Structural evolution of alkaline earth metal stannates MSnO3 (M = Ca, Sr, and Ba) photocatalysts for hydrogen production

Fulan Zhonga, Huaqiang Zhuangb, Quan Guc and Jinlin Long*b
aNational Engineering Research Center of Chemical Fertilizer Catalyst, Fuzhou University, Fuzhou 350002, P. R. China
bState Key Laboratory of Photocatalysis on Energy and Environment, School of Chemistry, Fuzhou University, Fuzhou 350116, P. R. China. E-mail: jllong@fzu.edu.cn; Tel: +86-591-83731234-6504
cKey Laboratory of Applied Surface and Colloid Chemistry, Ministry of Education, School of Chemistry and Chemical Engineering, Shaanxi Normal University, Xi'an, 710062, People's Republic of China

Received 2nd March 2016 , Accepted 21st April 2016

First published on 25th April 2016


Abstract

The alkaline earth metal stannates MSnO3 (M = Ca, Sr, and Ba) photocatalysts with different morphologies are successfully prepared by hydrothermal method and their photocatalytic activities are evaluated by photocatalytic reforming of ethanol/water solution to hydrogen. All of the as-prepared samples are characterized in detail by X-ray diffraction (XRD), ultraviolet-visible diffuse reflectance (UV-vis DRS), N2 physical adsorption, transmission electron microscopy (TEM), scanning electron microscopy (SEM), and X-ray photoelectron spectroscopy (XPS) before and after the photocatalytic hydrogen production to illustrate the effect of the photoreaction on the surface structure, and thus to make the photocatalysis clear. The results reveal that the greatest photocorrosion occurs on the surface of CaSnO3, SrSnO3, and BaSnO3 samples. And the formed surface species have great influence on H2 production from ethanol/water solution. The photocatalytic reaction can transform CaSnO3 into CaSn(OH)6, producing CaSn(OH)6/CaSnO3 composite where the photogenerated charges can be more efficiently separated and transferred, consequently enhancing the hydrogen evolution. As for SrSnO3, the photocorrosion can cause the formation of Sn2+ self-doped SrSnO3 nanoparticles on the surface to increase the hydrogen production efficiency. Unlike CaSnO3 and SrSnO3, the photocatalytic activity of BaSnO3 is gradually decreased due to the conversion of BaSnO3 to BaCO3. As expected, the H2 evolution rate decreased in the order of CaSnO3 > SrSnO3 > BaSnO3 under UV light irradiation. It is well demonstrated in the present work that CaSnO3 is a potential photocatalyst for the photocatalytic reforming of ethanol/water solution to hydrogen.


1. Introduction

Photocatalytic hydrogen production is an ideal technology to solve the current energetic and environmental problems. At present, a number of studies have focused on the development of photocatalysts that can work for the reaction. Among the various photocatalysts, semiconductor materials used for photocatalytic hydrogen production have been most intensively studied because of their unique electrical behavior. However, the reaction efficiency doesn't meet the requirements of practical application, owing to two key factors: (1) the light absorption efficiency of the developed inorganic and organic semiconductor is low and cannot satisfy the actual requirements. (2) The interaction between light and solid catalyst is unclear. Semiconductor materials used to the chemical process must satisfy two requirements in theory. Firstly, the bandgap energy is greater than 1.23 eV. Secondly, the band structure must be compatible with the reduction and oxidation potentials of water. In addition to the thermodynamic factors, the kinetic factors concluding the over-potential of hydrogen release and the separation and migration of charge carriers, etc. should be considered for design and development of highly efficient photocatalysts.

