Aditi
Vijay
,
Kadambari
Bairagi
and
Sonalika
Vaidya
*
Institute of Nano Science and Technology, Knowledge City, Sector 81, Sahibzada Ajit Singh Nagar, Punjab 140306, India. E-mail: svaidya@inst.ac.in
First published on 26th April 2022
This study focuses on relating the structure of perovskite oxides with their properties and activities and provides a comparative study of the three members of the Sr–Ti–O system for photocatalytic hydrogen evolution. The three oxides focused on in this study are based on perovskite structure viz. SrTiO3 and SrO–(SrTiO3)n (n = 1 and 2). We have successfully synthesized these three oxides through a methodology that combined the polymeric citrate precursor method with the hydrothermal method. Their crystal structure, morphology, and optical properties (absorption and photoluminescence) were systematically explored. SrTiO3 belonged to a class of cubic perovskite while Sr2TiO4 (n = 1) and Sr3Ti2O7 (n = 2) belonged to layered Ruddlesden–Popper based perovskite oxides. We observed the cube-shaped morphology for nanostructured SrTiO3 and layered morphology for Ruddlesden–Popper based oxides, Sr2TiO4 and Sr3Ti2O7. The photocatalytic hydrogen evolution performance of these nanostructured oxides was investigated. Amongst the three nanostructured oxides, the maximum amount of hydrogen was evolved with Sr3Ti2O7 as the photocatalyst. These results were supported by photoluminescence, time-resolved photoluminescence, and photoelectrochemical studies.
Over the past few decades, several semiconductor oxides such as TiO2, ZnO, AgGaO2, and BiVO4 have been discovered as efficient catalysts for photocatalytic hydrogen evolution.8–14 Among all the developed photocatalysts, perovskite structures i.e. ABO3 have gained much popularity because of their composition which can be easily modulated at A and B sites.15–17 Among the various perovskite semiconductor oxides, SrTiO3 is widely studied as a photocatalyst18–20 for its outstanding structural stability and compositional flexibility. Ruddlesden Popper-based layered perovskites, with a general formula of AO–(ABO3)n, (which can be also written as (An+1BO3n+1) or A′2[An−1BnO3n+1]) have attracted much attention in the field of photocatalytic water splitting due to their physical properties which can be modified by intercalation and ion-exchange.21,22 These structures consist of a large number of active sites such as the B site and the AO layer site to facilitate the reaction. Several metal ions can be doped23,24 at the A and B sites in these structures which modulates the bandgap of the oxide towards the visible region thereby enhancing their photocatalytic activity.
Fig. 1 shows the crystal structure of SrTiO3 and SrO–(SrTiO3)n (n = 1 and 2). SrTiO3 belongs to the class of perovskite that has a cubic structure composed of corner shared TiO6 octahedra with Sr, present in the holes formed from the cuboctahedron symmetry. When n = 1 and 2, the structure belongs to a class of oxide known as Ruddlesden–Popper phases. The two-lower symmetry 2D-Ruddlesden–Popper oxides, Sr2TiO4 (n = 1) and Sr3Ti2O7 (n = 2) have body-centered tetragonal symmetry. These structures are composed of stacked nSrTiO3 perovskite layers separated by a SrO rock-salt type layer, as shown in Fig. 1.
Fig. 1 Schematic showcasing the crystal structure of SrTiO3 and SrO–(SrTiO3)n for n = 1 (Sr2TiO4) and n = 2 (Sr3Ti2O7). |
Here, we discuss a comparative study of the performance of nanostructured SrTiO3 and SrO–(SrTiO3)nviz. Sr2TiO4 (n = 1) and Sr3Ti2O7 (n = 2) towards photocatalytic hydrogen evolution reactions. Out of these three perovskite oxides, Sr2TiO4 and Sr3Ti2O7 belong to I4/mmm, tetragonal symmetry whereas SrTiO3 belongs to the Pmm space group. The photocatalytic property of these nanostructured oxides is investigated through hydrogen evolution reactions under UV-visible light irradiation. Photoluminescence, time-resolved photoluminescence, and photoelectrochemical studies have been carried out to see the role of the perovskite structure in influencing photocatalytic activity. To the best of our knowledge, there are no reports on a comparative study of the three kinds of structures on photocatalytic hydrogen evolution performance. However, there is one report25 where the ratio of Sr/Ti in SrTiO3 was varied and their photocatalytic HER was studied for different ratios from 1.00 to 1.25. The photocatalytic hydrogen evolution activity of these three perovskite oxides has been studied separately.26–28 In this study, we have tried to highlight that the choice of structure (crystallographic), amongst the cubic and layered perovskites (showcased here with the Sr–Ti–O system), could be an efficient way for the development of a catalyst for a hydrogen evolution reaction with improved performance.
