Protonated Oxide, Nitrided, and Reoxidized K2La2Ti3O10 Crystals

Oct 26, 2016 - Synopsis. Protonated oxide, nitrided, and reoxidized layered K2La2Ti3O10 crystals (H2La2Ti3O10 and H2La2Ti3O10-3/2xNx) are studied for ...
0 downloads 3 Views 7MB Size
Research Article pubs.acs.org/journal/ascecg

Protonated Oxide, Nitrided, and Reoxidized K2La2Ti3O10 Crystals: Visible-Light-Induced Photocatalytic Water Oxidation and Fabrication of Their Nanosheets Kenta Kawashima,†,‡ Mirabbos Hojamberdiev,† Hajime Wagata,† Kunio Yubuta,§ Kazunari Domen,∥ and Katsuya Teshima*,†,⊥ †

Department of Environmental Science and Technology, Faculty of Engineering, Shinshu University, 4-17-1 Wakasato, Nagano 380-8553, Japan ‡ McKetta Department of Chemical Engineering, The University of Texas at Austin, Austin, Texas 78712, United States § Institute for Materials Research, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan ∥ Department of Chemical System Engineering, School of Engineering, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan ⊥ Center for Energy and Environmental Science, Shinshu University, 4-17-1 Wakasato, Nagano 380-8553, Japan S Supporting Information *

ABSTRACT: Protonated lanthanum titanium oxide H2La2Ti3O10 and oxynitride H2La2Ti3O10−3/2xNx crystals were synthesized from the oxide, nitrided, and reoxidized layered K2La2Ti3O10 crystals prepared by solid-state reaction through proton exchange. Here, we investigated the holding time of nitridation of oxide K2La2Ti3O10 crystals influencing their crystal structure, shape, and absorption wavelength and band gap energy. The XRD and SEM results confirmed that the crystal structure and plate-like shape of the parent oxide were maintained after nitridation at 800 °C for 10 h, and the color of crystals was changed from white to dark green. However, no clear absorption edges were observed in the UV−vis diffuse reflectance spectra of the nitrided crystals due mainly to the reduced titanium species (Ti3+), which act as the recombination center of the photogenerated charge carriers. To decrease the amount of the reduced titanium species, the nitrided crystals were further reoxidized at 400 °C for 6 h. After partial reoxidation, the absorption intensity in the longer wavelength region was reduced, and the absorption edges appeared at about 449−460 nm. The photocatalytic activity for the water oxidation halfreaction was evaluated only for the protonated samples. The protonated reoxidized K2La2Ti3O10 crystals showed the O2 evolution rate of 180 nmol·h−1 (for the photocatalytic water oxidation) under visible-light irradiation, and the unexpected photocatalytic decomposition of N2O adsorbed onto the photocatalyst surfaces was observed for the protonated oxide and protonated nitrided layered K2La2Ti3O10 crystals. Furthermore, lanthanum titanium oxide [La2Ti3O10]2− and oxynitride [La2Ti3O10−3/2xNx]2− nanosheets were successfully fabricated by proton exchange and mechanical exfoliation (sonication) of the oxide, nitrided, and reoxidized K2La2Ti3O10 crystals. The TEM results revealed that the lateral sizes of the fabricated nanosheets grown along the ⟨001⟩ direction are 270−620 nm. Apparently, the colloidal suspensions of the fabricated nanosheets showed a Tyndall effect, implying their good dispersion and stability for several weeks in water. KEYWORDS: Layered structure, Water oxidation, Oxynitride, Perovskite, Nanosheet, Visible light



INTRODUCTION Since the first report of the Honda−Fujishima effect,1 various types of photocatalytic materials have been researched for water and air purification, antifouling, antifogging, antibacterial, sterilization, and hydrogen production. Recently, layered perovskite oxides, such as K2La2Ti3O10,2 ABi2Nb2O9 (A = Ca, Sr, and Ba),3 RbLnTa2O7 (Ln = La, Pr, Nd, and Sm),4 and ABi 2 Ta 2 O 9 (A = Ca, Sr, and Ba), 5 were studied as photocatalysts for water splitting to generate hydrogen by © 2016 American Chemical Society

utilizing solar energy. Moreover, their derivatives can easily be designed by exchanging the cations of the interlayer with other cations.6 The layered Ruddlesden−Popper phase K2La2Ti3O10 having triple perovskite layers made up of octahedral TiO6 is a Received: June 14, 2016 Revised: October 7, 2016 Published: October 26, 2016 232

DOI: 10.1021/acssuschemeng.6b01344 ACS Sustainable Chem. Eng. 2017, 5, 232−240

Research Article

ACS Sustainable Chemistry & Engineering tetragonal compound with space group I4/mmm and lattice parameters of a = b = 3.8767(1) Å, c = 29.824(1) Å, and α = β = γ = 90° (Figure 1, the crystal structure was drawn using

for different times. Also, the nitrided K 2 La 2 Ti 3 O 10 (K2La2Ti3O10−3/2xNx) crystals were further reoxidized in order to decrease the excess amount of reduced species of Ti3+ to Ti4+. Here, we also discuss the effect of protonation on photocatalytic activity of the oxide, nitrided, and reoxidized K2La2Ti3O10 crystals for water oxidation activity under visible light and the exfoliation of the oxide, nitrided, and reoxidized K2La2Ti3O10 crystals into their nanosheets.