A large number of semiconductor materials were already found to be active for the hydrogen production under UV irradiation, mainly including metal-oxide semiconductors based on d0 (such as Ti, Zr, W, and Ta)1–4 and d10 (such as In, Ga, Ge, and Sn),5–8 cerium oxide based semiconductor,9 and non-oxide semiconductors such as ZnS, AgBr, and InP.10–12 A majority of tin-based oxides,7,13–16 such as ZnSnO3, CaSnO3, SrSnO3, and BaSnO3, were proved to be the active photocatalysts due to the strong oxidation and reduction ability at the same time, on the basis of three points as follows: (1) they have a large band-gap energy. (2) The conduction band position is more negative than the reduction potential of hydrogen. (3) The valence band position is more positive than the oxidation potential of the majority of the organic pollutants. Thus, the band structure of the alkaline earths metal stannates MSnO3 (M = Ca, Sr and Ba) completely satisfies the conditions of the hydrogen evolution and thus have received extensive attention. Chen et al.15 synthesized the dumbbell and rod-shaped SrSnO3, and investigated the photocatalytic activity of the water reduction. Omeiri et al.17 successfully synthesized BaSnO3−δ photocatalyst, and studied the activity of hydrogen production under visible light. Zhang et al. reported16 that alkaline earth metal stannates MSnO3 had a good activity for hydrogen production with UV light, and the photocatalytic activity followed the order of CaSnO3 > SrSnO3 > BaSnO3. The results showed that the activity sequence was consistent with the band gap size, and the difference of two potential values between the location of valence band (VBCaSnO3 > VBSrSnO3 >VBBaSnO3) and the reduction potential of H+/H2 of alkaline earth metal stannates MSnO3. However, the alkaline earth metal stannates MSnO3 photocatalysts often suffer from the great surface photocorrosion under UV irradiation, leading to the change in surface structure of catalyst. It may be one of important parameters making the photocatalytic activity different. So, it cannot be fully clarifies the nature of photocatalytic hydrogen production if we consider only from the band structure of alkaline earth metal stannates. A complete research of the evolution of the surface structure is necessary for the in-depth and comprehensive understanding of the photocatalysis of alkaline earth metal stannates MSnO3.

In this work, the alkaline earth metal stannates MSnO3 were synthesized by using the hydrothermal method. The physical and chemical properties of the materials before and after the reaction were characterized by X-ray diffraction (XRD), ultraviolet-visible diffuse reflectance (UV-vis DRS), N2 physical adsorption, transmission electron microscopy (TEM), scanning electron microscopy (SEM), and X-ray photoelectron spectroscopy (XPS). Photocatalytic hydrogen production from ethanol/water solution as a model photoreaction was used to evaluate the photocatalytic properties of the as-synthesized photocatalysts and the corresponding reaction mechanism was also proposed to clearly explain the different photocatalytic activity of alkaline earth metal stannates under the same conditions.

2. Experimental

2.1 Preparation of catalysts

The MSnO3 (M = Ca, Sr, and Ba) photocatalysts were prepared by hydrothermal method. Typically, 3.9 g CaCl2 (or SrCl2·6H2O = 9.4 g, BaCl2·2H2O = 8.6 g) and 12.3 g SnCl4·5H2O were first dissolved into 350 mL deionized water with the stirring speed of 800 rpm at room temperature for 30 min and marked A solution. Thereafter, 16 g NaOH was dissolved into 100 mL deionized water with the stirring speed of 500 rpm at room temperature for 30 min and marked B solution. Thirdly, the B solution was poured into the A solution at the stirring speed of 1000 rpm for 1 h. The obtained white turbid liquid was denoted as C. Finally, 80 mL C was transferred to 100 mL autoclave and heated at 180 °C for 24 h, and then cooled to room temperature. The resulting precipitates were filtered, washed with deionized water, dried, and then calcined at 700 °C for 6 h.

2.2 Catalyst characterizations

The X-ray diffraction (XRD) measurements were performed on a Bruker D8 Advance X-ray diffractometer at 40 kV and 40 mA, using Cu Kα1 radiation (λ = 1.5406 Å). The specific surface areas were determined at 77 K with a micromeritics ASAP 2010 instrument. UV-vis DRS spectra were obtained on a Varian Cary 500 Scan UV-VIS-NIR spectrophotometer using BaSO4 as a reference. XPS spectra were carried out on a VG ESCALAB 250XPS system with a monochromatized Al Kα X-ray source (15 kV 200 W 500 μm, pass energy = 20 eV). All binding energies were referenced to the C 1s peak at 284.6 eV of surface adventitious carbon. Transmission electron microscopy (TEM) images were obtained by a JEOL model JEM 2010 EX instrument at the accelerating voltage of 200 kV. The surface morphologies of the as-prepared catalysts were observed using a field emission scanning electron microscope (FESEM: America, FEI Nova NanoSEM230).