For the synthesis of SrTiO3, TTIP (12 mmol) was added to a mixture of ethylene glycol (25 mL) and methanol (10 mL). To this aqueous solution of citric acid in 5 mL water and solid Sr(NO3)2 (12 mmol) were added. The amount of citric acid taken was equivalent to that of TTIP. The resultant mixture was heated at 130 °C for 4 hours to form a single-phase transparent solution. Afterward, the pH of this solution was adjusted to 13 by adding 5 M NaOH (5 mL). The resulting solution was transferred into a Teflon vessel followed by hydrothermal treatment at 200 °C for 48 h. After the Teflon vessel was cooled down to room temperature, the obtained gel was vacuum dried at 200 °C for 16 h. The products were washed several times with glacial acetic acid (to remove carbonate present as an impurity in the sample), deionized (DI) water, and ethanol, and then dried overnight at 70 °C. SrTiO3 was finally synthesized by calcining the dried powder at 750 °C for 5 h.
For the synthesis of Sr2TiO4 and Sr3Ti2O7, the exact procedure was followed as for SrTiO3 except for the ratio of Sr and Ti precursor and heating temperature. For Sr2TiO4, the ratio of Sr and Ti precursor was kept at 2:1 (12 mmol of Sr(NO3)2 and 6 mmol of TTIP) and calcined first at 650 °C and then at 1000 °C for 12 h, whereas for Sr3Ti2O7 the ratio was kept at 3:2 (9 mmol of Sr(NO3)2 and 6 mmol of TTIP) and calcined first at 650 °C and then at 950 °C for 6 h.
y = A1e(−x/τ1) + A2e(−x/τ2) + y0 | (1) |
(2) |
(3) |
Photoelectrochemical studies were performed on a PGSTAT-30 (Autolab) electrochemical workstation using a standard three-electrode system consisting of Ag/AgCl (3 M KCl) as the reference electrode, platinum wire as the counter electrode, and Sr–Ti–O samples deposited on a glass substrate coated with fluorine-doped tin oxide (FTO) as the working electrode. For the synthesis of the working electrode, the catalyst was drop cast on a glass substrate coated with FTO with an area of 1 cm × 1 cm. For the preparation of the catalyst ink, 10 mg of the catalyst was dispersed in 200 μL of isopropyl alcohol containing 10 μL of Nafion resin solution through the ultra-sonication method. Here, 0.1 M Na2SO4 (pH = 7) was taken as an electrolyte, and saturated with argon for 30 minutes to remove dissolved oxygen. A 350 W Xe lamp was used as the source of irradiation. LSV (linear sweep voltammetry) curves were obtained in the range from 0 to −1 V vs. Ag/AgCl at a scan rate of 10 mV s−1 under the light. Electrochemical impedance spectroscopy (EIS) was carried out at −0.35 V vs. Ag/AgCl in the frequency range from 0.1 Hz to 100 kHz under light.