EXPERIMENTAL SECTION

Growth of K2La2Ti3O10 Crystals. The layered K2La2Ti3O10 crystals were grown by a solid-state reaction using reagent-grade KNO3, La2O3, and TiO2 (>98%, Wako Pure Chemical Industries, Ltd.) according to the experimental procedure reported previously elsewhere.9 A stoichiometric mixture of KNO3, La2O3, and TiO2 were dry mixed manually using an agate mortar and pestle for 15 min. An excess amount of KNO3 (25 mol %) was added to compensate potassium volatilized at high temperature. The mixture (10 g) was placed in a platinum crucible with a capacity of 30 cm3 and closed loosely with a platinum lid. The mixture-containing platinum crucible was heated at 1000 °C for 6 h at a heating rate of 50 °C·h−1, cooled to 500 °C at a cooling rate of 150 °C·h−1, and then cooled naturally to room temperature. Preparation of Protonated Forms of Oxide, Nitrided, and Reoxidized K2La2Ti3O10 Crystals. To obtain the layered oxynitride crystals, 0.5 g of the K2La2Ti3O10 crystals grown by a solid-state reaction was placed on an alumina plate and heated at 800 °C for 3, 5, 7, and 10 h at a heating rate of 600 °C·h−1 under an NH3 flow (200 mL·min−1) in a horizontal tubular furnace and cooled naturally to room temperature. The K2La2Ti3O10 crystals nitrided at 800 °C for 10 h were then reoxidized at 400 °C for 6 h at a heating rate of 600 °C· h−1. The protonated forms of oxide, nitrided, and reoxidized K2La2Ti3O10 crystals were obtained by immersing 0.25 g of crystals in 50 mL of 0.1 M HCl solution for 7 days, washing with deionized water several times, and drying at room temperature. Fabrication of Nanosheets. The nanosheets of oxide, nitrided (10 h), and reoxidized K2La2Ti3O10 crystals were fabricated through a proton exchange process by immersing 0.25 g of crystals in 50 mL of 0.1 M HCl solution for 7 days, washing with deionized water several times, and drying at room temperature and exfoliation by sonicating 0.01 g of the protonated crystals in 5 mL of deionized water. Photocatalytic Water Oxidation Activity Test. The photocatalytic O2 evolution reactions were carried out with 100 mg of protonated oxide, nitrided or reoxidized K2La2Ti3O10 crystals, and 200 mg of La2O3 (pH buffer) in 300 mL of a 10 mM AgNO3 (>99%, Wako Pure Chemical Industries, Ltd.) aqueous solution (sacrificial electron scavenger) in a glass cell connected to a closed circulation system under visible-light irradiation (300W Xe lamp with a λ > 420 nm cutoff filter and cold mirror), and the light intensity was 200 mW·cm−2. The evolved gases were detected by a gas chromatograph (GC-8A, TCD, Ar gas carrier, Shimadzu). Characterization. The X-ray diffraction (XRD) patterns of crystal samples were recorded on a MiniflexII X-ray diffractometer (Rigaku). The crystal morphology was observed by using a JSM-7600F scanning electron microscope (SEM, JEOL). The crystal orientations of the fabricated nanosheets were elucidated by an EM-002B high-resolution transmission electron microscope (HR-TEM, TOPCON) at an acceleration voltage of 200 kV. The ultraviolet−visible (UV−vis) diffuse reflectance spectra of crystal samples were measured using a V670 spectrophotometer (JASCO). Their absorption edges and band gap energies were estimated from the UV−vis diffuse reflectance spectra by using Kubelka−Munk theory. The reoxidation temperature of K2La2Ti3O10 crystals nitrided at 800 °C for 10 h was determined by simultaneous thermogravimetry and differential thermal analysis (TGDTA, Thermo plus EVO2, Rigaku) by heating 0.02 g of sample from room temperature up to 1200 °C at a heating rate of 10 °C·min−1 under synthetic air flow. The surface chemical compositions of the protonated samples were analyzed by X-ray photoelectron spectroscopy (XPS, JPS-9010MC, JEOL) using a nonmonochromatic Mg Kα

Figure 1. Schematic illustration of the crystal structure of K2La2Ti3O10.

VESTA 3).7−9 Having a band gap energy (Eg) of about 3.5 eV,2,6 K2La2Ti3O10 has shown photoluminescence,10 upconversion luminescence,11 and photocatalytic properties.2 Particularly, cocatalyst-loaded K2La2Ti3O10 (Ni−K2La2Ti3O10 and Cr−Ni−K2La2Ti3O10) showed high photocatalytic activity for overall water splitting under ultraviolet (UV) light irradiation.2,6,12 The ultimate aim of the study of photocatalytic water splitting is to achieve efficient energy conversion from solar energy to chemical energy (H2) by using visible-light-active photocatalysts. Theoretically, the sun light (λ < 1100 nm) can be utilized for photocatalytic a water-splitting reaction.13 As K2La2Ti3O10 has a wide band gap energy of 3.5 eV, it can only be excited under UV light (λ < 350 nm).2,6 To improve its photocatalytic activity by extending its response to the visiblelight region, K2La2Ti3O10 was doped with different cations and anions, i.e., K2−xLa2Ti3−xNbxO10 (0 ⩽ x ⩽ 1),14 zinc- and vanadium-doped K2La2Ti3O10,15,16 K2La2Ti3−xMxO10+δ (M = Fe, Ni, and W),17 nitrogen-doped K2La2Ti3O10,18 and Sn2+ and N3−-substituted K2La2Ti3O10,19 to engineer the band gap of K2La2Ti3O10. Furthermore, K2La2Ti3O10 was also composited with narrow band gap semiconductors, such as CdS,20 ZnIn2S4,21 and BiOBr,22 to improve its photocatalytic activity under visible-light irrigation. Recently, the protonated layered compounds have also shown an enhanced photocatalytic water splitting activity compared with nonprotonated counterparts due to the hydration of the interlayer cations and molecules.23 Protonated K2La2Ti3O10 (H2La2Ti3O10), which was composited with other semiconductors, such as TiO2, Fe2O3, and CdS, has also exhibited an enhanced photocatalytic activity because of the effective charge separation achieved by the coupling of two semiconductors having different band gap energy levels and the short travel distance of the photogenerated charge carriers, diffusing to reach the interface.24−26 Although the nitrogen-doped K2La2Ti3O10 crystals have been fabricated by immersing in aqueous NH3 solution or heating with urea,18,19 the subsequent nitridation of the K2La2Ti3O10 crystals under an NH3 flow at high temperature in order to form the oxynitride crystals of K2La2Ti3O10−3/2xNx has not been reported yet. In this study, we have synthesized K2La2Ti3O10 crystals by solid-state reaction9 and investigated the nitridation behavior of K2La2Ti3O10 crystals under an NH3 flow at 800 °C 233

DOI: 10.1021/acssuschemeng.6b01344 ACS Sustainable Chem. Eng. 2017, 5, 232−240

Research Article

ACS Sustainable Chemistry & Engineering X-ray source. The XPS profiles were fitted using a Gaussian− Lorentzian function, and the peak positions were normalized by positioning the C 1s peak at 284.5 eV.