2.3 Photocatalytic activity measurements

The photocatalytic reaction of hydrogen production was performed in a closed gas-recirculation system equipped with an inner irradiation quartz action vessel. In this typical experiment, 0.1 g photocatalyst and 5 mL ethanol used as the sacrificial reagent was suspended in 155 mL of distilled water by magnetic stirring and the temperature of the glass reaction cell was sustained at about 20 °C by circulating water. Prior to the reaction, the system was evacuated by a mechanical pump and then filled with 101 kPa of high-purity Ar (>99.99%). This process was repeated three times in order to completely remove O2 and CO2 from the system. Photocatalyst was dispersed in the ethanol/water by stirring with a magnetic stirrer and irradiated under a 125 W high-pressure Hg lamp with a main wavelength of 365 nm (the light intensity is ca. 17.3 mW cm−2 and the irradiation area is ca. 100 cm2). The temperature of the solution was controlled at 293 ± 0.1 K by circulating condensate water through an interlayer around the reactor. The gas compositions were monitored by a GC 112A equipped with a thermal conduction detector (TCD) and TDX-01 column. The apparent quantum efficiency (AQE) was calculated by the H2 production using the following equation:
 
image file: c6ra05614h-t1.tif(1)

3. Results and discussion

3.1. Characterizations of MSnO3 samples

Fig. 1A represented the XRD patterns of MSnO3 (M = Ca, Sr, and Ba) samples. It was found that the samples consisted of a single phase of CaSnO3, SrSnO3 and BaSnO3 (JCPDS no. 31-0312, 22-1442, and 15-0780) with a perovskite structure, respectively. No other peaks derived from impurities were observed, suggesting the good crystalline for the as-prepared samples.

The UV-visible absorption spectra of the as-prepared MSnO3 samples were showed in Fig. 1B. Three samples displayed a typical absorption feature of semiconductor in the UV region. The band-edge absorption of MSnO3 (M = Ca, Sr, and Ba) samples is positioned at 307, 324, and 407 nm, respectively, with increasing the cationic radius. According to the computing method in literatures,16 the light absorption can be described as follows:

αhν = A(Eg)n/2


image file: c6ra05614h-f1.tif
Fig. 1 (A) XRD patterns and (B) UV-vis DRS over MSnO3 (M = Ca, Sr, and Ba) samples.

Therein, α, , A, and Eg defined the coefficient of absorption, the energy of incident light, the constant, and the width of forbidden band, respectively. CaSnO3 is a direct band gap semiconductor with n = 1. SrSnO3 and BaSnO3 are indirect band gap semiconductors with n = 4. So the corresponding forbidden band widths of MSnO3 (M = Ca, Sr, and Ba) samples are 4.30, 4.10, and 3.12 eV, respectively, as shown in Table 1. The data are different slightly from those reported in literatures.15–17 This may be caused by the preparation method and the crystalline grain. These UV-vis DRS results indicated that the as-prepared MSnO3 (M = Ca, Sr, and Ba) samples prepared under the same hydrothermal conditions have differences in optical absorption properties, which will be closely related to the different photocatalytic activities for H2 production below.

Table 1 The physicochemical characteristics of MSnO3 (M = Ca, Sr, and Ba) samples
  CaSnO3 SrSnO3 BaSnO3
BET (m2 g−1) 13.5 16.9 5.9
Eg (eV) 4.30 4.10 3.12
H2 production rate (μmol h−1) 27.4 4.2 2.4


Therefore, the as-prepared MSnO3 samples were further characterized by SEM, as demonstrated in Fig. 2. The surface morphology of CaSnO3 sample presented inhomogenous cubic structure with 1–4 μm (Fig. 2a and b). The surface morphology of SrSnO3 sample has a nanorod-structure (Fig. 2c and d), while the BaSnO3 sample was ruleless nano-particles (Fig. 2e and f). The results suggested that the processes of the crystal growth for three samples prepared under the same hydrothermal conditions were different, resulting in different morphologies. Thus, we inferred that the surface areas for three samples followed the order of SrSnO3 > CaSnO3 > BaSnO3, as verified by the N2 absorption isotherms. The surface areas of CaSnO3, SrSnO3, and BaSnO3 samples were 13.5, 16.9, and 5.9 m2 g−1, respectively, as shown in Table 1.


image file: c6ra05614h-f2.tif
Fig. 2 SEM images of CaSnO3 (a and b), SrSnO3 (c and d), and BaSnO3 (e and f) samples.