Raman spectra of the three nanostructured oxides i.e. SrTiO3, Sr2TiO4, and Sr3Ti2O7 are shown in Fig. 3. The Raman spectra obtained for the perovskite SrTiO3 matched well with the previous reports.29,30 The spectra consist of a low-frequency band present at 77 cm−1 which was assigned to doubly degenerate modes, Eg. Second order Raman bands were also observed between 200–400 cm−1 and 600–800 cm−1. No first-order bands were observed in the Raman spectra, as expected for the cubic structure. According to the report by Nilsen et al.29 second-order Raman scattering is due to the creation and destruction of two phonons which can originate from anywhere in the Brillouin zone. The authors observed a second-order band or overtone for SrTiO3 at 369 cm−1 which was attributed to the combination of various bands including the TO4–TA, TO4–TOl, and 2TO2 bands whereas the band at 684 cm−1 was assigned to the 2TO3 overtone. It has been previously reported31 that there are four Raman active modes, A1g, and Eg, observed for layered Ruddlesden–Popper oxides with n = 1, Sr2TiO4. The Raman bands present at 121, 203, 182, and 571 cm−1 were assigned to the Eg, A1g, Eg, and A1g modes respectively (Fig. 3). The broader band observed between 400–450 cm−1 and around 700 cm−1 indicates the presence of a second-order band or defect-induced excitations which may originate at the oxygen sublattice. The Raman spectra of Sr3Ti2O7 (Fig. 3) show bands at 178, 198, 274, 500, and 633 cm−1 which can be assigned to the A1g, Eg, Eg, A1g, and A1g modes respectively which is consistent with a previous report.32 A defect-induced excitation was also observed at 92 cm−1. The band corresponding to the A1g mode was observed to be more intense in both Sr2TiO4 and Sr3Ti2O7, which could be due to symmetric stretching of the oxygen lattice.33
The high resolution XPS spectra of strontium (Sr(3d), Fig. S1a–c, ESI†), oxygen (O(1s), Fig. S2a–c, ESI†) and titanium, (Ti(2p), Fig. S3a–c, ESI†), was obtained for SrTiO3 Sr2TiO4 and Sr3Ti2O7. Peaks at binding energies of 133.7 eV (135.5 eV), 133.9 eV (135.5 eV) and 133.6 eV (135.4 eV) were observed in the Sr 3d5/2 (3d3/2) spectra of SrTiO3 (Fig. S1a, ESI†), Sr2TiO4 (Fig. S1b, ESI†) and Sr3Ti2O7 (Fig. S1c, ESI†) respectively. Fig. S2a–c (ESI†) show the high-resolution O 1s spectra for SrTiO3, Sr2TiO4, and Sr3Ti2O7. The peaks were fitted with two Gaussian peaks with binding energies at 529.7 and 531.2 eV for SrTiO3, 529.3 and 531.4 eV for Sr2TiO4 and 529.8 and 531.77 eV for Sr3Ti2O7. The peak at lower energy can be attributed to the metal–oxygen bond i.e. the presence of O2− ions in the crystal structure whereas the peak at higher energy can be related to the oxygen vacancies. Fig. S3a–c (ESI†) shows the high-resolution spectra of Ti(2p) for SrTiO3, Sr2TiO4, and Sr3Ti2O7 respectively. The presence of Ti3+ was observed along with the Ti4+ ion after fitting of the peaks. The peaks centered at 458.3 eV (464.2 eV), 458.1 eV (453.8 eV) and 458.5 eV (464.1 eV) were observed for the Ti4+ 2p3/2 (2p1/2) spectra of SrTiO3, Sr2TiO4 and Sr3Ti2O7 respectively. Ti3+ 2p3/2 (2p1/2) spectra of SrTiO3, Sr2TiO4 and Sr3Ti2O7 were obtained at 456.4 eV (462.8 eV), 456.8 eV (462.1 eV) and 456.6 eV (461.9 eV) respectively. Thus the presence of peaks corresponding to Ti3+ also confirms the presence of an oxygen vacancy in the lattice. Based on the area under the peak, the order for the ratio of oxygen vacancies:M–O and Ti3+:Ti4+ was found to be Sr2TiO4 > Sr3Ti2O7 > SrTiO3.
TEM studies for SrTiO3 showed nanocubes with a size of 80–100 nm (Fig. 4a). The HRTEM image (Fig. 4b) shows lattice fringes with a spacing of 0.281 nm, corresponding to the (110) plane. A rectangular sheet-like morphology was observed for both Ruddlesden–Popper based oxides i.e. Sr2TiO4 (Fig. 4c) and Sr3Ti2O7 (Fig. 4e). The size along one dimension of these sheets was observed to be ∼250 nm and ∼350 nm for Sr2TiO4 and Sr3Ti2O7 respectively. The lattice fringes corresponding to a spacing of 0.276 nm for Sr2TiO4 (Fig. 4d) and 0.273 nm for Sr3Ti2O7 (Fig. 4f) were observed which corresponded to the (110) plane. The BET surface area of all the samples viz. SrTiO3, Sr2TiO4, and Sr3Ti2O7 were observed to be 15 m2 g−1, 6.2 m2 g−1, and 14 m2 g−1 respectively.