the XRD patterns of crystals nitrided for 7 and 10 h slightly shifted to a lower 2θ angle owing to the expansion of the lattice volume toward a- and b-axes. As the crystal radius of N3− (157 pm) is larger than that of O2− (126 pm), this increase in the lattice volume may be caused by the presence of partially substituted nitrogen at oxygen sites as well as the insertion of nitrogen at interstitial sites in the lattice of triple perovskite sheets ([La2Ti3O10]2−).28,29 It should be noted that the layered structure of K2La2Ti3O10 was maintained even after hightemperature nitridaiton for longer period. The SEM images of the as-synthesized and nitrided K2La2Ti3O10 crystals are shown in Figure 3. As shown, the



RESULTS AND DISCUSSION Figure 2 shows the XRD patterns of the layered K2La2Ti3O10 crystals synthesized by a solid-state reaction and K2La2Ti3O10

Figure 2. XRD patterns of K2La2Ti3O10 crystals synthesized by solidstate reaction (a) and K2La2Ti3O10 crystals nitrided at 800 °C for 3 (b), 5 (c), 7 (d), and 10 h (e).

crystals nitrided at 800 °C for different holding times (3, 5, 7, and 10 h) under an NH3 flow (200 mL·min−1). The XRD pattern of the oxide crystals is identified as an anhydrous K2La2Ti3O10 phase (ICDD PDF 81-1167) with a minor unknown phase. The XRD patterns of crystals nitrided at 800 °C for 3 and 5 h contain diffraction peaks assignable to K2La2Ti3O10 and K2La2Ti3O10·1.6H2O (ICDD PDF 82-0208). It is thought that the formation of the layered K2La2Ti3O10· 1.6H2O as an impurity phase may possibly be stemmed from the insertion of water molecules, which were formed during the nitridation process, into the interlayer spacing of the layered K2La2Ti3O10, resulting in the expansion of c-axis, as described below:6,8

Figure 3. SEM images of K2La2Ti3O10 crystals synthesized by solidstate reaction (a) and K2La2Ti3O10 crystals nitrided at 800 °C for 3 (b), 5 (c), 7 (d), and 10 h (e).

K 2La 2Ti3O10 + 1/2x N2 + 3/2x H 2 → K 2La 2Ti3O10 − 3/2x Nx + 3/2x H 2O

as-synthesized oxide crystals have a platelet shape and irregular lateral size ranging from 0.8 to 5.8 μm (with thickness of approximately 150−230 nm). After nitridation, the nitrided crystals maintained the shape of oxide crystals, indicating that the K2La2Ti3O10 crystals are stable under high-temperature nitridation for long period. Figure 4a shows the UV−vis diffuse reflectance spectra of the as-synthesized and nitrided K2La2Ti3O10 crystals. The color of the as-synthesized K2La2Ti3O10 crystals is pale yellow with an absorption edge wavelength at about 364 nm and a narrower band gap energy of about 3.41 eV compared with that of the previously reported K2La2Ti3O10 crystals (Eg = 3.63−3.69 eV).18,19 This difference possibly resulted from the incorporation of a small amount of nitrogen into the lattice during solid-state synthesis of K2La2Ti3O10. After nitridation, the absorption of crystals presumably shifted upward in the visiblelight region (λ = 350−800 nm), their absorption edge eventually disappeared, and the color gradually altered from pale yellow to dark green with increasing the nitridation time. This upward shift in the visible-light region and color change can be explained by the following three reasons: (i) negative shift of the valence band due to the partial substitution of oxygen by nitrogen forming the hybrid orbitals of O 2p and N

(1)

K 2La 2Ti3O10 − 3/2x Nx + 1.6H 2O → K 2La 2Ti3O10 − 3/2x Nx ·1.6H 2O

(2)

In contrast, the XRD patterns of crystals nitrided at 800 °C for 7 and 10 h have the diffraction peaks matching with those of anhydrous K2La2Ti3O10 and the unknown phase. Interestingly, the K2La2Ti3O10·1.6H2O phase completely disappeared after prolonging the nitridation time to 7 and 10 h, implying that most of the interlayer water was evaporated. Compared to that of the as-synthesized K2La2Ti3O10 crystals, the 002 diffraction peak at 5.92° of the K2La2Ti3O10 crystals nitrided at 800 °C for 7 and 10 h shifted toward a lower 2θ angle due to the small amount of remaining interlayer water molecules. In the XRD patterns of crystals nitrided for 3 and 5 h, the 110 diffraction peak at 32.64° of the K2La2Ti3O10 crystals slightly shifted to higher 2θ angle compared to that of the as-synthesized K2La2Ti3O10 crystals, which might be associated with the superposition of the 110 diffraction peaks at 32.64 and 32.80° from the K2La2Ti3O10 and K2La2Ti3O10·1.6H2O crystals, respectively.27 On the other hand, the 110 diffraction peak in 234

DOI: 10.1021/acssuschemeng.6b01344 ACS Sustainable Chem. Eng. 2017, 5, 232−240

Research Article

ACS Sustainable Chemistry & Engineering

Figure 4. (a) UV−vis diffuse reflectance spectra of K2La2Ti3O10 crystals synthesized by solid-state reaction (blue) and K2La2Ti3O10 crystals nitrided at 800 °C for 3 (green), 5 (orange), 7 (purple), and 10 h (red). (b) TG-DTA curves of K2La2Ti3O10 crystals nitrided at 800 °C for 10 h.

Figure 5. XRD patterns of reoxidized K2La2Ti3O10 crystals nitrided at 800 °C for 3 (a), 5 (b), 7 (c), and 10 h (d).

2p, (ii) absorption of interstitial nitrogen impurities having an isolated N 2p narrow band above the O 2p valence band in the forbidden band of K2La2Ti3O10−3/2xNx, and (iii) absorption of the d-d transition of the reduced Ti3+ species (oxygen vacancies) caused by nitrogen substitution.18,19,29−31 For instance, K2La2Ti3O10 crystals nitrided for 3 h have a broadened peak at 550 nm due to the absorption of interstitial nitrogen,30 and the absorption intensity in longer wavelength (λ = 600− 800 nm) became higher due to the d-d transition of the reduced Ti3+ species in K2La2Ti3O10 crystals nitrided for 5, 7, and 10 h. As known, the defects such as the reduced Ti3+ species may act as a recombination center for the photogenerated electrons and holes, causing a decrease in the photocatalytic activity.31 To determine the optimum partial reoxidation temperature for oxidizing Ti3+ into Ti4+ without destroying the layered structure of K2La2Ti3O10−3/2xNx, the reoxidation behavior of the K2La2Ti3O10 crystals nitrided at 800 °C for 10 h under an NH3 flow was studied by TG-DTA (Figure 4b) from 25 to 1300 °C in synthetic air. After TG-DTA analysis, the color of the K2La2Ti3O10−3/2xNx crystals changed from dark green to white because of reoxidation.32 The weight loss from 25 to 300 °C corresponds to the desorption of 0.56 wt % adsorbed water on the crystal surface (25−100 °C) and release of 0.45 wt % interlayer water (100−300 °C).8 A significant weight loss and an endothermic peak observed in the range of 300−350 °C can be attributed to the desorption of water contained in the oxide matrix of [La2Ti3O10]2−.33 The weight change from 350 to 940 °C is associated with the transformation from oxynitride to oxide (i.e., resubstitution of 3O2− for 2N3−) through two processes: (i) formation of the dinitrogen-containing intermediate compound (K2La2Ti3O10(N2)1/2x) and oxidation of Ti3+ into Ti4+ (350−570 °C) and (ii) the release of dinitrogen (N2) gas from K2La2Ti3O10(N2)1/2x crystals (570−940 °C):32