XPS spectra of O 1s, Sn 3d, Ca 2p, Sr 3d, and Ba 3d obtained over the as-prepared MSnO3 samples were shown in Fig. 3A. The O 1s XPS spectra of CaSnO3 samples can be fitted three peaks at 529.8, 531.3, and 532.6 eV. The binding energy at 529.8 eV is ascribed to lattice oxygen, 531.3 eV for the surface hydroxyl, and 532.5 eV for adsorbed water.18,19 Similarly, the O 1s core level peak of SrSnO3 and BaSnO3 samples are also composed by lattice oxygen, surface hydroxyl, and adsorbed oxygen. However, the binding energy of different lattice oxygen decreases in the order of CaSnO3 > SrSnO3 > BaSnO3. This maybe originates from the different cation electronegativity that affects the binding energy of lattice oxygen. Fig. 3B presented the Sn 3d XPS spectra of MSnO3 samples at ca. 486.4 and 494.8 eV that corresponded to Sn 3d5/2 and Sn 3d3/2, respectively. The different value of two binding energies is 8.4 eV. As reported by Kwoka et al.,20 it can be obtained that the Sn element exist three oxidizing steps corresponding to the following binding energy values: Sn0 (485.0 eV), Sn2+ (485.9 eV) and Sn+4 (486.6 eV). Recently, our group detailedly studied the valence of Sn element by designing and preparing the Sn-grafted TiO2, Sn-grafted Ru/TiO2 and Sn, Ni-grafted TiO2 photocatalysts.21–23 Based on the previous reports and the above results, they clearly demonstrate that the valence of Sn in MSnO3 (M = Ca, Sr, and Ba) is +4. As shown in Fig. 3C, the Ca 2p3/2 and Ca 2p1/2 binding energies are 346.3 and 349.9 eV, respectively. And the difference in these two values is 3.6 eV. The Sr 3d5/2 and Sr 3d3/2 binding energies of SrSnO3 sample are 132.9 and 134.4 eV, respectively. The Ba 3d5/2 and Ba 3d3/2 binding energies of BaSnO3 sample are located at 779.4 and 794.9 eV, respectively. This is in good consistency with that reported in literature.18


image file: c6ra05614h-f3.tif
Fig. 3 XPS of spectra of MSnO3 (M = Ca, Sr, and Ba) samples. (A) O 1s, (B) Sn 3d, (C) Ca 2p, Sr 3d, Ba 3d.

3.2 Photocatalytic activity for H2 production

The photocatalytic activities for H2 production over the as-prepared MSnO3 photocatalysts were compared under UV light irradiation. As seen from Fig. 4A, it appeared that the MSnO3 samples displayed very low photocatalytic activity for H2 evolution before the photocatalytic reaction 1 h, whereas the photocatalytic activity was significantly enhanced as the reaction progressed, suggesting that there existed an induction process at the initial phase of photocatalytic reaction. This result is in good consistency with the phenomenon of H2 production from photocatalytic splitting water over ZnSn(OH)6 and Zn2SnO4 catalysts in our previous research work.24,25 Compared to the SrSnO3 and BaSnO3 samples, the CaSnO3 samples showed significantly enhanced photocatalytic activity, depending on the irradiation time. It can see from Fig. 4B that the H2 evolution rate over CaSnO3 sample reached 27.4 μmol h−1, followed by a sharp decrease of 4.2 and 2.4 μmol h−1 over SrSnO3 and BaSnO3 ones, respectively. It can be estimated by the eqn (1) that their AQE values are equal to 0.48%, 0.073%, and 0.042%, respectively. As expected, the H2 evolution rate decreases in the order CaSnO3 > SrSnO3 > BaSnO3. Obviously, the CaSnO3 sample is more active for photocatalytic reforming of ethanol/water solution to hydrogen than those of SrSnO3 and BaSnO3 samples.
image file: c6ra05614h-f4.tif
Fig. 4 (A) Time course of evolved H2 over the as-prepared MSnO3 (M = Ca, Sr, and Ba) photocatalysts under UV light irradiation. (B) The hydrogen evolution rate from the ethanol solutions over the as-prepared MSnO3 (M = Ca, Sr, and Ba) photocatalysts.

The crystal structure, electronic structure and optical excitation of MSnO3 (M = Ca, Sr, and Ba) catalysts have been studied by Zhang et al.16 Meanwhile, they clarified the relationships between the structure, the optical excitation property, and the photocatalytic H2 production. They thought that the photocatalytic activities of hydrogen-releasing rate for MSnO3 (M = Ca, Sr, and Ba) samples associated with not only the conduction band position, but the migration excitation energy. Through the analysis of the activity, two issues were presented: (1) why is the hydrogen-releasing rate small at the initial phase? And as the reaction proceeds, the activity increased suddenly. (2) The photocatalytic activities of MSnO3 (M = Ca, Sr, and Ba) samples matched with the conduction band position. However, the relationship of the activity and the conduction band position is not linear. In fact, the conduction band position and the migration excitation energy of photocatalyst are the reflection of the capabilities of the photo-electronic reduction and migration. This is one of proofs of the catalysts with high activity, but it is not equivalent to the photocatalytic hydrogen-releasing activity. The effect of the structure on the activity cannot fully explain the difference in photocatalytic activity. The photocatalytic hydrogen-releasing activities are very closely related to the reaction kinetics and the catalyst surface microstructure during the reaction. Therefore, the surface microscopic process of CaSnO3, SrSnO3, and BaSnO3 photocatalysts during photocatalytic reforming of ethanol/water solution to H2 would be detailedly characterized below.

