Next Article in Journal
Effect of Domain Size, Boundary, and Loading Conditions on Mechanical Properties of Amorphous Silica: A Reactive Molecular Dynamics Study
Next Article in Special Issue
Sonochemical Synthesis of Copper-doped BiVO4/g-C3N4 Nanocomposite Materials for Photocatalytic Degradation of Bisphenol A under Simulated Sunlight Irradiation
Previous Article in Journal
Morpho-Structural, Thermal and Mechanical Properties of PLA/PHB/Cellulose Biodegradable Nanocomposites Obtained by Compression Molding, Extrusion, and 3D Printing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Construction of a MgSn(OH)6 Perovskite/SnO2 Type-II Heterojunction: A Highly Efficient Photocatalyst towards Photodegradation of Tetracycline

1
Department of Biological and Chemical Engineering, Cooperative Innovation Center of Lipid Resources and Children’s Daily Chemicals, Chongqing University of Education, Chongqing 400067, China
2
College of Chemistry and Chemical Engineering, Chongqing University of Technology, 69 Hongguang Rd., Lijiatuo, Banan District, Chongqing 400054, China
3
Analytical and Testing Center of Chongqing University, Chongqing 401331, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(1), 53; https://doi.org/10.3390/nano10010053
Submission received: 14 November 2019 / Revised: 20 December 2019 / Accepted: 21 December 2019 / Published: 24 December 2019
(This article belongs to the Special Issue Fabrication of Heterostructure Nanomaterials for Catalysis)

Abstract

:
Using solar energy to remove antibiotics from aqueous environments via photocatalysis is highly desirable. In this work, a novel type-II heterojunction photocatalyst, MgSn(OH)6/SnO2, was successfully prepared via a facile one-pot in situ hydrothermal method at 220 °C for 24 h. The obtained heterojunctions were characterized via powder X-ray diffraction, Fourier-transform infrared spectroscopy, transmission electron microscopy, and ultraviolet-visible diffuse reflectance spectroscopy. The photocatalytic performance was evaluated for photodegradation of tetracycline solution under ultraviolet irradiation. The initial concentration of tetracycline solution was set to be 20 mg/L. The prepared heterojunctions exhibited superior photocatalytic activity compared with the parent MgSn(OH)6 and SnO2 compounds. Among them, the obtained MgSn(OH)6/SnO2 heterojunction with MgCl2·6H2O:SnCl4·5H2O = 4:5.2 (mmol) displayed the highest photocatalytic performance and the photodegradation efficiency conversion of 91% could be reached after 60 min under ultraviolet irradiation. The prepared heterojunction maintained its performance after four successive cycles of use. Active species trapping experiments demonstrated that holes were the dominant active species. Hydroxyl radicals and superoxide ions had minor effects on the photocatalytic oxidation of tetracycline. Photoelectrochemical measurements were used to investigate the photocatalytic mechanism. The enhancement of photocatalytic activity could be assigned to the formation of a type-II junction photocatalytic system, which was beneficial for efficient transfer and separation of photogenerated electrons and holes. This research provides an in situ growth strategy for the design of highly efficient photocatalysts for environmental restoration.

Graphical Abstract

1. Introduction

Pharmaceutical residues in wastewater are a severe threat to the ecological environment. Misuse and overuse of antibiotics can lead to high antibiotic resistance and multi-resistant strains of microorganisms [1,2,3,4]. As a broad-spectrum antibiotic, tetracycline (TC) is widely used for disease treatment in humans, and it is frequently detected in various bodies of water owing to its high solubility. Removal of TC from wastewater is important. Semiconductor-based photocatalysis, an advanced oxidation method, is the most promising technique for harvesting solar energy to alleviate environmental problems [5,6]. The discovery and design of photocatalysts that are stable, efficient, inexpensive, and ecofriendly is of importance [7,8]. Among the numerous photocatalytic materials, perovskite-based materials have excellent stability, crystalline structures with high symmetry, and diverse chemical and physical properties [9,10].
Recently, a perovskite-type hydroxide, MSn(OH)6 (M = Mg, Mn, Fe, Co, Ni, Cu, Zn, Sr, or Ba), was prepared via the hydrothermal method and showed excellent potential as a photocatalyst for environmental remediation [11,12]. In the crystalline structure of this compound, the metal ions are octahedrally coordinated with OH to form M(OH)6 and Sn(OH)6 polyhedra that connect with each other by sharing O atoms to build the whole crystalline structure. Because they have different Shannon radii, these hydroxides exhibit multiple crystalline symmetries. Compounds containing Mg, Ca, Mn, Fe, Co, Ni, Cu, Zn, or Cd adopt the rock salt ReO3 structure with cubic or tetragonal symmetry [13,14,15,16]. In contrast, those containing Sr or Ba exhibit a distorted perovskite structure in the monoclinic family with the space group P21/n [14]. The wide variety of compositions and constituent elements in these hydroxide perovskites offers plenty of room to study their photocatalytic performance. Considerable effort has been devoted to study their applications in environmental and energy catalysis research. Nanocubes of CaSn(OH)6 and CuSn(OH)6 with a d10-d10 electronic configuration exhibit high degradation rates for methylene blue [17,18,19]. The CoSn(OH)6 catalyst has high activity and selectivity for the photoreduction of CO2 to CO [20]. MgSn(OH)6 and ZnSn(OH)6 could be applied to degradation of gaseous C6H6 and reforming of ethanol to H2 and CH4 [11,21]. These studies have shown that hydroxide perovskites have excellent potential for photocatalysis. However, few studies have investigated the ability of this family of compounds to remove antibiotics in photocatalysis.
Generally, single photocatalysts have some drawbacks, such as the rapid recombination of photogenerated electrons and holes and relatively small reactive sites, which will limit their practical application [22]. Construction of heterostructures is an effective method to overcome these shortcomings and numerous reports have demonstrated that the interface formed between two different semiconductors could accelerate carrier transfer, and the recombination rate of the spatially separated photogenerated carriers can be reduced, resulting in greatly improved photocatalytic performance [23,24,25,26]. A high-quality heterojunction with a good lattice match is highly desired, and could be fabricated by co-sharing of one or several of the same atoms between the formed coherent or semi-coherent interface.
In this work, SnO2 was investigated for coupling with MgSn(OH)6 because of its chemical stability, low exciton energy, and outstanding conductivity. The simulated electron and hole mobilities along the x-direction were reported to be ~2966 and 66 cm2/Vs, respectively. Because of the large difference in transport behavior, SnO2 has been widely used as an efficient photocatalyst [27]. Furthermore, at the atomic level, an intimately contacted interface by co-sharing of the Sn atoms would be expected and can be fabricated via a one-pot in situ hydrothermal treatment between MgSn(OH)6 and SnO2. The observed SnO2 nanoparticles were loaded tightly on the surface of the MgSn(OH)6 nanocubes. As a result, photoinduced carriers in the heterojunction were effectively separated and the photocatalytic activity was remarkably improved for degradation of a TC solution under ultraviolet (UV) irradiation. Finally, a possible photocatalytic mechanism for the type-II heterojunction is proposed and discussed.

2. Materials and Methods

2.1. Synthesis of SnO2, MgSn(OH)6, and MgSn(OH)6/SnO2 Heterojunctions

The following materials were applied for this experiment: MgCl2·6H2O (Aladdin, 98%, Shanghai, China), SnCl4·5H2O (Aladdin, 99%, Shanghai, China), tetracycline (Adamas-beta, 97%+, Shanghai, China), NaOH (Aladdin, 96%, Shanghai, China), ethyl alcohol (Aladdin, 99.7%, Shanghai, China), 2-propanol (IPA) (Aladdin, 99.7%, Shanghai, China), disodium ethylenediaminetetraacetic acid (EDTA) (Aladdin, 99.5%, Shanghai, China), and 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) (Aladdin, 98%, Shanghai, China). All of the above chemicals were directly utilized without any further purification.
To obtain pure phase SnO2 (SO), 4 mmol SnCl4·5H2O was dissolved in 15 mL of water. Then, 4 mL NaOH (0.5 g/mL) was added to adjust pH to ~7 and the mixture was further stirred for 5 min. Next, the mixture was transferred into a 25-mL Teflon-lined steel autoclave and hydrothermally heated at 220 °C for 24 h.
For preparation of pure MgSn(OH)6 (MSOH), 4 mmol MgCl2·6H2O, 4 mmol SnCl4·5H2O, and 12 mL NaOH (0.5 g/mL) were added to 15 mL of deionized water to adjust pH to ~11 and magnetically stirred for further 10 min. The obtained slurry was hydrothermally treated at 220 °C for 24 h in a Teflon-lined steel autoclave.
Heterojunctions of MgSn(OH)6/SnO2 with different mass fractions were prepared by an in situ hydrothermal formation strategy. The mass fraction of SnO2/MgSn(OH)6 was adjusted to 10%, 20%, 30%, or 40% by controlling the amount of SnCl4·5H2O added, and the obtained products were labeled as MSOH-SO-1 (SnCl4·5H2O/MgCl2·6H2O = 4.4/4 mmol), MSOH-SO-2 (SnCl4·5H2O/MgCl2·6H2O = 4.8/4 mmol), MSOH-SO-3 (SnCl4·5H2O/MgCl2·6H2O = 5.2/4 mmol), and MSOH-SO-4 (SnCl4·5H2O/MgCl2·6H2O = 5.6/4 mmol), respectively. Taking MSOH-SO-3 as an example, 4 mmol MgCl2·6H2O and 5.2 mmol SnCl4·5H2O were dispersed in 15 mL of deionized water, followed by addition of were 4 mL NaOH (0.5 g/mL) to adjust pH to ~8. The mixture was further magnetically stirred at room temperature for 30 min and then put into a 25-mL Teflon-lined steel autoclave and kept in an oven at 220 °C for 24 h.
All the products were washed with distilled water several times and then dried at 70 °C for 10 h.

2.2. Characterization

Powder X-ray diffraction (PXRD) was performed on a PANalytical X’pert powder diffractometer equipped with a PIXcel detector and using Cu Kα radiation (40 kV and 40 mA). A scan step width of 0.01° and a scan rate of 0.1° s−1 were applied to record patterns in the 2θ range of 6°–90°. Transmission electron microscopy (TEM), high-angle annular dark field imaging (HAADF), and energy-dispersive X-ray spectrometry of the MgSn(OH)6/SnO2 heterojunctions were performed using a Talos F200S G2 microscope to investigate the microstructures of the samples. Ultraviolet-Visible diffuse reflectance spectroscopy (UV–Vis DRS) data were collected at room temperature using a powder sample with BaSO4 as a standard on a Shimadzu UV-3150 spectrophotometer over the spectral range of 200–800 nm. Fourier transform infrared (FT-IR) spectra were collected using a Nicolet 360 spectrometer with a 2 cm–1 resolution in the range of 500–4000 cm−1. The photocatalysts were fixed within a pressed KBr pellet. For instance, 1 mg heterojunction MSOH-SO-3 within 100 mg KBr were pressed at 15 MPa for 15 min, forming 13 mm pellets. Total dissolved organic carbon (TOC) was determined via a TOC analyzer (SHIMADZU, TOC-L CPB). The BET specific surface areas were investigated by means of N2 adsorption-desorption at 77 K using a Quantachrome QuadraWin and the specific surface areas were determined according to the BET method in the relative pressure range p/p0 = 0.069~0.249. Electrochemical measurements were conducted on a CHI 660E workstation. A Pt plate, calomel electrode, and MSOH-SO-3 coated on indium tin oxide (ITO) were used as the counter electrode, reference electrode, and working electrode, respectively, in a three-electrode cell. For a typical preparation of working electrode, 10 mg of catalyst samples were dispersed in 500 μL of water/ethanol (240/250) mixed solvent containing 10 μL of 5 wt % Nafion and sonicated for 30 min. Then, a certain amount of the catalyst ink was loaded onto an ITO electrode. Electrochemical impedance spectroscopy (EIS) was carried out using an alternating voltage with an amplitude of 5 mV over the frequency range of 105 Hz to 0.1 Hz and an open circuit voltage in 0.5 M Na2SO4. For analysis of the transient photocurrent responses, a 175 W high-pressure Hg lamp and Na2SO4 were used as the light source and electrolyte, respectively. Mott–Schottky curves were measured in a Na2SO4 solution with an amplitude perturbation of 5 mV and frequency of 1000 Hz.

2.3. Photocatalytic Activity Measurements

The photocatalytic activities of the products were evaluated for degradation of a TC solution. The light source was an external 175 W high-pressure mercury lamp with a primary wavelength of 365 nm. The lamp was positioned 9 cm above the reaction vessel and the light intensity on the surface of the suspension was approximately 11 mW/cm2. The reactant solution was maintained at room temperature using a stream of cool water during the photocatalytic reaction. Before irradiation, the photocatalyst powder (30 mg) and TC solution (20 mg L−1, 100 mL) were fully stirred in the dark for 1 h to establish the adsorption–desorption equilibrium and the pH was measured to be ~5.5. Then, the solution was exposed to the light. Aliquots (5 mL) of the suspension was taken at set intervals (20 min) and separated by centrifugation. Then, the concentration of the TC solution was determined by UV–Vis spectrometry at 355 nm and the process of photocatalytic performance was carried out within 60 min. Considering the small loss of the catalysts in the recycling process, several batches of repeated experiments for each cycle were performed. Then, the catalysts were collected and mixed to maintain the weight of 30 mg for each test. The degradation rate could be calculated by the formula
Photodegadation (%) = 1− C/C0
where C0 was the absorbance of the initial solution and C was the absorbance of solution at a given time after the photocatalytic reaction. Trapping experiments of the active species were carried out using 30 mg of MSOH-SO-3 and 100 mL of TC solution (20 mg/L). The reactive intermediate participating in the degradation process was identified by using different sacrificial agents. Such as, 10 mL of 2-propanol (IPA), 0.1 mmol disodium ethylenediaminetetraacetic acid (EDTA), and 0.1 mmol 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) were added sequentially to trap ·OH, h+, and ·O2 radicals, respectively. TOC analysis was carried out at every 20 min to check the degree of mineralization over heterojunction MSOH-SO-3 under UV light irradiation.

3. Results and Discussion

Photocatalysts synthesized from SnO2, MgSn(OH)6, and the MSOH-SO heterojunction were characterized by PXRD (Figure 1). Peaks at 19.7°, 22.8°, 32.5°, and 52.6° matched well with the crystalline planes (111), (200), (220), and (420), respectively, that belong to the cubic phase MgSn(OH)6 (JCPDS 074-0366). Major diffraction peaks at 26.7° (110 plane), 33.8° (101 plane), and 51.7° (211 plane) were indexed to the SnO2 phase (JCPDS 021-1250). Notably, the peak intensity of SnO2 gradually strengthened with increasing SnCl4·5H2O concentration, indicating that the amount of SnO2 loaded on MgSn(OH)6 increased. The structures of these photocatalysts were confirmed by FT-IR spectroscopy. Specifically, for SnO2, a broad peak located at ~618 cm−1 was attributed to the antisymmetric vibrational mode of Sn-O-Sn [28]. For MgSn(OH)6, absorption peaks in the range 3000–3500 cm−1 were ascribed to O–H stretching vibration modes and lattice modes at 1175 cm−1 corresponded to the O–H bending vibrations. Peaks in the range of 500–1000 cm−1 were attributed to the Mg–O and Sn–O bond stretching vibration modes [29]. All of the main characteristic peaks of MgSn(OH)6 and SnO2 were observed in MSOH-SO. On the basis of the PXRD and FT-IR observations, we concluded that both SnO2 and MgSn(OH)6 existed in the heterojunctions.
The typical interface structure of the synthesized heterojunction MSOH-OH-3 was further characterized by HAADF–TEM observation. The MgSn(OH)6 nanocubes were combined with randomly oriented nanoparticles of SnO2 (Figure 2a). Furthermore, well-ordered lattice fringes with a lattice spacing 0.38 nm could be ascribed to the (200) plane of MgSn(OH)6. The interplanar distances of 0.32 nm and 0.25 nm could be assigned to the (110) and (101) planes of the SnO2 nanoparticles, respectively. Elemental mapping via HAADF–energy-dispersive X-ray spectroscopy (Figure 2b) confirmed the uniform distributions of Mg, Sn, and O elements in the MSOH-SO-3 heterojunction. The contents of Sn and O elements were much higher than that of Mg, which showed that a heterojunction formed between MgSn(OH)6 and SnO2 (Figure S1). All of the above observations demonstrated that there was close contact at the interface of the synthesized heterojunction, which might be favorable for promoting the separation of photogenerated carriers at the interface during the photocatalytic reaction. Besides, we also measured the BET specific surface areas of the as-synthesized samples, heterojunction MSOH-SO-3, SO, and MSOH. As presented in Figure S2, the results showed that the calculated specific surface areas of sample MSOH-SO-3 and SO were 114.2 m2/g and 81.7 m2/g, respectively. However, the obtained specific surface area of MSOH was too small to ignore.
To investigate the optical absorption properties and forbidden band gap of the synthesized photocatalysts, the UV–Vis DRS spectra of MgSn(OH)6, MSOH-SO-3, and SnO2 were recorded (Figure 3). All the photocatalysts showed an obvious absorption edge in the UV region. The absorption edges for MgSn(OH)6 and SnO2 were at approximately 303 nm and 343 nm, respectively. The corresponding band gaps estimated from the curve of ( α h ν ) 2 versus the photon energy (hν) were 4.09 eV (Figure 3b) and 3.61 eV (Figure 3d), respectively. The MSOH-SO-3 heterojunction exhibited a blue shift compared with SnO2 (Figure 3c), which would facilitate the separation of photoinduced carriers and improve the photocatalytic performance.
The photocatalytic activities of the prepared catalysts were evaluated for photodegradation of a TC solution under UV irradiation (Figure 4). Blank experiments showed that UV light had little effect on photodegradation of the TC solution (Figure 4a). The degradation rates of pure SnO2 and MgSn(OH)6 were 44% and 34%, respectively. However, when SnO2 was combined with MgSn(OH)6 to form a heterojunction, the photocatalytic activity was enhanced. Among the heterojunctions, MSOH-SO-3 exhibited the best photocatalytic ability and the photodegradation rate reached nearly 91% in 60 min, which was much larger than that of state-of-art UV light responsive ZnO [30]. The corresponding changes in the characteristic absorption of the TC solution are shown in Figure S3. The behavior of the photocatalytic reaction was described well by a pseudo-first order model [31,32,33]
ln C C 0 = k t ,
where c 0 and c are the TC concentrations in solution at time 0 and t, respectively; and k is the fitted kinetic rate constant. A linear relationship was observed between ln C 0 C and t (Figure 4b), which revealed that the synthesized catalysts for photodegradation of TC obeyed the model perfectly. The fitted rate constants of MgSn(OH)6 and SnO2 were 0.006 min−1 and 0.009 min−1, respectively (Figure 4c). The k values of the heterojunctions were all higher than those of the parent compounds, and the corresponding rate constants of MSOH-SO-1, MSOH-SO-2, MSOH-3, and MSOH-SO-4 were 0.018 min−1, 0.024 min−1, 0.030 min−1, and 0.028 min−1, respectively. Among the heterojunctions, MSOH-SO-3 had the highest k, which was approximately five times that of MgSn(OH)6 and 3.3 times that of SnO2. The photocatalytic activity was further evidence of successful preparation of the heterojunctions and an efficient synergistic effect at the interface. The total organic carbon (TOC) performance was also performed to check mineralization ability over heterojunction MSOH-SO-3. As indicated in Figure S4, after 60 min irradiation, the TOC removal efficiency of mineralizing TC molecule could reach 35%.
To evaluate its practicality as a photocatalyst, the stability of MSOH-SO-3 was assessed in four successive cycles of reuse in photocatalytic experiments. The photocatalytic performance showed no obvious loss after four cycles (Figure 4d) and the observed PXRD patterns remained the same (Figure S5). Statistically, the standard deviations of kinetic rate constants of the four cycling performance is 0.00299. These results show that the MSOH-SO-3 heterojunction is a stable and reusable photocatalyst for the degradation of TC under UV irradiation.
The photocurrent response and EIS showed a close association between the migration, transfer, and recombination of photoinduced electron–hole pairs in the photocatalysts. The photocurrent intensities of MgSn(OH)6, SnO2, and MSOH-SO-3 were measured for four cycles of intermittent (on–off) UV irradiation (Figure 5a). The photocurrent intensities of MgSn(OH)6 were lower than 0.4 µA/cm2, whereas the photocurrent intensities of SnO2 were lower than 0.7 µA/cm2. After the coupling of SnO2 with MgSn(OH)6, a clear enhancement of the photocurrent response was observed, which suggested that the electric field at the interface of the heterojunction could accelerate the movement of carriers and greatly promote separation of the photoinduced electrons and holes. Additionally, EIS was conducted to understand the carrier transfer resistance using the arc radius [34]. Nyquist plots with smaller arc radii indicate a low impedance and a high carrier transfer efficiency. MSOH-SO-3 had the smallest arc radius compared with the parent compounds MgSn(OH)6 and SnO2 (Figure 5b). The low electron transfer resistance of MSOH-SO-3 confirmed successful fabrication of the MgSn(OH)6/SnO2 heterojunction, which could effectively improve the efficiency of electron–hole pair separation and increase the photocurrent intensity and photodegradation activity.
Trapping experiments of the reactive species during the photocatalytic process were performed to investigate the photocatalytic mechanism. For this, MSOH-SO-3. EDTA, TEMPO, and IPA were applied separately as scavengers to determine the concentrations of h+, ·O2, and ·OH, respectively. Introduction of TEMPO and IPA in the photocatalytic process had little effect on the degradation of the TC solution (Figure 6), indicating that ·O2 and ·OH play minor roles in the photodegradation process. In contrast, following the addition of EDTA, a dramatic inhibition of the photocatalytic efficiency was observed and the degradation rate decreased rapidly to 27%. Thus, it could be inferred that h+ was an active participant in photodegradation of TC with the MSOH-SO-3 heterojunction.
The band edge positions of MgSn(OH)6 and SnO2 were determined using the Mott–Schottky technique. Generally, the quasi Fermi level (flat band potential V f b ) can be obtained from the x-intercept of a Mott–Schottky plot ( 1 C 2 = 0 ) as a function of the applied potential according to the following formula [35,36]
1 C 2 = ( 2 ε r ε 0 N d e ) × ( V V f b κ B T e ) ,
where C is the space charge capacitance; ε r and ε 0 are the dielectric constant of the semiconductor and permittivity in vacuum, respectively; e is the electronic charge; N d is the carrier density; and V, V f b ,   κ B , and T are the applied voltage, flat-band potential, Boltzmann constant, and temperature, respectively. The flat band potential V f b was set to 0.1 V below the conduction band (CB) minimum or above the valence band (VB) maximum for n-type and p-type semiconductors, respectively [37]. The Mott–Schottky plots of MgSn(OH)6 and SnO2 (Figure 7a,b) had positive slopes, which indicated that they were n-type semiconductors. The estimated V f b of MgSn(OH)6 and SnO2 were −1.155 V and −0.655 V versus the saturated calomel electrode (SCE), respectively, and could be calibrated to normal hydrogen electrode potentials as [38,39]
V R H E = V S C E + 0.059 p H + V S C E 0 ,
where V R H E is the calibrated potential versus the reversible hydrogen electrode(RHE), V S C E 0 is 0.245 V, and V S C E is obtained from experimental values. Thus, the V f b of MgSn(OH)6 and SnO2 were −0.505 and −0.005 V, respectively, versus RHE after calibration. Therefore, the calibrated CB minima of MgSn(OH)6 and SnO2 were −0.605 and −0.105 V, respectively. According to the optical band gaps from the UV–Vis DRS curves, the valence band maxima of MgSn(OH)6 and SnO2 were 3.485 and 3.505 V, respectively.
In light of the above experimental results and analysis, a possible photocatalytic mechanism for TC degradation using the MSOH-SO-3 heterojunction is proposed in Figure 8. According to the calculated band position, MgSn(OH)6 and SnO2 exhibited type-II band alignment, and an energy bias was naturally generated at the contact interfaces between the two compounds. This facilitated the separation and transfer of photoinduced carriers. Under UV irradiation, both MgSn(OH)6 and SnO2 could be excited to produce electrons and holes. The excited electrons on the CB of MgSn(OH)6 and LOMO level of TC molecule tended to flow down to the CB of SnO2 across the interface. Meanwhile, the remaining holes on the VB of SnO2 moved to the VB of MgSn(OH)6 and HOMO level of TC molecule. This transfer mechanism greatly prolonged the relaxation time of the photoinduced carriers, which agreed with the photoelectrochemical measurement results. The accumulated electrons in the CB position of SnO2 were more positive than O2/·O2, thus ·O2 would theoretically not participate in the photocatalytic reaction. By contrast, the accumulated holes h+ in MgSn(OH)6 could capture OH to generate ·OH thermodynamically. However, there was no large decrease in the degradation of the TC solution in the presence of IPA according to our reactive species trapping experiments. These results indicated that the produced · OH would not be favorable for the oxidation of TC, as reported earlier [40]. Thus, the h+ present in the VB of MgSn(OH)6 could directly oxidize the TC solution to the corresponding degradation products. Therefore, formation of the type-II heterojunction that accelerated the carrier separation could be considered as the primary factor for the enhanced photocatalytic activity towards the TC degradation.

4. Conclusions

In summary, a series of novel MSOH-SO heterojunction photocatalysts containing n-type MgSn(OH)6 and n-type SnO2 were successfully prepared via a facile one-pot in situ method. The heterojunction MSOH-SO-3 exhibited outstanding photodegradation activity for TC solution and the maximum photodegradation rate was 91% after UV irradiation within 60 min. This rate was approximately 5 times that of MgSn(OH)6 and 3.3 times that of SnO2. The enhancement of the photocatalytic performance could be attributed to the successful preparation of a type-II heterojunction between MgSn(OH)6 and SnO2, which greatly enhanced the efficient separation of photoinduced carriers based on the structural analysis, photoelectrochemical measurement and reactive species detecting experiments. The photoinduced h+ radicals played a dominant role in the photocatalytic degradation of a TC solution over the MSOH-SO-3 heterojunction photocatalyst. This research might provide a facile way for construction highly efficient heterojunction photocatalysts.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/1/53/s1, Figure S1: elemental mapping of heterojunction MSOH-SO-3; Figure S2: The specific surface area determination performed by BET measurement with N2 adsorption-desorption. (a) Heterojunction MSOH-SO-3 (b) SnO2 (SO); Figure S3: UV-visible absorption of TC as a function of irradiation time over the heterojunction MSOH-SO-3; Figure S4: The TOC degradation of TC solution under different UV-light irradiation. (TOC0: initial TOC value; TOCt: TOC value at photodegradation time t); Figure S5: XRD patterns of sample MSOH-SO-3 of cycling runs on the photocatalytic reaction of degrading TC solution.

Author Contributions

All authors made important contributions on this work. Specifically, Y.L. and D.Y. proposed this topic and design of the project; B.Z. completed the characterization part. X.T., Q.Y., and Y.D. completed the experimental part; Y.W. analyzed the results; D.Y., B.Z., and Y.L. composed the manuscript. All authors participated in the discussions of the results. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by the National Natural Science Foundation of China (no. 11904039), the Basic and Frontier Research Project of Chongqing Science and Technology Commission (nos. cstc2018jcyjAX0827, cstc2019jcyj-msxmX0187), the project of Scientific and Technological Research Program of Chongqing Municipal Education Commission (nos. KJQN201801135). This project was funded by Children’s Research Institute of National Center for Schooling Development Program and Chongqing University of Education (no. CSDP19FS01109) and the project of Science and Technology Collaborative Innovation Platform Construction of Chongqing University of Education (no. 2017XJPT01), and the Cultivation for National Science Foundation of Chongqing University of Education (no. 18GZKP01).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bakkeren, E.; Huisman, J.S.; Fattinger, S.A.; Hausmann, A.; Furter, M.; Egli, A.; Slack, E.; Sellin, M.E.; Bonhoeffer, S.; Regoes, R.R.; et al. Salmonella Persisters Promote the Spread of Antibiotic Resistance Plasmids in the Gut. Nature 2019, 573, 276–280. [Google Scholar] [CrossRef] [PubMed]
  2. Yelin, I.; Snitser, O.; Novich, G.; Katz, R.; Tal, O.; Parizade, M.; Chodick, G.; Koren, G.; Shalev, V.; Kishony, R. Personal clinical history predicts antibiotic resistance of urinary tract infections. Nat. Med. 2019, 25, 1143–1152. [Google Scholar] [CrossRef] [PubMed]
  3. Yoshida, M.; Reyes, S.G.; Tsuda, S.; Horinouchi, T.; Furusawa, C.; Cronin, L. Time Programmable Drug Dosing Allows the Manipulation, Suppression and Reversal of Antibiotic Drug Reversal of Antibiotic Drug Resistance in Vitro. Nat. Commun. 2017, 8, 15589–15599. [Google Scholar] [CrossRef] [PubMed]
  4. Amado, S.H.; Coque, T.M.; Baquero, F.; Martinez, J.L. Defining and Combating Antibiotic Resistance from One Health and Global Health Perspectives. Nat. Microbiol. 2019, 4, 1432–1442. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, P.; Blaney, L.; Cagnetta, G.; Huang, J.; Wang, B.; Wang, Y.J.; Deng, S.B.; Yu, G. Degradation of Ofloxacin by Perylene Diimide Supramolecular Nanofiber Sunlight Driven Photocatalyst. Environ. Sci. Technol. 2019, 53, 1564–1575. [Google Scholar] [CrossRef]
  6. Yan, J.F.; Peng, J.L.; Lai, L.; Ji, F.Z.; Zhang, Y.H.; Lai, B.; Chen, Q.X.; Yao, G.; Chen, X.; Song, L.P. Activation CuFe2O4 by Hydroxylamine for Oxidation of Antibiotic Sulfamethoxazole. Environ. Sci. Technol. 2018, 52, 14302–14310. [Google Scholar] [CrossRef]
  7. Zhang, X.; Zhao, X.D.; Wu, D.H.; Jing, Y.; Zhou, Z. MnPSe3 Monolayer: A Promising 2D Visible Light Photohydrolytic Catalyst with High Carrier Mobility. Adv. Sci. 2016, 3, 1600062–1600066. [Google Scholar] [CrossRef] [Green Version]
  8. Zhang, X.; Zhang, Z.H.; Wu, D.H.; Zhang, X.; Zhao, X.D.; Zhou, Z. Computational Screening of 2D Materials and Rational Design of Heterojunctions for Water Splitting Photocatalysis. Small Methods 2018, 2, 1700359–1700367. [Google Scholar] [CrossRef]
  9. Hailili, R.; Wang, Z.Q.; Gong, X.Q.; Wang, C.Y. Octahedral-shaped Perovskite CaCu3Ti4O12 with Dual Defects and Coexposed {(001), (111)} Facets for Visible Light Photocatalysis. Appl. Catal. B 2019, 254, 86–97. [Google Scholar] [CrossRef]
  10. Kong, J.J.; Yang, T.; Rui, Z.B.; Ji, H.B. Perovskite Based Photocatalysts for Organic Contaminants Removal: Current Status and Future Perspectives. Catal. Today 2019, 327, 47–63. [Google Scholar] [CrossRef]
  11. Huang, D.W.; Fu, X.L.; Long, J.L.; Jiang, X.L.; Chang, L.; Meng, S.G.; Chen, S.F. Hydrothermal Synthesis of MSn(OH)6 (M = Co, Cu, Fe, Mg, Mn and Zn) and Their Photocatalytic Activity for the Destruction of Gaseous Benzene. Chem. Eng. J. 2015, 269, 168–179. [Google Scholar] [CrossRef]
  12. Wang, M.; Cao, X.L.; Huang, Y.F.; Guo, C.S.; Huang, L.J.; Yin, S.; Sato, T. Solvent Free Mechanochemical Synthesis of Well Dispersed Single Crystalline Zinc Hydroxystannate and Their Photocatalytic Properties. CrystEngComm 2012, 14, 2950–2953. [Google Scholar] [CrossRef]
  13. Luo, Y.P.; Chen, J.; Liu, J.W.; Shao, Y.; Li, X.F.; Li, D.Z. Hydroxide SrSn(OH)6: A New Photocatalyst for Degradation of Benzene and Rhodamine B. Appl. Catal. B 2016, 182, 533–540. [Google Scholar] [CrossRef]
  14. Mizoguchi, H.; Bhuvanesh, N.S.P.; Kim, Y.I.; Ohara, S.; Woodward, P.M. Hydrothermal Crystal Growth and Structure Determination of Double Hydroxides LiSb(OH)6, BaSn(OH)6 and SrSn(OH)6. Inorg. Chem. 2014, 53, 10570–10577. [Google Scholar] [CrossRef]
  15. Kramer, J.W.; Kelly, B.; Manivannan, V. Synthesis of MSn(OH)6 (where M = Mg, Ca, Zn, Mn or Cu) materials at room temperature. Cent. Eur. J. Chem. 2010, 8, 65–69. [Google Scholar] [CrossRef]
  16. Jena, H.; Kutty, K.V.G.; Kutty, T.R.N. Ionic Transport and Structural Investigations on MSn(OH)6 (M = Ba, Ca, Mg, Co, Zn, Fe and Mn) Hydroxide Perovskites Synthesized by Wet Sonochemical Methods. Mater. Chem. Phys. 2004, 88, 167–179. [Google Scholar] [CrossRef]
  17. Baeissa, E.S. Novel Pd/CaSn(OH)6 Nanocomposite Prepared by Modified Sonochemical Method for Photocatalytic Degradation of Methylene Blue Dye. J. Alloys Compd. 2014, 590, 303–308. [Google Scholar] [CrossRef]
  18. Liu, T.Y.; Ma, X.L.; Yang, L.F.; Li, H.; Li, H.H.; Lee, S.W.; Wang, Y.H. Highly Enhanced Photocatalytic Activity of CaSn(OH)6 through Tuning CaSn(OH)6/SnO2 Heterostructural Interaction and Optimizing Fe3+ Doping Concentration. Appl. Catal. B 2017, 217, 256–264. [Google Scholar] [CrossRef]
  19. Dong, S.Y.; Cui, L.F.; Zhao, Y.L.; Wu, Y.W.; Xia, L.J.; Su, X.F.; Zhang, C.Y.; Wang, D.; Guo, W.; Sun, J.H. Crystal Structure and Photocatalytic Properties of Perovskite MSn(OH)6 (M = Cu and Zn) Composites with d10-d10 configuration. Appl. Surf. Sci. 2019, 463, 659–667. [Google Scholar] [CrossRef]
  20. Gao, Y.; Ye, L.; Cao, S.Y.; Chen, H.; Yao, Y.N.; Jiang, J.; Sun, L.C. Perovskite Hydroxide CoSn(OH)6 Nanocubes for Efficient Photoreduction of CO2 to CO. ACS Sustain. Chem. Eng. 2018, 6, 781–786. [Google Scholar] [CrossRef]
  21. Fu, X.L.; Leung, D.Y.C.; Wang, X.X.; Xue, W.W.; Fu, X.Z. Photocatalytic Reforming of Ethanol to H2 and CH4 over ZnSn(OH)6 nanocubes. J. Hydrog. Energy 2011, 36, 1524–1530. [Google Scholar] [CrossRef]
  22. Li, M.Y.; Tang, Y.B.; Shi, W.L.; Chen, F.Y.; Shi, Y.; Gu, H.C. Design of Visible Light Response Core-Shell Fe2O3/CuBi2O4 Heterojunctions with Enhanced Photocatalytic Activity Towards the Degradation of Tetracycline: Z-scheme Photocatalytic Mechanism Insight. Inorg. Chem. Front. 2018, 5, 3148–3154. [Google Scholar] [CrossRef]
  23. Kudo, A.; Miseki, Y. Heterogeneous Photocatalyst Materails for Water Splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, H.L.; Zhang, L.S.; Chen, Z.G.; Hu, J.Q.; Li, S.J.; Wang, Z.H.; Liu, J.S.; Wang, X.C. Semiconductor Heterojunction Photocatalysis: Design, Construction and Photocatalytic Performances. Chem. Soc. Rev. 2014, 43, 5234–5244. [Google Scholar] [CrossRef]
  25. Zhou, P.; Yu, J.G.; Jaroniec, M. All Solid State Z-Scheme Photocatalytic Systems. Adv. Mater. 2014, 6, 4920–4935. [Google Scholar] [CrossRef]
  26. Ong, W.J.; Tan, L.L.; Ng, Y.H.; Yong, S.T.; Chai, S.P. Graphitic Carbon Nitride (g-C3N4) Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer To Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  27. Li, Y.Y.; Yang, Q.M.; Wang, Z.M.; Wang, G.Y.; Zhang, B.; Zhang, Q.; Yang, D.F. Rapid Fabrication of SnO2 Nanoparticle Photocatalyst: Computational Understanding and Photocatalytic Degradaton of Organic Dye. Inorg. Chem. Front. 2018, 5, 3005–3014. [Google Scholar] [CrossRef]
  28. Selvan, R.K.; Perelshtein, I.; Perkas, N.; Gedanken, A. Synthesis of Hexagonal Shaped SnO2 Nanocrystals and SnO2@C Nanocomposites for Electrochemical Redox Supercapacitors. J. Phys. Chem. C 2008, 112, 1825–1830. [Google Scholar] [CrossRef]
  29. Sandesh, S.; Kristachar, P.K.R.; Manjunathan, P.; Halgeri, A.B.; Shanbhag, G.V. Synthesis of Biodiesel and Acetins by Transesterification Reactions using Novel CaSn(OH)6 Heterogeneous Base Catalyst. Appl. Catal. A Gen. 2016, 523, 1–11. [Google Scholar] [CrossRef]
  30. Thi, V.H.T.; Lee, B.K. Great Improvement on Tetracycline Removal using ZnO Rod-activated Carbon Fiber Composite Prepared with a Facile Microwave method. J. Haz. Mater. 2017, 324, 329–339. [Google Scholar]
  31. Xie, Z.J.; Feng, Y.P.; Wang, F.L.; Chen, D.N.; Zhang, Q.X.; Zeng, Y.Q.; Lv, W.Y.; Liu, G.G. Construction of Carbon Dots Modified MoO3/g-C3N4 Z-scheme Photocatalyst with Enhanced Visible Light Photocatalytic Activity for the Degradation of Tetracycline. Appl. Catal. B Environ. 2018, 229, 96–104. [Google Scholar] [CrossRef]
  32. Hailili, R.; Wang, Z.Q.; Li, Y.X.; Wang, Y.H.; Sharma, V.K.; Gong, X.Q.; Wang, C.Y. Oxygen Vacancies Induced Visible Light Photocatalytic Activities of CaCu3Ti4O12 with Controllable Morphologies for Antibiotic Degradation. Appl. Catal. B Environ. 2018, 221, 422–432. [Google Scholar] [CrossRef]
  33. Babu, B.; Harish, V.V.N.; Koutavarapu, R.; Shim, J.; Yoo, K. Enhanced Visible Light Active Photocatalytic Performance Using CdS Nanorods Decorated with Colloidal SnO2 Quantum Dots: Optimization of Core Shell Nanostructure. J. Ind. Eng. Chem. 2019, 76, 476–487. [Google Scholar] [CrossRef]
  34. Yang, Y.; Zeng, Z.T.; Zeng, G.M.; Huang, D.L.; Xiao, R.; Zhang, C.; Zhou, C.Y.; Xiong, W.P.; Wang, W.J.; Cheng, M.; et al. Ti3C2 Mxene/porous g-C3N4 Interfacial Schottky Junction for Boosting Spatial Charge Separation in Photocatalytic H2O2 Production. Appl. Catal. B Environ. 2019, 258, 117956–117966. [Google Scholar] [CrossRef]
  35. Gelderman, K.; Lee, L.; Donne, S.W. Flat Band Potential of a Semiconductor Using the Mott Schottky Equation. J. Chem. Edu. 2007, 84, 685–688. [Google Scholar] [CrossRef]
  36. Boltersdorf, J.; Zoellner, B.; Fancher, C.M.; Jones, J.L.; Maggard, P.A. Single and Double Site Substitutions in Mixed Metal Oxides: Adjusting the Band Edges Toward the Water Redox Couples. J. Phys. Chem. C 2016, 120, 19175–19188. [Google Scholar] [CrossRef]
  37. Ke, J.; Liu, J.; Sun, H.Q.; Zhang, H.Y.; Duan, X.G.; Liang, P.; Li, X.Y.; Tade, M.O.; Liu, S.M.; Wang, S.B. Facile Assembly of Bi2O3/Bi2S3/MoS2 n-p Heterojunction with Layered n-Bi2O3 and p-MoS2 for Enhanced Photocatalytic Water Oxidation and Pollutant Degradation. Appl. Catal. B Environ. 2017, 200, 47–55. [Google Scholar] [CrossRef]
  38. Lin, R.; Wan, J.W.; Xiong, Y.; Wu, K.L.; Cheong, W.C.; Zhou, G.; Wang, D.S.; Peng, Q.; Chen, C.; Li, Y.D. Quantitative Study of Charge Carrier Dynamics in Well-Defined WO3 Nanowires and Nanosheets: Insight into the Crystal Facet Effect in Photocatalysis. J. Am. Chem. Soc. 2018, 140, 9078–9082. [Google Scholar] [CrossRef]
  39. Xu, Z.Q.; Guan, Z.G.; Yang, J.J.; Li, Q.Y. Band Positions and Photoelectrochemical Properties of Solution Processed Silver Substituted Cu2ZnSnS4 Photocathode. ACS Appl. Energy Mater. 2019, 2, 2779–2785. [Google Scholar] [CrossRef]
  40. Lin, Y.M.; Li, D.Z.; Hu, J.H.; Xiao, G.C.; Wang, J.X.; Li, W.J.; Fu, X.Z. Highly Efficient Photocatalytic Degradation of Organic Pollutants by PANI-Modified TiO2 Composite. J. Phys. Chem. C 2012, 116, 5764–5772. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns with miller indices of several strongest peaks; (b) FT-IR spectra of SnO2, MgSn(OH)6, and MSOH-SO-3 heterojunction.
Figure 1. (a) XRD patterns with miller indices of several strongest peaks; (b) FT-IR spectra of SnO2, MgSn(OH)6, and MSOH-SO-3 heterojunction.
Nanomaterials 10 00053 g001
Figure 2. (a) HRTEM image; (b) HAADF and EDX spectrum of the MSOH-SO-3 heterojunction.
Figure 2. (a) HRTEM image; (b) HAADF and EDX spectrum of the MSOH-SO-3 heterojunction.
Nanomaterials 10 00053 g002
Figure 3. (a) UV–Vis DRS and plots of ( α h ν ) 2 versus the photon energy ( h ν ) for the band gap energies of (b) MgSn(OH)6, (c) MSOH-SO-3, and (d) SnO2.
Figure 3. (a) UV–Vis DRS and plots of ( α h ν ) 2 versus the photon energy ( h ν ) for the band gap energies of (b) MgSn(OH)6, (c) MSOH-SO-3, and (d) SnO2.
Nanomaterials 10 00053 g003
Figure 4. (a) Photodegradation of TC solution with the synthesized catalysts under UV irradiation. (b) Linear fitting with a pseudo-first order reaction model for TC photodegradation with the synthesized samples. (c) The fitted kinetic constant for photocatalytic reaction of TC solution. (d) Results for four cycles of reuse of MSOH-SO-3 for TC photodegradation.
Figure 4. (a) Photodegradation of TC solution with the synthesized catalysts under UV irradiation. (b) Linear fitting with a pseudo-first order reaction model for TC photodegradation with the synthesized samples. (c) The fitted kinetic constant for photocatalytic reaction of TC solution. (d) Results for four cycles of reuse of MSOH-SO-3 for TC photodegradation.
Nanomaterials 10 00053 g004
Figure 5. (a) Transient photocurrent responses under UV–Vis irradiation. (b) EIS Nyquist plots of MgSn(OH)6, SnO2, and MSOH-SO-3.
Figure 5. (a) Transient photocurrent responses under UV–Vis irradiation. (b) EIS Nyquist plots of MgSn(OH)6, SnO2, and MSOH-SO-3.
Nanomaterials 10 00053 g005
Figure 6. Trapping experiment for the reactive species during TC photodegradation over the MSOH-SO-3 heterojunction.
Figure 6. Trapping experiment for the reactive species during TC photodegradation over the MSOH-SO-3 heterojunction.
Nanomaterials 10 00053 g006
Figure 7. Mott–Schottky curves of (a) MgSn(OH)6 and (b) SnO2.
Figure 7. Mott–Schottky curves of (a) MgSn(OH)6 and (b) SnO2.
Nanomaterials 10 00053 g007
Figure 8. Schematic of the carrier transfer in a type-II heterojunction between MgSn(OH)6 and SnO2.
Figure 8. Schematic of the carrier transfer in a type-II heterojunction between MgSn(OH)6 and SnO2.
Nanomaterials 10 00053 g008

Share and Cite

MDPI and ACS Style

Li, Y.; Tian, X.; Wang, Y.; Yang, Q.; Diao, Y.; Zhang, B.; Yang, D. In Situ Construction of a MgSn(OH)6 Perovskite/SnO2 Type-II Heterojunction: A Highly Efficient Photocatalyst towards Photodegradation of Tetracycline. Nanomaterials 2020, 10, 53. https://doi.org/10.3390/nano10010053

AMA Style

Li Y, Tian X, Wang Y, Yang Q, Diao Y, Zhang B, Yang D. In Situ Construction of a MgSn(OH)6 Perovskite/SnO2 Type-II Heterojunction: A Highly Efficient Photocatalyst towards Photodegradation of Tetracycline. Nanomaterials. 2020; 10(1):53. https://doi.org/10.3390/nano10010053

Chicago/Turabian Style

Li, Yuanyuan, Xiaofang Tian, Yaoqiong Wang, Qimei Yang, Yue Diao, Bin Zhang, and Dingfeng Yang. 2020. "In Situ Construction of a MgSn(OH)6 Perovskite/SnO2 Type-II Heterojunction: A Highly Efficient Photocatalyst towards Photodegradation of Tetracycline" Nanomaterials 10, no. 1: 53. https://doi.org/10.3390/nano10010053

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

Article Metrics

Back to TopTop