as K2La2Ti3O10 and K2La2Ti3O10·1.6H2O phases. After partial reoxidation, the peak intensities of the (001) plane of the layered K2La2Ti3O10·1.6H2O crystals at 5.25° decreased owing to the release of interlayer water partially converting K2La2Ti3O10·1.6H2O into K2La2Ti3O10. In addition, the diffraction peaks of the (110) plane of the K2La2Ti3O10 crystals at 32.64° slightly shifted to lower 2θ angle compared to that of the as-nitrided K2La2Ti3O10 crystals probably because of the elimination of the superposition of the 110 diffraction peaks of the K2La2Ti3O10 and K2La2Ti3O10·1.6H2O crystals. In the case of K2La2Ti3O10 crystals nitrided at 800 °C for 7 and 10 h, the XRD patterns of the partially reoxidized crystals can be well indexed as K2La2Ti3O10 without any diffraction peaks of impurity phases. After partial reoxidation, the diffraction peaks of the (110) plane of K2La2Ti3O10 crystals at 32.64° in the XRD patterns slightly shifted to a higher 2θ angle compared to that of K2La2Ti3O10 crystals nitrided for 7 and 10 h (parents) because of the decrease in the lattice volume. The possible reason for the decrease in the lattice volume is a partial release of interstitial and substantial nitrogen existing in the [La2Ti3O10]2− lattice during partial reoxidation process. The XRD results indicate that the nitrogen introduced into the layered structure of K2La2Ti3O10−3/2xNx is maintained even after reoxidation. Figure 6 shows the SEM images of nitrided K2La2Ti3O10 crystals partially reoxidized at 400 °C for 6 h. As shown, the platelet shape and size of the K2La2Ti3O10−3/2xNx crystals are

K 2La 2Ti3O10 − 3/2x Nx + 3/4xO2 → K 2La 2Ti3O10 (N2)1/2x (3)

K 2La 2Ti3O10 (N2)1/2x → K 2La 2Ti3O10 + 1/2x N2

(4)

From 940 to 1300 °C, another weight loss of 1.4 wt % and an exothermic peak are noted because of the partial decomposition of K2La2Ti3O10 to K2O and La2Ti3O9.9 From the TG-DTA results, 400 °C was found to be an optimum partial reoxidation temperature. Figure 5 shows the XRD patterns of the nitrided K2La2Ti3O10 crystals partially reoxidized at 400 °C for 6 h. For K2La2Ti3O10 crystals nitrided at 800 °C for 3 and 5 h, the XRD patterns of the partially reoxidized crystals are identified

Figure 6. SEM images of reoxidized K2La2Ti3O10 crystals nitrided at 800 °C for 3 (a), 5 (b), 7 (c), and 10 h (d). 235

DOI: 10.1021/acssuschemeng.6b01344 ACS Sustainable Chem. Eng. 2017, 5, 232−240

Research Article

ACS Sustainable Chemistry & Engineering

change are due to the reoxidation of Ti3+ into Ti4+. The clear absorption edges of the partially reoxidized crystals are found to be at about 449−460 nm. Hence, the narrower band gap energies of the partially reoxidized crystals are approximately 2.70−2.76 eV compared to that of the as-synthesized K2La2Ti3O10 crystals (Eg = 3.41 eV). The reduction in the optical band gap resulted from the negative shift of the valence band by partial substitution of nitrogen. Furthermore, the band gap energies of the partially reoxidized K2La2Ti3O10−3/2xNx crystals in this study are narrower than that of the nitrogendoped K2La2Ti3O10 crystals (Eg = 3.44−3.59 eV) reported previously.18,19 Presumably, this noticeable difference is accounted for by moving the valence band more upward because of the higher amount of nitrogen introduced in the K2La2Ti3O10 crystals. It is thought that high-temperature nitridation under an NH3 flow is more effective than immersing in an aqueous NH3 solution or heating with urea to introduce a higher amount of nitrogen into the K2La2Ti3O10 crystals. Figure 8 shows the TEM and lattice images, selected-area electron diffraction (SAED) patterns, and colloidal suspensions of the oxide, nitrided, and reoxidized nanosheets fabricated thorough the proton exchange and exfoliation of the assynthesized, nitrided (at 800 °C for 10 h), and reoxidized (at 400 °C for 6 h) K2La2Ti3O10 crystals. As mentioned above, the K2La2Ti3O10 crystals have a platelet shape and irregular lateral size ranging from 0.8 to 5.8 μm (with thickness of approximately 150−230 nm). According to the TEM results, the reduction in the lateral size of the fabricated nanosheets was observed (270−620 nm) due to the mechanical shear stress under sonication.34 As a single set of diffraction spots in the SAED patterns was observed for the fabricated nanosheets, and no grain boundaries were found in their lattice images, evidencing that the fabricated nanosheets have a singlecrystalline nature. Indexing of the SAED patterns obtained

maintained without any significant changes after partial reoxidation. However, the surfaces of the reoxidized crystals became rougher due to the introduction and release of nitrogen during the nitridation and reoxidation processes. The UV−vis diffuse reflectance spectra of the partially reoxidized K2La2Ti3O10−3/2xNx crystals are shown in Figure 7.

Figure 7. UV−vis diffuse reflectance spectra of K2La2Ti3O10 crystals synthesized by solid-state reaction (blue) and reoxidized K2La2Ti3O10 crystals nitrided at 800 °C for 3 (green), 5 (orange), 7 (purple), and 10 h (red).