3.3 Structural evolution

To elucidate the difference of the photocatalytic activities for H2 production over the as-prepared MSnO3 photocatalysts, XRD was first used to investigate their structural evolution. The XRD patterns of CaSnO3 sample before and after photocatalytic reaction were compared in Fig. 5A. As mentioned above, the XRD patterns consisted of a single phase of CaSnO3 with a perovskite structure before the photocatalytic H2 production. After UV illumination 10 h, some new diffraction peaks, indexed as CaSn(OH)6 (JCPDS-09-0030), appeared in the XRD, suggesting the generation of CaSn(OH)6 after the photocatalytic H2 production. The microstructural change during photocatalytic reaction was further characterized by TEM, as shown in Fig. 6. The TEM images of the cube CaSnO3 was gradually destroyed with the increase of illumination time. The results indicated that the phenomenon of light corrosion appeared in the photocatalytic reforming of ethanol/water solution to H2, resulting in the formation of CaSn(OH)6 crystals.
image file: c6ra05614h-f5.tif
Fig. 5 XRD pattern changes of MSnO3 samples before and after photocatalytic H2 production.

image file: c6ra05614h-f6.tif
Fig. 6 TEM images of CaSnO3 over the photocatalytic process. (A) 5 h, (B) 15 h, (C) 20 h.

As such, the crystal structure of SrSnO3 before and after the photocatalytic H2 production has also been changed (Fig. 5B). Before the photocatalytic reaction, the SrSnO3 sample is good crystalline with the perovskite structure. After illumination 10 h, the diffraction peak of Sn (JCPDS-04-0673) appeared in the XRD, indicating the generation of metal Sn after reaction. Unlike CaSnO3 sample, the precursor of SrSn(OH)6 was not generated.

Fig. 5C presented the XRD pattern of BaSnO3 sample before and after the photocatalytic H2 production. After reaction, in addition to the characteristic diffraction peaks of BaSnO3, the diffraction peaks of Sn and BaCO3 were also detected, suggesting the formation of Sn and BaCO3 during the photocatalytic H2 production. C2H5OH can provide the only C source in the reaction. Therefore, it was inferred that the mineralized product CO2 was generated in the photocatalytic reaction, which was evidenced by adsorption of 0.01 M Ba(OH)2 solution. Alkaline substances in water reacted with CO2 to generate CO32−, and then combined with Ba2+ to generate BaCO3.

3.4 XPS analysis

To further illustrate the structural evolution, XPS was used to investigate their chemical states. XPS spectra of O 1s, Sn 3d, Ca 2p, Sr 3d, and Ba 3d obtained over the as-prepared MSnO3 samples before and after photocatalytic reaction were characterized. As for CaSnO3 sample (Fig. 7A), the intensity of O 1s located at 529.9 eV was significantly weakened after reaction. However, the intensity of O 1s located at 531.3 eV obviously increased. The results showed that the surface lattice oxygen of CaSnO3 was reduced, and hydroxyl oxygen species were increased. Meanwhile, the chemical environment of Ca 2p and Sn 3d changed due to the crystal phase transition during the reaction. The positions of Ca 2p and Sn 3d were slightly shifted to some extent (Fig. 7B and C). In the absence of alcohol as sacrificial agent, the surface structure of CaSnO3 was similarly changed. This phenomenon verified the transition from CaSnO3 to CaSn(OH)6 under UV irradiation.
image file: c6ra05614h-f7.tif
Fig. 7 XPS spectra change of CaSnO3 over irradiation process. (A) Before reaction; (B) ethanol as sacrificial agent; (C) pure water.

Fig. 8 compared the XPS spectra of O 1s, Sn 3d, and Sr 3d obtained over SrSnO3 sample before and after reaction. Similarly, the O 1s XPS spectra of the lattice oxygen (529.9 eV) drastically reduced after photocatalytic reaction (Fig. 8A). As seen from Fig. 8B, there appeared a new Sn 3d5/2 peak located 484.8 eV on the spectrum of Sn 3d after photocatalytic reaction, which can be attributed to metal Sn.26–28 The results proved that Sn was formed during photocatalytic reaction due to the photocorrosion on the surface of SrSnO3 sample, which is consistent with the XRD results. However, at the absence of ethanol as the case of sacrificial agent, the peaks of Sn on Sn 3d spectrum was very small, indicating that the presence of ethanol can promoting the generation of metal Sn. The position of Sr 3d peak moved toward the high binding energy after the photocatalytic reaction. It indicated that there were electron-withdrawing species on the catalyst surface, increasing the binding energy of Sr 3d peak. The reason is possibly that a large number of surface defects resulted in the increase of Sn2+ under UV irradiation, changing the chemical environment of Sr, thus increasing the binding energy.


image file: c6ra05614h-f8.tif
Fig. 8 XPS spectra change of SrSnO3 over irradiation process. (A) Before reaction; (B) ethanol as sacrificial agent; (C) pure water.

