005
Figure 6. Band edge positions of the pure and codoped NaNbO3. Reproduced with permission from the authors of [45]. Copyright (2019), John Wiley & Sons, Inc., Hoboken, NJ, USA. NHE, normal hydrogen electrode.
Figure 6. Band edge positions of the pure and codoped NaNbO3. Reproduced with permission from the authors of [45]. Copyright (2019), John Wiley & Sons, Inc., Hoboken, NJ, USA. NHE, normal hydrogen electrode.
Catalysts 10 00126 g006
Figure 7. Photocatalytic process over orthorhombic CuNb2O6, comparing the presence and absence of near-infrared (NIR)-cutoff filter. C.B. and V.B. correspond to conduction band and valence band, respectively. AM 1.5G defines the employed irradiation conditions following the ASTM G-173-03 Standard. Reproduced with permission from the authors of [49]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Figure 7. Photocatalytic process over orthorhombic CuNb2O6, comparing the presence and absence of near-infrared (NIR)-cutoff filter. C.B. and V.B. correspond to conduction band and valence band, respectively. AM 1.5G defines the employed irradiation conditions following the ASTM G-173-03 Standard. Reproduced with permission from the authors of [49]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Catalysts 10 00126 g007
Figure 8. SEM images of (a) Nb3O7(OH) and (b,c) Ti(IV) modified samples with the titanium content averaged over the morphology indicated relative to niobium (scale bar of 1 μm). Reprinted with permission from the authors of [51]. Copyright (2019) American Chemical Society, Washington, DC, USA.
Figure 8. SEM images of (a) Nb3O7(OH) and (b,c) Ti(IV) modified samples with the titanium content averaged over the morphology indicated relative to niobium (scale bar of 1 μm). Reprinted with permission from the authors of [51]. Copyright (2019) American Chemical Society, Washington, DC, USA.
Catalysts 10 00126 g008
Figure 9. Scheme of Z-scheme water splitting of Bi4NbO8Cl coupled with Ru/SrTiO3/Rh photocatalyst via Fe3+/Fe2+ redox mediator. Reprinted with permission from the authors of [58]. Copyright (2019) American Chemical Society, Washington, DC, USA.
Figure 9. Scheme of Z-scheme water splitting of Bi4NbO8Cl coupled with Ru/SrTiO3/Rh photocatalyst via Fe3+/Fe2+ redox mediator. Reprinted with permission from the authors of [58]. Copyright (2019) American Chemical Society, Washington, DC, USA.
Catalysts 10 00126 g009
Figure 10. H2 evolution over Pt/Rb2NdNb2O6N in DMSO/TEOA solution containing 1 mL of water. Reprinted with permission from the authors of [59]. Copyright (2019) American Chemical Society, Washington, DC, USA.
Figure 10. H2 evolution over Pt/Rb2NdNb2O6N in DMSO/TEOA solution containing 1 mL of water. Reprinted with permission from the authors of [59]. Copyright (2019) American Chemical Society, Washington, DC, USA.
Catalysts 10 00126 g010
Figure 11. Reaction rate of the hydrogen photoproduction using different alcohols. Me, Et, and Is stand for methanol, ethanol, and 2-propanol (isopropanol), respectively. Reproduced with permission from the authors of [68]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Figure 11. Reaction rate of the hydrogen photoproduction using different alcohols. Me, Et, and Is stand for methanol, ethanol, and 2-propanol (isopropanol), respectively. Reproduced with permission from the authors of [68]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Catalysts 10 00126 g011