As compared to the UV−vis spectra of the nitrided K2La2Ti3O10 crystals, the absorption intensity of the partially reoxidized K2La2Ti3O10−3/2xNx crystals was reduced in the longer wavelength region (λ = 440−800 nm), and the color significantly changed from dark green to yellow. The reduction in the absorption intensity in longer wavelength and color

Figure 8. TEM and HRTEM images, SAED patterns, and the colloidal suspensions of the oxide (a), nitrided (b), and reoxidized nanosheets (c) fabricated through proton exchange and exfoliation processes. 236

DOI: 10.1021/acssuschemeng.6b01344 ACS Sustainable Chem. Eng. 2017, 5, 232−240

Research Article

ACS Sustainable Chemistry & Engineering

compositions in H2La2Ti3O10 and H2La2Ti3O10−3/2xNx, and potassium is present with a low concentration, indicating successful K+/H+ ion exchange. The La 3d and 4d, Ti 2p, O 1s, and N 1s XPS core-level spectra of the protonated oxide, nitrided, and reoxidized K2La2Ti3O10 crystals are shown in Figure S3. The binding energies of La 3d (835.0 and 839.1 eV for La 3d5/2, and 851.9 and 856.0 eV for La 3d3/2) for the protonated oxide, nitrided, and reoxidized K2La2Ti3O10 crystals indicate the presence of La3+.36 In all the protonated crystals, three components are required to fit the La 4d peak. The high intensity peaks at 102.5 and 105.7 eV are associated with La 4d5/2 and La 4d3/2, respectively, and the low intensity broad peak at around 109.4 eV probably corresponds to a bonding satellite for La 4d3/2.36 The Ti 2p XPS core-level spectra of the protonated oxide, nitrided, and reoxidized K2La2Ti3O10 crystals showed the Ti 2p3/2 and Ti 2p1/2 peaks at around 458.7 and 464.4 eV, respectively, which can be recognized as the Ti4+ cations of the O − Ti bond.37 Here, the peaks at 455.7 and 462.2 eV in the Ti 2p XPS core-level spectra of the protonated nitrided K2La2Ti3O10 crystals are also originated from Ti 2p3/2 and Ti 2p1/2, which can be attributed to the reduced Ti species, such as Ti3+ cations of the O−Ti−N bond.37 This difference indicates that only the protonated nitrided K2La2Ti3O10 crystals may contain oxygen vacancies on the crystal surface. In the O 1s spectra of all the protonated crystals, the two components centered at around 530.7 and 532.7 eV can be assigned to the O2− anions (O−Ti linkage) and carbonate-like species (O−C linkage), respectively.38 Regarding the N 1s XPS core-level spectra of all the protonated crystals, the deconvoluted peaks can be classified into four: (i) substituted nitrogen of the O− Ti−N bond (around 396.8 eV), (ii) surface-adsorbed NH2 (399.5 eV), (iii) surface-adsorbed NHx bond (400.3 eV), and (iv) nitrogen species of the Ti−O−N or Ti−N−O bonds or surface-adsorbed NOx (400.8 eV), respectively.19,39−41 The XPS results confirmed that the protonated nitrided K2La2Ti3O10 crystals contain a higher amount of the substituted nitrogen compared to the protonated oxide and protonated reoxidized K2La2Ti3O10 crystals. The photocatalytic water oxidation activities of the protonated oxide, nitrided, and reoxidized K2La2Ti3O10 crystals were evaluated under visible-light irradiation (λ > 420 nm) in an aqueous AgNO3 solution, and the reaction time courses of O2 evolution are plotted in Figure 10. It is worth noting that this is the first report on the photocatalytic sacrificial O2 evolution over layered H2La2Ti3O10 and H2La2Ti3O10−3/2xNx

with the incident beam along the [001] zone axis direction reveals that the diffraction spots correspond to the (200), (110), (020), and (1−10) faces, indicating that the fabricated nanosheets were grown along the ⟨001⟩ direction. The in-plane surfaces of the fabricated nanosheets lie on the (001) plane, and the side surfaces are the (100), (010), and (110) planes. As shown in Figure 8, all the suspensions of the fabricated nanosheets clearly show a Tyndall effect, suggesting that the fabricated nanosheets have a good dispersion and are stable in water for several weeks.35 Figure 9 shows the XRD patterns of the protonated oxide, nitrided, and reoxidized K2La2Ti3O10 crystals. The XRD

Figure 9. XRD patterns and SEM images of protonated oxide (a), nitrided (b), and reoxidized K2La2Ti3O10 crystals (c).

patterns of all the protonated oxide, nitrided, and reoxidized crystals are identified as a H2La2Ti3O10 phase (ICDD PDF 48− 0983) with a minor unknown phase. This confirms that the interlayer K+ ions of the oxide, nitrided, and reoxidized K2La2Ti3O10 crystals were successfully exchanged with H+ ions during proton exchange. The structure of H2La2Ti3O10 in a layered perovskite type is closely related to that of parent K2La2Ti3O10, and the space group of the H2La2Ti3O10 crystal structure is the same as that of K2La2Ti3O10 (I4/mmm).9 However, the diffraction peak of the (002) plane of H2La2Ti3O10 shifted to a lower 2θ angle compared to that of K2La2Ti3O10 because of the expansion of the c-axis. An increase in the c-parameter is due to the presence of weaker interactions (hydrogen bonds) between the interlayers of [La2Ti3O10]2− unit layers in H2La2Ti3O10.25 Figure 9 shows the SEM images of the protonated oxide, nitrided, and reoxidized K2La2Ti3O10 crystals. After proton exchange, all the protonated crystals maintained the platelet shape of their parent oxide, nitrided, and reoxidized K2La2Ti3O10 crystals. The distribution of elements and chemical compositions of the protonated oxide, nitrided, and reoxidized K2La2Ti3O10 crystals were analyzed by energydispersive X-ray spectroscopy (EDS) and are shown in Figures S1 and S2 in the SI. The EDS elemental mapping images clearly show that the La, Ti, O, and N elements are uniformly distributed throughout the protonated oxide, nitrided, and reoxidized crystals, implying the absence of any impurities. The EDS spectra shown in Figure S2 in the SI confirm that the La and Ti elements are present with nearly stoichiometric

Figure 10. Reaction time courses for O2 and N2 evolution from aqueous AgNO3 solution over the protonated oxide (a), nitrided (b), and reoxidized K2La2Ti3O10 crystals (c). 237