Fig. 9 also displayed the XPS spectra of BaSnO3 samples before and after the photocatalytic reaction. As seen from the O 1s spectrum of BaSnO3 samples (Fig. 9A), the peak of the lattice oxygen (529.3 eV) decreased after reaction due to photocorrosion on the sample surface. Notably, there was no presence of the surface lattice oxygen in the sample after the photocatalytic H2 production. However, the adsorbed water of the surface chemistry disappeared after the water photolysis. Similarly, the surface of BaSnO3 sample formed the self-doped Sn2+. The peak of Sn 3d shifted slightly on the XPS spectra. The XPS peaks of metal Sn were not clearly seen, but observed from the XRD pattern after photocatalytic reforming of ethanol, implying that the Sn generated was distributed in the form of inhomogeneous particulates on the catalyst surface. As seen from Fig. 9C, there appeared the peak of Ba 3d5/2 located at 779.9 eV after the photocatalytic reaction, attributed to BaCO3 in Ba 3d5/2.29–31 This result was consistent with the XRD results.


image file: c6ra05614h-f9.tif
Fig. 9 XPS spectra change of BaSnO3 over irradiation process. (A) Before reaction; (B) ethanol as sacrificial agent; (C) pure water.

3.5 The pH change of the solution during photocatalytic H2 production

To better illustrate the corresponding reaction mechanism, the pH values of the solution were monitored during the process of photocatalytic reforming of ethanol/water solution to H2. Fig. 10 presented the pH change of the solution during photocatalytic H2 production over MSnO3. As seen from Fig. 10, the pH value of the solution for three catalysts was about 6.0 at the initial stages, indicating the weak acidity. The pH value did not substantially increase under the light irradiation within one hour, thus corresponding photocatalytic hydrogen-releasing rate was relatively small. With increasing the irradiation time, the pH value of the solution increased. For the CaSnO3 sample, the pH value of the solution was stable to be about 8.6 when the exposure time was more than 6 h. As for the SrSnO3 sample, the pH value of the reaction solution continued to increase during the light irradiation, and the increasing rate of the pH value became small after light irradiation of 5 h.
image file: c6ra05614h-f10.tif
Fig. 10 The pH change of photocatalytic reaction over CaSnO3, SrSnO3, and CaSnO3 samples.

The pH value was 10.3 after light irradiation of 10 h. When the photocatalyst was BaSnO3, the pH value of the solution changed with “S” type vs. the illumination time. And the pH value of the solution was 10.1 after reaction. It is obviously implied that the solution for three samples was alkaline after the photocatalytic reaction.

3.6 Reaction mechanism

After clarifying the structural evolution and the process of photocatalytic reaction in detail, the corresponding reaction mechanism was proposed to clearly explain the different photoactivity of MSnO3 samples under the same conditions. According to the TEM results mentioned above, the images of the cube CaSnO3 were gradually destroyed with the increase of illumination time. The CaSnO3 surface reacted with water to generate CaSn(OH)6, as verified by XRD and XPS above. CaSn(OH)6 is alkaline that also conforms to the pH value above. CaSn(OH)6 can be reformed by ethanol to generate H2 under UV light.24 The photocatalytic activity for H2 production over CaSn(OH)6 was far less than that over CaSnO3 one. However, the sharp increase of the activity was related to the species of CaSn(OH)6 generated during photocatalytic reforming of ethanol/water solution. This is mainly because the nanocomposite structure CaSnO3/CaSn(OH)6 was formed through the in situ coupling CaSn(OH)6 with CaSnO3, promoting the efficient separation of photogenerated charge, consequently enhancing the photocatalytic efficiency. The process of photocatalytic reforming of ethanol/water solution to H2 is as follows:
image file: c6ra05614h-t2.tif