Figure 12. (a) UV/vis absorption spectra of SnO2 NTs and (Sn1−xNbx)O2 NTs (x = 0.05, 0.1) and (b) theoretical H2 gas and measured H2 and O2 gases, during 24 h chronoamperometry test of (Sn0.95Nb0.05)O2/N performed in 0.5 M H2SO4 solution at 0.75 V (vs. RHE) and 26 °C. Reproduced with permission from the authors of [70]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Figure 12. (a) UV/vis absorption spectra of SnO2 NTs and (Sn1−xNbx)O2 NTs (x = 0.05, 0.1) and (b) theoretical H2 gas and measured H2 and O2 gases, during 24 h chronoamperometry test of (Sn0.95Nb0.05)O2/N performed in 0.5 M H2SO4 solution at 0.75 V (vs. RHE) and 26 °C. Reproduced with permission from the authors of [70]. Copyright (2019), Elsevier, Amsterdam, Netherlands.
Catalysts 10 00126 g012
Figure 13. Band gap (Eg) and band positions of Nb-based semiconductors vs. normal hydrogen electrode (NHE). On the right side, the standard potentials of several redox couples are presented for comparison.
Figure 13. Band gap (Eg) and band positions of Nb-based semiconductors vs. normal hydrogen electrode (NHE). On the right side, the standard potentials of several redox couples are presented for comparison.
Catalysts 10 00126 g013
Figure 14. TEM images of (a) NaNbO3–cubic and (b) NaNbO3–orthorhombic. HR-TEM images of (c) NaNbO3–orthorhombic and (d) NaNbO3–cubic. Reproduced with permission from the authors of [85]. Copyright (2019), Royal Society of Chemistry, London, United Kingdom.
Figure 14. TEM images of (a) NaNbO3–cubic and (b) NaNbO3–orthorhombic. HR-TEM images of (c) NaNbO3–orthorhombic and (d) NaNbO3–cubic. Reproduced with permission from the authors of [85]. Copyright (2019), Royal Society of Chemistry, London, United Kingdom.
Catalysts 10 00126 g014
Figure 15. (a) X-ray diffraction (XRD) patterns of the as-prepared NaNbO3 samples. (b) Dependence of cubic and orthorhombic NaNbO3 contents on the annealing temperature. (c) UV/vis absorption spectra of the as-prepared NaNbO3 samples and the inset in the figure is the corresponding (αhν)1/2–hν curves. Reproduced with permission from the authors of [89]. Copyright (2019), Royal Society of Chemistry, London, United Kingdom.
Figure 15. (a) X-ray diffraction (XRD) patterns of the as-prepared NaNbO3 samples. (b) Dependence of cubic and orthorhombic NaNbO3 contents on the annealing temperature. (c) UV/vis absorption spectra of the as-prepared NaNbO3 samples and the inset in the figure is the corresponding (αhν)1/2–hν curves. Reproduced with permission from the authors of [89]. Copyright (2019), Royal Society of Chemistry, London, United Kingdom.
Catalysts 10 00126 g015
Table 1. Summary of the photocatalytic activity of niobates combined with distinct materials towards H2 evolution from aqueous solution.
Table 1. Summary of the photocatalytic activity of niobates combined with distinct materials towards H2 evolution from aqueous solution.
NiobateCo-CatalystLight SourceSacrificial ReagentH2 Formation RateRef.
HCa2Nb3O10CdS, Pt300 W Xe, >400 nmLactic acid52.6 μmol h−1[41]
HCa2Nb3O10CdS300 W Xe, >420 nmmethanol450 μmol g−1 h−1[42]
HNb3O8Cu2+300 W Xemethanol98.2 µmol h−1[43]
HNb3O8NiS300 W Xetriethanolamine1519.4 μmol g−1 h−1[44]
C-doped KNbO3MoS2, Pt300 W Xemethanol1300 µmol g−1 h−1[47]
H1.78Sr0.78Bi0.22Nb2O7Ni-CH3CH2NH2300 W Xemethanol372.67 μmol h−1[52]