Fig. 4 TEM images of (a) SrTiO3, (c) Sr2TiO4, and (e) Sr3Ti2O7. HRTEM images of (b) SrTiO3, (d) Sr2TiO4, and (f) Sr3Ti2O7. The inset shows the corresponding reduced FFT of the HRTEM image. |
The bandgap of the oxides was calculated using the Tauc equation (eqn (4)).
(4) |
Fig. 5 (a) Tauc plot and (b) schematic showcasing the energy diagram for SrTiO3, Sr2TiO4, and Sr3Ti2O7. |
Sample | n (amount of hydrogen gas evolved after 5 h of the reaction) (mol) | Apparent quantum yield (AQY) (%) |
---|---|---|
SrTiO3 | 12.23 × 10−6 | 0.018 |
Sr2TiO4 (SrO(SrTiO3)) | 22.6 × 10−6 | 0.03 |
Sr3Ti2O7 (SrO(SrTiO3)2) | 36.5 × 10−6 | 0.058 |
To ascertain the reason for the observed trend for the photocatalytic hydrogen evolution, photoluminescence (PL) and time-resolved photoluminescence studies were carried out for the three oxides. Fig. 7 shows the photoluminescence spectra for all the samples. Here, the samples were excited at a wavelength of 380 nm, and the emission spectra were recorded in the range of 395–700 nm. All the samples showed PL emission in the violet-blue region with a peak centered at 414 nm and 436 nm. It has been reported37 that the photoluminescence for SrTiO3 based oxides arises mostly due to a recombination of the electrons and holes that are trapped in the intermediate states (present within the bandgap). These intermediate states arise as a result of distortion, oxygen vacancies, etc. It has also been reported37 that the emission in the violet-blue region occurs due to the presence of a surface or shallow defects which may arise due to oxygen vacancies. Oxygen vacancies are also known to affect the catalytic behavior of oxides.5,38 The presence of a controlled concentration of defects (oxygen vacancies) is known to increase the photocatalytic efficiency of SrTiO3 towards hydrogen evolution.39 The oxygen vacancies act as electron donors which either result in an increased charge transport or a shift in the Fermi level towards the conduction band.39 Such a phenomenon is likely to improve the charge separation behavior of the oxide. In our studies, the presence of defect/oxygen vacancies was also confirmed by Raman and XPS studies. In addition to the presence of defects, it was also observed that the PL emission, corresponding to the defect, decreased in the following order SrTiO3 > Sr2TiO4 > Sr3Ti2O7. With the same excitation energy and no significant changes in the optical bandgap of the oxides, the decrease in the PL emission can be related to a decrease in the recombination of electron and hole pairs giving rise to radiative emission. To further investigate the lifetime of photo-induced charge carriers, time-resolved photoluminescence decay spectra were recorded. The data was fitted using eqn (1) and the average lifetime (τavg) was calculated using eqn (2). Fig. 8a–c show the second exponential decay fit of the time-resolved PL of the three oxides. The parameters obtained after fitting are listed in Table 2. The average lifetime was found to follow the order Sr3Ti2O7 > Sr2TiO4 > SrTiO3, for the Sr–Ti–O system. From the values obtained for τavg (ns), it can be concluded that the recombination of photo-induced charge carriers is delayed for Ruddlesden–Popper based oxides (Sr2TiO4 and Sr3Ti2O7) compared with SrTiO3. The longer lifetime implies that a large number of photo-induced electrons could reach the surface of the catalysts which would be captured by the H+ ions. Thus, the longer lifetime of the photo-induced charge carriers for Sr3Ti2O7 further supports the increased photocatalytic properties of Sr3Ti2O7. The presence of interlayers (SrO layer in our study) results in a reduction of the recombination of photogenerated charge carriers by separation of the electrons and holes thereby enhancing the photocatalytic water splitting reaction.40
Sample | A 1 | τ 1 (ns) | A 2 | τ 2 (ns) | Average lifetime τavg (ns) |