Hydrogen was produced by photocatalytic reforming of ethanol/water solution over SrSnO3 catalyst under ultraviolet excitation. And methane and carbon monoxide were released. According to XRD and XPS, the light corrosion phenomenon occurred on the SrSnO3 surface after the photocatalytic H2 production. Sn2+ was formed on the surface of SrSnO3 sample. At the same time, Sr(OH)2 was produced in the solution. This phenomenon can be explained by the following three facts: (1) the results of XRD and XPS (Fig. 5B and 8B) showed that Sn in catalyst was transformed into metal Sn after the photocatalytic reaction. The reason is that the dismutation reaction was generated due to the intermediate species Sn2+. (2) The positions of Sn 3d peak of Sn2+ and Sn4+were very close. And it is difficult to distinguish them from XPS. However, Sr 3d peak shifted toward higher binding energy, verifying the existence of Sn2+. (3) As seen from Fig. 10, the pH value of the solution increased continually, and the pH value was 10.1 after light irradiation 10 h, suggesting that the solution was alkaline after the photocatalytic H2 production. The presence of the surface Sn2+ enhanced the separation and migration of the photogenerated charge carriers, improving the photocatalytic activity. Therefore, the activity for H2 production increased obviously after the catalytic reaction 2 h, which is consistent with the photocatalytic activity in Fig. 4. According to the results above, we think that the pathway of photocatalytic H2 evolution over SrSnO3 sample is as follows:

image file: c6ra05614h-t3.tif

Like SrSnO3, the Sn on the surface of BaSnO3 sample was transformed into Sn2+ under UV irradiation. At the same time, Ba(OH)2 was produced, which is alkaline (Fig. 10). The activity for BaSnO3 catalyst increased after light irradiation of 1 h due to the presence of Sn2+. However, as shown in Fig. 4, the H2 release rate over BaSnO3 sample slowly increased and was much smaller than those of CaSnO3 and SrSnO3. This phenomenon may be caused by the generated BaCO3 due to the photocatalytic reaction. On one hand, the BaCO3 enrichment on the catalyst surface reduced the number of the surface active sites and the surface migration rate of the photogenerated carriers. On the other hand, it hindered the molecular adsorption on the catalyst surface. The process of the photocatalytic reforming of ethanol/water solution to H2 over BaSnO3 sample is as follows:

image file: c6ra05614h-t4.tif

Therefore, it is concluded in the present work that only CaSnO3 is a potential photocatalyst for the photocatalytic reforming of ethanol/water solution to hydrogen.

4. Conclusions

The alkaline earth metal stannates MSnO3 (M = Ca, Sr, and Ba) photocatalysts prepared under the same conditions showed different structures, optical absorption properties, morphology, and activities. The great photocorrosion occurs on the surface of CaSnO3, SrSnO3, and BaSnO3 samples. And the formed surface species have great influence on their photocatalytic activities for hydrogen production from ethanol/water solution. The in situ coupling of CaSnO3 and CaSn(OH)6 formed during photocatalytic reaction can greatly promote the efficient separation and transfer of photogenerated charges, consequently enhancing the photocatalytic efficiency. As for SrSnO3, the photocorrosion can cause the formation of Sn2+ self-doped SrSnO3 nanoparticles on the surface to improve the photocatalytic activity. Unlike CaSnO3 and SrSnO3, the photocatalytic activity of BaSnO3 is gradually decreased due to the conversion of BaSnO3 to BaCO3, decreasing the number of the surface active sites and the surface migration rate of the photogenerated carriers. As expected, the H2 evolution rate over CaSnO3 sample reached the value of 27.4 μmol h−1, followed by a sharp decrease of 4.2 and 2.4 μmol h−1 over SrSnO3 and BaSnO3, respectively. It is well demonstrated in the present work that CaSnO3 is more active for photocatalytic reforming of ethanol/water solution to hydrogen than those of SrSnO3 and BaSnO3.

Acknowledgements

This work was financially supported by the NSFC (Grants No. 21403035, 21373051).