DOI: 10.1021/acssuschemeng.6b01344 ACS Sustainable Chem. Eng. 2017, 5, 232−240

Research Article

ACS Sustainable Chemistry & Engineering

photocatalytic water oxidation (0.9 μmol) and the photocatalytic decomposition of N2O (1.8 μmol). It was already reported that the LaTiO2N crystals with a visible-light absorption edge of 600 nm could not show a high photocatalytic activity under the near band gap light irradiation (λ = 580 nm).47 Thus, the protonated reoxidized K2La2Ti3O10 crystals with an absorption edge of 460 nm could not show high amount of evolved O2 gas from the photocatalytic water oxidation half-reaction under the light irradiation used in this study (λ > 420 nm). From the above results, it is clear that only the protonated reoxidized K2La2Ti3O10 crystals are active during the photocatalytic water oxidation half-reaction because of the modified band structure (Eg = 2.70 eV) and less defects associated with reduced titanium species and anion vacancy.

crystals. After 5 h of photocatalytic water oxidation halfreactions with the protonated oxide, nitrided, and reoxidized K2La2Ti3O10 crystals, the total amounts of evolved gases are 21.5, 16.7, and 2.7 μmol for O2 gas and 40.6, 33.8, and 3.6 μmol for N2 gas. Even though the protonated oxide K2La2Ti3O10 crystals with an absorption edge wavelength at about 364 nm could not act as a photocatalyst under the present conditions, and the protonated nitrided K2La2Ti3O10 crystals having the greater amount of reduced titanium species (Ti3+), which act as the recombination center for the photogenerated charge carriers, should have lower photocatalytic activity,42 they have exhibited the O2 and N2 gas evolution after 5 h of the reactions. First, the partial photolysis of AgNO3 occurred immediately, and silver nanoparticles were formed on the surfaces of the protonated crystals under light irradiation because AgNO3 has a photosensitivity.43 According to the previous report, the TiO2 crystals coated with silver nanoparticles have exhibited the narrower bang gap energy (2.75 eV) compared to the pure TiO 2 crystals. 44 Therefore, the protonated oxide Ag/ H2La2Ti3O10 is also expected to show a narrower band gap energy. Furthermore, the photocatalytic activity of the protonated nitrided K2La2Ti3O10 crystals with silver nanoparticles can be improved due to the effective charge separation between H2La2Ti3O10−3/2xNx and Ag nanoparticles.45 Interestingly, the ratio of the evolved N2/O2 gases for the protonated oxide and protonated nitrided K2La2Ti3O10 crystals were found to be 1.9 and 2.0, respectively. It is known that the photocatalytic decomposition of nitrous oxide (N2O) can take place in the presence of silver.46,47 From the XPS results, it was confirmed that the adsorbed NOx species existed on the surfaces of all the protonated crystals. Therefore, it is believed that regarding the protonated oxide and nitrided K2La2Ti3O10 crystals, the entire reactions can mainly be assigned as the photocatalytic decomposition of N2O, as given by the following reaction:



CONCLUSIONS In summary, protonated lanthanum titanium oxide H2La2Ti3O10 and oxynitride H2La2Ti3O10−3/2xNx crystals were synthesized from the oxide, nitrided, and reoxidized layered K2La2Ti3O10 crystals. Although partial nitridation and reoxidation resulted in a noticeable change in absorption wavelength and band gap energy of K2La2Ti3O10 crystals, their crystal structure and morphology were maintained. Also, the oxide [La2Ti3O10]2− and oxynitride [La2Ti3O10−3/2xNx]2− nanosheets were fabricated from the oxide, nitrided, and reoxidized K2La2Ti3O10 crystals through proton exchange and mechanical exfoliation. Among three protonated samples (oxide, nitrided, and reoxidized), the protonated reoxidized K2La2Ti3O10 crystals exhibited high stability and O2 evolution rate of 180 nmol·h−1 after 5 h of the photocatalytic water oxidation half-reaction. Unexpectedly, the protonated oxide and protonated nitrided K2La2Ti3O10 crystals did not exhibit any photocatalytic activity for O2 evolution due to the photolysis of AgNO3 and photodecomposition of N2O during the reaction. Introducing nitrogen into the crystal lattice and partial reoxidizing of Ti3+ were found to be effective for enhancing the photocatalytic water oxidation activity of the layered K2La2Ti3O10 crystals.

2N2O → 2N2↑ + O2 ↑

After 30 min of the reaction, the O2 and N2 gas evolution rate gradually decreased due possibly to the coverage of the surfaces of the photocatalysts with silver nanoparticles with increasing the reaction time.46 In contrast, the ratio of the evolved N2/O2 gases for the protonated reoxidized K2La2Ti3O10 crystals was 1.3, indicating that its gas evolution process is different from those of the protonated oxide and protonated nitrided K2La2Ti3O10 crystals. As the protonated reoxidized K2La2Ti3O10 crystals have the higher absorption edge wavelength at about 460 nm and lesser amount of Ti3+, the photocatalytic O2 evolution reaction for this sample can be expressed as follows:



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.6b01344. EDS element mapping images, EDS spectra, and XPS core-level spectra of protonated oxide, nitrided, and reoxidized K2La2Ti3O10 crystals. (PDF)



4AgNO3 + 2H 2O → 4Ag + O2 ↑ + 4HNO3

Meantime, the partial photolysis of AgNO3 and the photocatalytic decomposition of N2O are also expected to occur. However, the amounts of evolved O2 and N2 gases of the protonated reoxidized K2La2Ti3O10 crystals were much lower than those of the protonated oxide and protonated nitrided K2La2Ti3O10 crystals. This difference is presumably related to the deposition of more silver nanoparticles on the photocatalyst surfaces in the early period of the reaction owing to the acceleration of the deposition rate of silver nanoparticles by the photocatalytic water oxidation reaction. However, the total amount of evolved O2 gas (2.7 μmol) of the protonated reoxidized K2La2Ti3O10 crystals was contributed by the

AUTHOR INFORMATION

Corresponding Author

*Phone: +81-26-269-5556. Fax: +81-26-269-5550. E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This research was supported in part by the Japan Technological Research Association of Artificial Photosynthetic Chemical Process (ARPChem). 238

DOI: 10.1021/acssuschemeng.6b01344 ACS Sustainable Chem. Eng. 2017, 5, 232−240

Research Article

ACS Sustainable Chemistry & Engineering



(21) Cui, W.; Guo, D.; Liu, L.; Hu, J.; Rana, D.; Liang, Y. Preparation of ZnIn2S4/K2La2Ti3O10 composites and their photocatalytic H2 evolution from aqueous Na2S/Na2SO3 under visible light irradiation. Catal. Commun. 2014, 48, 55−59. (22) Cui, W.; An, W.; Liu, L.; Hu, J.; Liang, Y. A novel nano-sized BiOBr decorated K2La2Ti3O10 with enhanced photocatalytic properties under visible light. J. Solid State Chem. 2014, 215, 94−101. (23) Maeda, K.; Mallouk, T. E. Comparison of two-and three-layer restacked Dion−Jacobson phase niobate nanosheets as catalysts for photochemical hydrogen evolution. J. Mater. Chem. 2009, 19, 4813− 4818. (24) Huang, Y.; Wu, J.; Li, T.; Hao, S.; Lin, J. Synthesis and photocatalytic properties of H2La2Ti3O10/TiO2 intercalated nanomaterial. J. Porous Mater. 2006, 13, 55−59. (25) Wu, J.; Huang, Y.; Li, T.; Lin, J.; Huang, M.; Wei, Y. Synthesis and photocatalytic properties of layered nanocomposite H2La2Ti3O10/ Fe2O3. Scr. Mater. 2006, 54, 1357−1362. (26) Zhang, L.; Zhang, W.; Lu, L.; Yang, X.; Wang, X. Synthesis, structure and photocatalytic reactivity of layered CdS/H2La2Ti3O10 nanocomposites. J. Mater. Sci. 2006, 41, 3917−3921. (27) Sato, M.; Toda, K.; Watanabe, J.; Uematsu, K. Structure determination and silver ion conductivity of layered perovskite compounds M2La2Ti3O10 (M = K and Ag). J. Chem. Soc. Jpn. 1993, 5, 640−646. (28) Wang, M.; Ren, F.; Zhou, J.; Cai, G.; Cai, L.; Hu, Y.; Wang, D.; Liu, Y.; Guo, L.; Shen, S. N doping to ZnO nanorods for photoelectrochemical water splitting under visible light: engineered impurity distribution and terraced band structure. Sci. Rep. 2015, 5, 12925. (29) Maeda, K.; Takata, T.; Domen, K. (Oxy)nitrides and Oxysulfides as Visible-Light-Driven Photocatalysts for Overall Water Splitting. In Energy Efficiency and Renewable Energy through Nanotechnology, Green Energy and Technology; Zang, L., Ed.; Springer Verlag: London, 2011; pp 487−529. (30) Irie, H.; Watanabe, Y.; Hashimoto, K. Nitrogen-concentration dependence on photocatalytic activity of TiO2−xNx powders. J. Phys. Chem. B 2003, 107, 5483−5486. (31) Kawashima, K.; Hojamberdiev, M.; Wagata, H.; Yubuta, K.; Oishi, S.; Teshima, K. Chloride flux growth of La2TiO5 crystals and nontopotactic solid-state transformation to LaTiO2N crystals by nitridation using NH3. Cryst. Growth Des. 2015, 15, 333−339. (32) Tessier, F.; Le Gendre, L.; Cheviré, F.; Marchand, R.; Navrotsky, A. Thermochemistry of a new class of materials containing dinitrogen pairs in an oxide matrix. Chem. Mater. 2005, 17, 3570− 3574. (33) Missyul, A. B.; Zvereva, I. A.; Palstra, T. T. M. The formation of the complex manganites LnSr2Mn2O7 (Ln = La, Nd, Gd). Mater. Res. Bull. 2012, 47, 4156−4160. (34) Gedanken, A. Using sonochemistry for the fabrication of nanomaterials. Ultrason. Sonochem. 2004, 11, 47−55. (35) Yuan, Y. J.; Yu, Z. T.; Liu, X. J.; Cai, J. G.; Guan, Z. J.; Zou, Z. G. Hydrogen photogeneration promoted by efficient electron transfer from iridium sensitizers to colloidal MoS2 catalysts. Sci. Rep. 2014, 4, 4050. (36) Sunding, M. F.; Hadidi, K.; Diplas, S.; Løvvik, O. M.; Norby, T. E.; Gunnæs, A. E. XPS characterisation of in situ treated lanthanum oxide and hydroxide using tailored charge referencing and peak fitting procedures. J. Electron Spectrosc. Relat. Phenom. 2011, 184, 399−409. (37) Maegli, A. E.; Sagarna, L.; Populoh, S.; Penkala, B.; Otal, E. H.; Weidenkaff, A. Optical and transport properties of LaTi1−xMx(O,N)3±δ (x = 0; 0.1, M = Nb5+, W6+) thin films prepared by plasma ammonolysis. J. Solid State Chem. 2014, 211, 106−112. (38) Shan, Z.; Archana, P. S.; Shen, G.; Gupta, A.; Bakker, M. G.; Pan, S. NanoCOT: low-cost nanostructured electrode containing carbon, oxygen, and titanium for efficient oxygen evolution reaction. J. Am. Chem. Soc. 2015, 137, 11996−12005. (39) Kim, Y. K.; Park, S.; Kim, K. J.; Kim, B. Photoemission study of N-doped TiO2 (110) with NH3. J. Phys. Chem. C 2011, 115, 18618− 18624.

REFERENCES

(1) Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37−38. (2) Ikeda, S.; Hara, M.; Kondo, J. N.; Domen, K.; Takahashi, H.; Okubo, T.; Kakihana, M. Preparation of K2La2Ti3O10 by polymerized complex method and photocatalytic decomposition of water. Chem. Mater. 1998, 10, 72−77. (3) Li, Y.; Chen, G.; Zhang, H.; Lv, Z. Band structure and photocatalytic activities for H2 production of ABi2Nb2O9 (A = Ca, Sr, Ba). Int. J. Hydrogen Energy 2010, 35, 2652−2656. (4) Machida, M.; Yabunaka, J.; Kijima, T. Synthesis and photocatalytic property of layered perovskite tantalates, RbLnTa2O7 (Ln = La, Pr, Nd, and Sm). Chem. Mater. 2000, 12, 812−817. (5) Li, Y.; Chen, G.; Zhang, H.; Li, Z.; Sun, J. Electronic structure and photocatalytic properties of ABi2Ta2O9 (A = Ca, Sr, Ba). J. Solid State Chem. 2008, 181, 2653−2659. (6) Takata, T.; Shinohara, K.; Tanaka, A.; Hara, M.; Kondo, J. N.; Domen, K. A highly active photocatalyst for overall water splitting with a hydrated layered perovskite structure. J. Photochem. Photobiol., A 1997, 106, 45−49. (7) Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272−1276. (8) Toda, K.; Watanabe, J.; Sato, M. Crystal structure determination of ion-exchangeable layered perovskite compounds, K2La2Ti3O10 and Li2La2Ti3O10. Mater. Res. Bull. 1996, 31, 1427−1435. (9) Gönen, Z. S.; Paluchowski, D.; Zavalij, P.; Eichhorn, B. W.; Gopalakrishnan, J. Reversible cation/anion extraction from K2La2Ti3O10: formation of new layered titanates, KLa2Ti3O9.5 and La2Ti3O9. Inorg. Chem. 2006, 45, 8736−8742. (10) Kudo, A.; Sakata, T. luminescent properties of nondoped and rare earth metal ion-doped K2La2Ti3O10 with layered perovskite structures: importance of the hole trap process. J. Phys. Chem. 1995, 99, 15963−15967. (11) Taniguchi, T.; Murakami, T.; Funatsu, A.; Hatakeyama, K.; Koinuma, M.; Matsumoto, Y. Reversibly tunable upconversion luminescence by host−guest chemistry. Inorg. Chem. 2014, 53, 9151−9155. (12) Thaminimulla, C. T. K.; Takata, T.; Hara, M.; Kondo, J. N.; Domen, K. Effect of chromium addition for photocatalytic overall water splitting on Ni−K2La2Ti3O10. J. Catal. 2000, 196, 362−365. (13) Matsuoka, M.; Kitano, M.; Takeuchi, M.; Tsujimaru, K.; Anpo, M.; Thomas, J. M. Photocatalysis for new energy production: recent advances in photocatalytic water splitting reactions for hydrogen production. Catal. Today 2007, 122, 51−61. (14) Huang, Y.; Wu, J.; Wei, Y.; Hao, S.; Huang, M.; Lin, J. Synthesis and photocatalytic activity of hydrated layered perovskite K2−xLa2Ti3−xNbxO10 (0 ⩽ x ⩽ 1) and protonated derivatives. Scr. Mater. 2007, 57, 437−440. (15) Ya-Hui, Y.; Qi-Yuan, C.; Zhou-Lan, Y.; Jie, L. Study on the photocatalytic activity of K2La2Ti3O10 doped with zinc (Zn). Appl. Surf. Sci. 2009, 255, 8419−8424. (16) Yang, Y.; Chen, Q.; Yin, Z.; Li, J. Study on the photocatalytic activity of K2La2Ti3O10 doped with vanadium (V). J. Alloys Compd. 2009, 488, 364−369. (17) Wang, B.; Li, C.; Hirabayashi, D.; Suzuki, K. Hydrogen evolution by photocatalytic decomposition of water under ultraviolet− visible irradiation over K2La2Ti3−xMxO10+δ perovskite. Int. J. Hydrogen Energy 2010, 35, 3306−3312. (18) Huang, Y.; Wei, Y.; Cheng, S.; Fan, L.; Li, Y.; Lin, J.; Wu, J. Photocatalytic property of nitrogen-doped layered perovskite K2La2Ti3O10. Sol. Energy Mater. Sol. Cells 2010, 94, 761−766. (19) Kumar, V.; Uma, S.; Govind. Investigation of cation (Sn2+) and anion (N3−) substitution in favor of visible light photocatalytic activity in the layered perovskite K2La2Ti3O10. J. Hazard. Mater. 2011, 189, 502−508. (20) Cui, W.; Liu, L.; Ma, S.; Liang, Y.; Zhang, Z. CdS-sensitized K2La2Ti3O10 composite: a new photocatalyst for hydrogen evolution under visible light irradiation. Catal. Today 2013, 207, 44−49. 239

DOI: 10.1021/acssuschemeng.6b01344 ACS Sustainable Chem. Eng. 2017, 5, 232−240

Research Article

ACS Sustainable Chemistry & Engineering (40) Qiao, Y.; Hu, X.; Liu, Y.; Chen, C.; Xu, H.; Hou, D.; Hu, P.; Huang, Y. Conformal N-doped carbon on nanoporous TiO2 spheres as a high-performance anode material for lithium-ion batteries. J. Mater. Chem. A 2013, 1, 10375−10381. (41) Fettkenhauer, C.; Wang, X.; Kailasam, K.; Antonietti, M.; Dontsova, D. Synthesis of efficient photocatalysts for water oxidation and dye degradation reactions using CoCl2 eutectics. J. Mater. Chem. A 2015, 3, 21227−21232. (42) Irie, H.; Watanabe, Y.; Hashimoto, K. Nitrogen−concentration dependence on photocatalytic activity of TiO2−xNx powders. J. Phys. Chem. B 2003, 107, 5483−5486. (43) Omrani, A. A.; Taghavinia, N. Photo-induced growth of silver nanoparticles using UV sensitivity of cellulose fibers. Appl. Surf. Sci. 2012, 258, 2373−2377. (44) Tunc, I.; Bruns, M.; Gliemann, H.; Grunze, M.; Koelsch, P. Bandgap determination and charge separation in Ag@TiO2 core shell nanoparticle films. Surf. Interface Anal. 2010, 42, 835−841. (45) Hu, H.; Ding, J.; Zhang, S.; Li, Y.; Bai, L.; Yuan, N. Photodeposition of Ag2S on TiO2 nanorod arrays for quantum dotsensitized solar cells. Nanoscale Res. Lett. 2013, 8, 10. (46) Sano, T.; Negishi, N.; Mas, D.; Takeuchi, K. Photocatalytic decomposition of N2O on highly dispersed Ag+ ions on TiO2 prepared by photodeposition. J. Catal. 2000, 194, 71−79. (47) Singh, R. B.; Matsuzaki, H.; Suzuki, Y.; Seki, K.; Minegishi, T.; Hisatomi, T.; Domen, K.; Furube, A. Trapped state sensitive kinetics in LaTiO2N solid photocatalyst with and without cocatalyst loading. J. Am. Chem. Soc. 2014, 136, 17324−17331.

240

DOI: 10.1021/acssuschemeng.6b01344 ACS Sustainable Chem. Eng. 2017, 5, 232−240