---|---|---|---|---|---|
SrTiO3 | 1800 | 37.7 ± 1.0 | 213 | 351 ± 8.6 | 202 |
Sr2TiO4 | 1646 | 38.0 ± 1.2 | 198 | 367 ± 8.8 | 215 |
Sr3Ti2O7 | 1619 | 40.3 ± 1.0 | 194 | 392 ± 10 | 230 |
To further evaluate the charge transfer behavior of the three oxides in the presence of light, photoelectrochemical studies were carried out. The photocurrent density (Fig. 9a) was found to be the highest for Sr3Ti2O7. The onset potential was observed to be ∼−0.40 V vs. Ag/AgCl, which is nearly the same for all three samples. The overpotential for the hydrogen evolution reaction (HER) for the three oxides was evaluated from the current–voltage curve at −0.1 mA cm−2vs. Ag/AgCl (Fig. 9b). The overpotential was found to follow the order SrTiO3 > Sr2TiO4 > Sr3Ti2O7 for the Sr–Ti–O system. Fig. 9c shows Nyquist plots for the three oxides. Based on Rct (charge transfer resistance), Rs (solution resistance), and a constant phase element with impedance, which is related to the angular frequency of the applied potential, ω (using eqn (5)),
(5) |
(6) |
(7) |
Fig. 9 (a) Photocurrent density, (b) overpotential, (c) Nyquist plot and (d) Mott–Schottky plot for SrTiO3, Sr2TiO4, and Sr3Ti2O7. |
Sample | R ct (kΩ) | R s (Ω) | CPE (μmho) | n (in constant phase element) |
---|---|---|---|---|
SrTiO3 | 20.6 | 36.5 | 16.0 | 0.940 |
Sr2TiO4 | 14.5 | 28.4 | 12.5 | 0.973 |
Sr3Ti2O7 | 12.3 | 26.5 | 12.1 | 0.974 |
Thus, from the trend observed for the photocatalytic hydrogen evolution there are a few interesting observations. One, the photocatalytic hydrogen evolution increases with the introduction of the SrO layer i.e. when comparing SrTiO3 with Sr2TiO4 (SrO(SrTiO3)). The SrO layer in Ruddlesden–Popper based layered perovskites44 is responsible for the dissociation of water. The oxygen site (apical oxygen) in SrTiO3 of Ruddlesden–Popper based layered perovskites favors the adsorption of hydrogen enabling its recombination with other adsorbed hydrogen to form H2. Wei et al.45 also observed that insertion of the SrO layer in SrTaO2N to form Sr2TaO3N (a Ruddlesden–Popper based oxynitride) resulting in the improved photocatalytic performance of the oxynitride. Second, with the increase in the SrTiO3 perovskite unit in Ruddlesden–Popper based layered perovskites i.e. on comparing Sr2TiO4 and Sr3Ti2O7, the photocatalytic activity was found to increase for Sr3Ti2O7. For photocatalytic water splitting, the active sites in Ruddlesden–Popper based layered perovskites are the B-site cations (Ti in Sr–Ti–O based systems), AO layers (SrO layer in Sr–Ti–O based systems), and defects arising due to oxygen vacancies. This arrangement of layers is known to suppress the charge carrier recombination, as evident from time-resolved photoluminescence studies, and promote charge carrier transfer, as evident from a decrease in charge transfer resistance.
Based on the study, it is also to be noted there are two important synergistic factors that affect the photocatalytic behavior of the oxides: (a) crystal structure and the arrangement of atoms, and (b) the presence of defects (oxygen vacancies). Thus, synergism in the role of the SrO layer and the SrTiO3 perovskite unit along with the layered morphology, low crystallite size, and presence of defects (oxygen vacancies), is presumed to have resulted in an improved photocatalytic performance for Ruddlesden–Popper based layered perovskite. This improved performance was evident from the reduced overpotential, low charge transfer resistance, and high charge carrier density for Ruddlesden–Popper based layered perovskites. Hence, the appropriate choice of the crystal structure from a series could result in attaining the desired factors required for designing an efficient catalyst for photocatalytic HER.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ma00097k |
This journal is © The Royal Society of Chemistry 2022 |