Notes and references

  1. A. Kudo and T. Kondo, J. Mater. Chem., 1997, 7, 777–780 RSC.
  2. K. Sayama and H. Arakawa, J. Photochem. Photobiol., A, 1994, 77, 243–247 CrossRef CAS.
  3. H. Kadowaki, N. Saito, H. Nishiyama, H. Kobayashi, Y. Shimodaira and Y. Inoue, J. Phys. Chem. C, 2006, 111, 439–444 CrossRef.
  4. H. Kato and A. Kudo, Chem. Phys. Lett., 1998, 295, 487–492 CrossRef CAS.
  5. A. Kudo and I. Mikami, J. Chem. Soc., Faraday Trans., 1998, 94, 2929–2932 RSC.
  6. J. Sato, H. Kobayashi, K. Ikarashi, N. Saito, H. Nishiyama and Y. Inoue, J. Phys. Chem. B, 2004, 108, 4369–4375 CrossRef CAS.
  7. C. W. Lee, D. W. Kim, I. S. Cho, S. Park, S. S. Shin, S. W. Seo and K. S. Hong, Int. J. Hydrogen Energy, 2012, 37, 10557–10563 CrossRef CAS.
  8. J. Sato, H. Kobayashi, N. Saito, H. Nishiyama and Y. Inoue, J. Photochem. Photobiol., A, 2003, 158, 139–144 CrossRef CAS.
  9. G. R. Bamwenda, T. Uesigi, Y. Abe, K. Sayama and H. Arakawa, Appl. Catal., A, 2001, 205, 117–128 CrossRef CAS.
  10. J. F. Reber and K. Meier, J. Phys. Chem., 1984, 88, 5903–5913 CrossRef CAS.
  11. N. Kakuta, N. Goto, H. Ohkita and T. Mizushima, J. Phys. Chem. B, 1999, 103, 5917–5919 CrossRef CAS.
  12. T. Ohmori, H. Mametsuka and E. Suzuki, Int. J. Hydrogen Energy, 2000, 25, 953–955 CrossRef CAS.
  13. C. Fang, B. Geng, J. Liu and F. Zhan, Chem. Commun., 2009, 2350–2352 RSC.
  14. W. F. Zhang, J. Tang and J. Ye, Chem. Phys. Lett., 2006, 418, 174–178 CrossRef CAS.
  15. D. Chen and J. Ye, Chem. Mater., 2007, 19, 4585–4591 CrossRef CAS.
  16. W. Zhang, J. Tang and J. Ye, J. Mater. Res., 2007, 22, 1859 CrossRef CAS.
  17. S. Omeiri, B. Hadjarab, A. Bouguelia and M. Trari, J. Alloys Compd., 2010, 505, 592–597 CrossRef CAS.
  18. N. Sharma, K. M. Shaju, G. V. S. Rao and B. V. R. Chowdari, J. Power Sources, 2005, 139, 250–260 CrossRef CAS.
  19. H. Wang, F. Sun, Y. Zhang, L. Li, H. Chen, Q. Wu and J. C. Yu, J. Mater. Chem., 2010, 20, 5641–5645 RSC.
  20. M. Kwoka, L. Ottaviano, M. Passacantando, S. Santucci, G. Czempik and J. Szuber, Thin Solid Films, 2005, 490, 36–42 CrossRef CAS.
  21. Q. Gu, J. Long, Y. Zhou, R. Yuan, H. Lin and X. Wang, J. Catal., 2012, 289, 88–99 CrossRef CAS.
  22. H. Huang, J. Lin, L. Fan, X. Wang, X. Fu and J. Long, J. Phys. Chem. C, 2015, 119, 10478–10492 CAS.
  23. Q. Gu, J. Long, L. Fan, L. Chen, L. Zhao, H. Lin and X. Wang, J. Catal., 2013, 303, 141–155 CrossRef CAS.
  24. X. Fu, D. Y. C. Leung, X. Wang, W. Xue and X. Fu, Int. J. Hydrogen Energy, 2011, 36, 1524–1530 CrossRef CAS.
  25. X. Fu, X. Wang, J. Long, Z. Ding, T. Yan, G. Zhang, Z. Zhang, H. Lin and X. Fu, J. Solid State Chem., 2009, 182, 517–524 CrossRef CAS.
  26. Y.-J. Hsu and S.-Y. Lu, J. Phys. Chem. B, 2005, 109, 4398–4403 CrossRef CAS PubMed.
  27. B. Tang, W. Dai, G. Wu, N. Guan, L. Li and M. Hunger, ACS Catal., 2014, 4, 2801–2810 CrossRef CAS.
  28. C. Zegadi, K. Abdelkebir, D. Chaumont, M. Adnane and S. Hamzaoui, Adv. Mater. Phys. Chem., 2014, 4, 93–104 CrossRef CAS.
  29. M. D. C. B. López, G. Fourlaris, B. Rand and F. L. Riley, J. Am. Ceram. Soc., 1999, 82, 1777–1786 CrossRef.
  30. M. Wegmann, L. Watson and A. Hendry, J. Am. Ceram. Soc., 2004, 87, 371–377 CrossRef CAS.
  31. Y. Li, P.-C. Su, L. M. Wong and S. Wang, J. Power Sources, 2014, 268, 804–809 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2016