KNb3O8g-C3N41000 W Hg λ > 400 nmdimethylhydrazine25.0 µmol h−1 g−1[53]
K3H3Nb10.8O30g-C3N41000 W Hg λ > 400 nmdimethylhydrazine27.0 µmol h−1 g−1[53]
HNb3O8g-C3N41000 W Hg λ > 400 nmdimethylhydrazine37.0 µmol h−1 g−1[53]
Ca2Nb2TaO10g-C3N4, Pt300 W Xe, >400 nmtriethanolamine43.54 µmol h−1[54]
Ba5Nb4O15g-C3N43 W LEDs 420 nmoxalic acid2673 μmol h−1 g−1[55]
AgNbO3g-C3N4, Pt300 W Xe, >420 nmmethanol88.0 µmol g−1 h−1[56]
KTa0.75Nb0.25O3g-C3N4, Pt300 W Xe, >420 nmmethanol86.2 μmol·g−1·h−1[57]
Table 2. Summary of Nb-doped photocatalysts employed for H2 evolution from aqueous solution.
Table 2. Summary of Nb-doped photocatalysts employed for H2 evolution from aqueous solution.
SemiconductorCo-CatalystLight SourceSacrificial ReagentH2 RateRef.
TiO2-300 W Xe > 420 nmethanol1146 µmol g−1[15]
TiO2Pd500 W Hg-Xe 420–680 nmmethanol~0.6 mmol g−1 h−1[62]
TiO2Pt500 W Hg-Xe 420–680 nmmethanol~0.1 mmol g−1 h−1[62]
TiS3-200 W halogenNa2SO3 (electrolyte)2.2 μmol min−1·cm2[63]
Cu5Ta11O30Pt1000 W Xe arc > 420 nmmethanol6 μmol h−1[64]
Ta3N5-400–700 nmwater41.4 µmol g−1 h−1[65]
KTaO3Pt300 W Xemethanol728 µmol g−1 h−1[66]
BiVO4-100 mW cm−2XeNaHCO3 (electrolyte)0.18 mmol h−1[67]
Table 3. Common products from CO2 reduction, their simplified half reactions, and redox potentials at pH = 7.
Table 3. Common products from CO2 reduction, their simplified half reactions, and redox potentials at pH = 7.
ReactionsE (V vs. RHE)
CO2 + 2e + 2H+ → CO + H2O−0.53
CO2 + 2e + 2H+ → HCOOH−0.61
CO2 + 4e + 4H+ → HCOH + H2O−0.48
CO2 + 6e + 6H+ → CH3OH + H2O−0.38
CO2 + 8e + 8H+ → CH4 + 2H2O−0.24
Table 4. Summary of the photocatalytic activity of Nb-based compounds combined with distinct materials towards CO2 reduction.
Table 4. Summary of the photocatalytic activity of Nb-based compounds combined with distinct materials towards CO2 reduction.
Nb-Based MaterialCo-CatalystLight SourceMain ProductsObserved RateRef.
NaNbO3Pt300 W XeCH4653 ppm h−1 g−1[84]
NaNbO3 Pt300 W Xe > 300 nmCH410.4 μmol h−1 m−2[85]
NaNbO3None300 W Xe > 420 nmCO, CH4, CH3OH, H2NA[87]
NaNbO3 heterojunctionPt300 W Xe > 300 nmCH4~10 μmol h−1 m−2[89]
HNb3O8None350 W XeCH43.5 μmol g−1 h−1[92]
SiO2-Pillared HNb3O8Pt350 W XeCH42.90 µmol g−1 h−1[93]
Nb2O5NoneUVC lamp, 0.167 mW cm−2CO, CH4, CH3COOH, HCOOH~10 μmol L−1 g−1[95]
C3N4/NaNbO3None300 W Xe λ > 420 nmCH4~6 µmol h−1 g−1[96]
C3N4/KNbO3Pt300 W Xe λ > 420 nmCH40.25 µmol h−1[97]
Re(I) polypyridyl complexes/KxH4−xNb6O17None300 W Xe λ > 420 nmCO2.9 *[104]
* Turnover frequency (TOF), h–1.

Share and Cite

MDPI and ACS Style

Nunes, B.N.; Lopes, O.F.; Patrocinio, A.O.T.; Bahnemann, D.W. Recent Advances in Niobium-Based Materials for Photocatalytic Solar Fuel Production. Catalysts 2020, 10, 126. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10010126

AMA Style

Nunes BN, Lopes OF, Patrocinio AOT, Bahnemann DW. Recent Advances in Niobium-Based Materials for Photocatalytic Solar Fuel Production. Catalysts. 2020; 10(1):126. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10010126

Chicago/Turabian Style

Nunes, Barbara Nascimento, Osmando Ferreira Lopes, Antonio Otavio T. Patrocinio, and Detlef W. Bahnemann. 2020. "Recent Advances in Niobium-Based Materials for Photocatalytic Solar Fuel Production" Catalysts 10, no. 1: 126. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10010126

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop