Next Article in Journal
Some Insights into the Use of Heterogeneous Copper Catalysts in the Hydroprocessing of Levulinic Acid
Next Article in Special Issue
Visible-Light-Driven Peroxymonosulfate Activation for Accelerating Tetracycline Removal Using Co-TiO2 Nanospheres
Previous Article in Journal
Non-Thermal Plasma Incorporated with Cu-Mn/γ-Al2O3 for Mixed Benzene Series VOCs’ Degradation
Previous Article in Special Issue
Design and Architecture of P-O Co-Doped Porous g-C3N4 by Supramolecular Self-Assembly for Enhanced Hydrogen Evolution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controlled Synthesis of Ag-SnO2/α-Fe2O3 Nanocomposites for Improving Visible-Light Catalytic Activities of Pollutant Degradation and CO2 Reduction

1
Key Laboratory of Functional Inorganic Materials Chemistry, School of Chemistry and Materials Science, International Joint Research Center for Catalytic Technology, Heilongjiang University, Ministry of Education, Harbin 150080, China
2
National Institute of Vacuum Science and Technology, Islamabad 45400, Pakistan
3
Department of Chemistry, Abdul Wali Khan University Mardan, Khyber Pakhtunkhwa 23200, Pakistan
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(4), 696; https://doi.org/10.3390/catal13040696
Submission received: 13 March 2023 / Revised: 30 March 2023 / Accepted: 31 March 2023 / Published: 4 April 2023

Abstract

:
The key to developing highly active α-Fe2O3-based photocatalysts is to improve the charge separation and efficiently utilize the electrons with sufficient thermodynamic energy. Herein, α-Fe2O3 nanosheets (FO) were synthesized using a metal-ion-intervened hydrothermal method and then coupled with SnO2 nanosheets (SO) to obtain SO/FO nanocomposites. Subsequently, nanosized Ag was selectively loaded on SO using the photo-deposition method to result in the ternary Ag-SO/FO nanocomposites. The optimal nanocomposite could realize the efficient aerobic degradation of 2,4-dichlorophenol as a representative organic pollutant under visible-light irradiation (>420 nm), exhibiting nearly six-fold degradation rates of that for FO. Additionally, the Ag-SO/FO photocatalyst is also applicable to the visible-light degradation of other organic pollutants and even CO2 reduction. By using steady-state surface photovoltage spectroscopy, fluorescence spectroscopy, and electrochemical methods, the photoactivity enhancement of Ag-SO/FO is principally attributed to the improved charge separation by introducing SO as an electron platform for the high-energy-level electrons of FO. Moreover, nanosized Ag on SO functions as a cocatalyst to further improve the charge separation and facilitate the catalytic reduction. This work provides a feasible design strategy for narrow-bandgap semiconductor-based photocatalysts by combining an electron platform and a cocatalyst.

1. Introduction

With the rapid development of industry, human beings are faced with increasingly serious environmental pollution. Among various forms of pollution, water pollution caused by discarding organic pollutants into water reservoirs is found throughout the world, which damages the natural ecosystem and harms the health of the living [1,2]. In particular, the chlorophenols and corresponding derivatives such as 2,4-dichlorophenol (2,4-DCP) are the representative organic pollutants for water bodies, which are commonly used in manufacturing herbicides, fungicides, antiseptics, and wood preservations [3,4,5]. The US Environmental Protection Agency has listed 2,4-DCP as one of the most toxic pollutants with potential carcinogenic and mutagenic risks to human health [6,7,8]. Therefore, it is highly desirable to develop green and efficient techniques to degrade organic pollutants, such as 2,4-DCP, in water.
Semiconductor photocatalysis, capable of degrading organic pollutants by the photocatalytic reaction using sustainable solar energy, is regarded as one of the most promising technologies. The key to realizing such a process is to design and fabricate efficient photocatalyst materials. For semiconductors such as photocatalysts, visible-light responsive ones are advantageous in terms of light absorption [9,10,11]. Taking the hematite α-Fe2O3 as an example, it is a narrow-bandgap metal oxide semiconductor (2.2 eV), which is stable, economic, and eco-friendly, and hence has been extensively investigated as a promising photocatalyst candidate [12,13,14]. Although the morphology control of two-dimensional (2D) α-Fe2O3 to obtain corresponding nanosheets could enlarge the specific surface area and improve the charge separation to a certain extent by shortening the charge transfer distance, its photocatalytic efficiency is still low due to poor charge separation, the unavailability of electrons with sufficient thermodynamic energy, and the lack of catalytic sites for target reactions [15,16,17].
For the photogenerated electrons of α-Fe2O3 nanosheets (denoted as FO for short) excited under visible-light irradiation, only some of them possess a high energy level above the conduction band (CB) of FO which could meet the thermodynamic requirement for the reduction of reactants such as O2, etc. [18]. However, these high-energy-level electrons (HELEs) tend to quickly relax to the CB bottom and then recombine with the holes [19]. Therefore, it is rational to efficiently use the HELEs to realize the photoactivity enhancement of FO under visible-light irradiation. Currently, one of the most applied strategies is to introduce another semiconductor to couple with FO, forming a heterojunction. Traditional wide-bandgap semiconductors (WBS), such as ZnO, TiO2, and SnO2, with relatively high CB potentials (<~ 0 eV) become suitable choices [20,21,22,23]. Thus, the coupling of WBSs with FO would function as an electron platform for the HELEs, resulting in improved charge separation. Considering that SnO2 has high electron mobility and suitable CB potential at around 0 eV, it is feasible to construct dimension-matched 2D/2D SnO2/α-Fe2O3 heterojunctions (denoted as SO/FO) [20,24].
Moreover, catalytic efficiency is another crucial factor in determining photocatalytic performance [18]. If photocatalytic aerobic organic pollutant degradation is the target reaction, the catalytic process for the O2 photoreduction significantly affects the overall degradation efficiency [25]. Therefore, it is necessary to introduce an effective cocatalyst on SO in SO/FO to further enhance photoactivity. Noble metals, such as Ag, Pt, and Au nanoparticles (NPs), have been reported as favorable cocatalysts in photocatalysis [26,27,28], performing the dual functions of electron acceptors and catalytic sites for activating reactants. Given that the economic cost of Ag is relatively low, selectively loaded Ag NPs on SO in the SO/FO heterojunctions would further facilitate electron transfer and catalytic reduction.
Based upon the above design scheme, herein, ultrathin FO was synthesized using the metal-ion-intervened hydrothermal method, which was coupled with SO obtained via the hydrothermal method to result in 2D/2D SO/FO heterojunctions. Subsequently, nanosized Ag was selectively loaded on SO to eventually obtain the ternary Ag-SO/FO nanocomposites. Under visible-light irradiation (>420 nm), the optimal Ag-SO/FO photocatalyst showed nearly six-fold photoactivity enhancement compared with FO for the aerobic degradation of 2,4-DCP. The steady-state surface photovoltage spectra, fluorescence spectra, and electrochemical results, etc., indicate that the enhanced photoactivity is due to improved charge separation by introducing SO as the electron platform and nanosized Ag as an electron acceptor. Additionally, nanosized Ag could facilitate catalytic reduction as a cocatalyst. This work improves the design scheme for efficient heterojunction photocatalysts.

2. Results

2.1. The Effect of Co-Modification by SO and Nanosized Ag on Photocatalytic Performance

As described in the experimental section, the FO-based nanocomposites were synthesized and then evaluated for photocatalytic aerobic degradation of organic pollutant 2,4-DCP under visible-light irradiation. As shown in Figure 1a, pure FO demonstrated a degradation rate of around 16%.
After coupling SO, the photoactivity of xSO/FO samples, where x% indicates the mass ratio of SO to FO, show obvious enhancement compared with FO (Figure S1a). In particular, 5SO/FO was the best binary sample and provided a degradation rate of 51%. Based on 5SO/FO, Ag NPs with different amounts were photo-deposited to obtain the ternary yAg-5SO/FO nanocomposites, where y% indicates the mass ratio of Ag relative to 5SO/FO. Among all Ag-modified samples, 2Ag-5SO/FO showed the largest degradation rate, which reached approximately six-fold that of FO (Figure S1b). According to the above results, the co-modification of SO and nanosized Ag with suitable amounts could effectively enhance the photocatalytic activities of FO. Considering that stability is another critical index used to evaluate photocatalysts, the recyclability of 2Ag-5SO/FO as the optimal photocatalyst was examined for five consecutive runs. As shown in Figure 1b, no obvious photoactivity decrease could be found after 5 runs for 10 h, indicating that the 2Ag-5SO/FO photocatalyst possesses favorable stability under reaction conditions.

2.2. Structural and Surface Characterization

The structural information of as-synthesized FO-based nanocomposites was collected by several characterization methods. The XRD patterns of FO, SO, 5SO/FO, and 2Ag-5SO/FO are demonstrated in Figure S2a. As shown in Figure S2a, the diffraction peaks of FO were 33.2, 35.62, 39.2, 40.8, 49.5, 54.0, 63.9, and 66.0°, which agree well with the ones of the standard α-Fe2O3 JCPDS No. 33-0664. For SO, the peaks at 26.6, 33.8, 37.9, 51.8, 54.7, 57.8, and 61.9° are ascribed to SnO2 (JCPDS No. 41-1445) [14,17]. The peaks for 5SO/FO were consistent with those for FO. This indicates that the introduction of a tiny amount of SO does not change the crystalline structure of FO. Similarly, with the further introduction of nanosized Ag, for 2Ag-5SO/FO, no extra peaks appeared except those for FO [29]. Moreover, the light absorption of the FO-based samples was investigated by UV–vis DRS. Figure S2b depicts that neither SO nor nanosized Ag affected the light absorption of FO.
Furthermore, the morphology of typical samples was assured by TEM. As shown in Figure S3a, FO takes on 2D morphology, in agreement with [16,17]. Additionally, Figure S3b verifies that SO is also a 2D nanosheet, whose size is smaller than that of FO (Figure S3b). A heterojunction structure could be observed for 5SO/FO (Figure S3c), indicating the successful combination between FO and SO. For 2Ag-5SO/FO, in addition to the heterojunctions, such as 5SO/FO, the Ag NPs with an average diameter of around 5 nm were observed on SO (Figure 2a), evidencing the selective location of nanosized Ag on SO. Next, the chemical state of nanosized Ag was revealed by XPS. The XPS peaks at the binding energies of 373.9 and 367.8 eV were attributed to Ag 3d3/2 and Ag 3d5/2, respectively (Figure 2b) [30]. In summary, the 2Ag-5SO/FO nanocomposites were synthesized by co-modifying with SO and nanosized Ag, respectively.

2.3. Photogenerated Charge Separation and Catalytic Function

The charge separation is the key factor determining the photoactivity for semiconductor photocatalysis. Accordingly, the charge separation properties of typical samples were characterized to investigate the effect of SO and nanosized Ag as modifiers. When a photocatalyst capable of oxidizing H2O is excited in the aqueous environment, it will produce hydroxyl radicals as captured by the coumarin to produce 7-hydroxycoumarin, whose fluorescence signal can be quantified to reflect the charge separation situation of the as-tested sample. The fluorescence spectra (FS) were demonstrated in Figure 3a and Figure S4.
5SO/FO presents the largest FS signal among xSO/FO samples, indicating that the loading of SO on FO improves the charge transfer by forming a heterojunction. After introducing nanosized Ag, for yAg-5SO/FO samples 2Ag-5SO/FO displays the largest signal, implying that an optimal amount of nanosized Ag could further improve the charge separation. Moreover, steady-state surface photovoltage spectroscopy (SS-SPS) is an advanced photophysical technique which can directly reveal the photogenerated charge separation and recombination by means of the surface potential difference of a semiconductor before and after illumination. Generally, the SPS intensity and charge separation are positively correlated. As shown in Figure 3b, the SS-SPS signal intensity sequence is FO < 5SO/FO < 2Ag-5SO/FO, consistent with the FS results [18].
Additionally, the electrochemical impendence spectra (EIS) and linear sweep voltammetry (LSV) curves were collected for typical samples under visible-light irradiation. As shown in Figure S5a and Figure 4a inset, 5SO/FO displayed the lowest semicircle radii among the SO-modified samples, suggesting that 5SO/FO has higher charge mobility/transport than FO. In Figure S5b, the introduction of nanosized Ag further reduces the capacitive arc radii of 5SO/FO, proving that Ag facilitates electron transfer. The LSV curves of FO, 5SO/FO, and 2Ag-5SO/FO are shown in Figure 4a. The photocurrent density signal intensities show the sequence of FO < 5SO/FO < 2Ag-5SO/FO, demonstrating the positive effects of SO and Ag on charge separation. Based upon the above results on the charge separation, it can be found that SO does improve the charge separation as the design scheme. When irradiated by visible light, only FO with a narrow bandgap could be excited. Normally, the photoelectrons of FO would be excited to different energy levels, some of which possess a high-energy level, denoted as HELEs, and hence capable of inducing photoreduction thermodynamically. However, the HELEs would relax to the valance band of FO quickly to recombine with the holes. Therefore, the improved charge separation of 5SO/FO indicates that the HELEs transfer from FO to SO, where SO functions as an electron platform [19,20]. In addition, the further improved charge separation for 2Ag-5SO/FO is caused by nanosized Ag on SO acting as an electron acceptor to facilitate the electron transfer. Accordingly, the introduction of both electron platform and electron acceptor effectively improves the charge separation to enhance the photoactivity.
Except for light absorption and charge separation, the catalytic efficiency for O2 photooxidation is another essential factor affecting the photoactivity for the aerobic degradation of organic pollutants. Thus, the electrochemical reduction curves in the O2 bubbled system were collected to investigate the catalytic function for O2 reduction (Figure 4b). It was demonstrated that the onset potential gradually decreased in the sequence of 2Ag-5FO/SO < 5SO/FO < FO. This indicates that the introduction of SO benefits the catalytic oxidation of O2. Noteworthily, 2Ag-5SO/FO showed the best catalytic function for O2 reduction, which is attributed to the catalytic role of nanosized Ag.

2.4. Discussion

To further reveal the catalytic process mechanism for the visible-light aerobic degradation of 2,4-DCP, the trapping experiments were performed for typical samples by adding different scavengers (Figure S6). Specifically, ethylenediamine tetra-acetic acid disodium salt (EDTA-2Na), isopropyl alcohol (IPA), and benzoquinone (BQ) were used to trap the holes, hydroxyl radicals, and superoxo species, respectively. As shown in Figure S6, the existence of either EDTA-2Na and IPA significantly decreased the visible-light degradation rates of 2,4-DCP for FO, 5SO/FO and 2Ag-5SO/FO. This proves that holes and hydroxyl radicals are the main active species that participate the degradation of 2,4-DCP.
The applicability of fabricated samples was also checked for the representative industrial organic dyes BPA and MO (Figure 5a and Figures S7 and S8). Under identical reaction conditions with those for the photocatalytic aerobic 2,4-DCP degradation, the degradation rates of FO, 5SO/FO, and 2Ag-5SO/FO for BPA reached 13%, 35%, and 61%, respectively. On another aspect, the degradation rates of FO, 5SO/FO, and 2Ag-5SO/FO for MO reached 16%, 45%, and 88%, respectively. Compared with FO, 2Ag-5SO/FO realized 4.7-fold and 6-fold photoactivity enhancement for the BPA and MO degradation, respectively. The excellent photocatalytic performances of 2Ag-5SO/FO photocatalyst originate from the favorable charge separation and sufficient thermodynamic energy, and the catalytic function of Ag NPs [31]. Therefore, it can be inferred that 2Ag-5SO/FO might be capable of catalyzing visible-light CO2 reduction. As expected, under visible light irradiation, the CO and CH4 production rates for 2Ag-5SO/FO achieved 7.6 µmol g−1 h−1 and 1.5 µmol g−1 h−1, respectively, whose CO2 conversion rate was nearly 8-fold of that for FO (Figure 5b). The above photocatalytic performances indicate that the design principle of Ag-SO/FO photocatalysts based on improving charge separation is applicable for both the aerobic degradation of organic pollutants and the CO2 reduction.
Based on the above, the overall photocatalytic mechanism of the Ag-SO/FO photocatalyst is illustrated in Scheme 1. Under the visible-light irradiation, the photoelectrons of excited FO possessing sufficient thermodynamic energy transferred to the CB of as-coupled SO by crossing the interface, which further transferred to supported Ag NPs, resulting in improved charge separation. Moreover, nanosized Ag could function as the cocatalyst to activate the reactants. For the aerobic degradation of the organic pollutants, Ag could adsorb and activate O2 to facilitate the catalytic reduction of O2 with the consequence of speeding up electron consumption, which is beneficial to charge separation. Accordingly, the resultant holes and hydroxyl radicals would dominate the degradation of organic pollutants [32,33].

3. Materials and Methods

All the chemicals and reagents were of analytical grade and used as received.

3.1. Synthesis of FO

Ultrathin FO was prepared through the metal-ion-intervened hydrothermal method as previously reported [17]. In particular, Fe(NO3)3·9H2O (0.40 g) and Al2(SO4)3·18H2O (0.16 g) were dispersed in the deionized (DI) water (10 mL), which was then magnetically stirred for 10 min to obtain a uniform liquid mixture. Subsequently, triethylamine (3 mL) was added dropwise under continuous stirring for 30 min. The resulting mixture was transferred to 50 mL Teflon-lined stainless-steel autoclaves and heated at 160 °C for 24 h, then cooled naturally. The product was collected by centrifugation, cleaned with absolute ethanol and DI water several times, and finally dried in the oven at 60 °C for 12 h. Then, FO was crushed to powder and calcined at 350 °C for 2 h.

3.2. Synthesis of SO

The hydrothermal method was utilized to synthesize SnO2 nanosheets abbreviated as (SO). In the typical synthesis procedure, 4.5 mM of SnCl2·2H2O was added to 15 mL of DI H2O and magnetically stirred for 15 min at room temperature to form the transparent solution. Then, 30 mL of 0.4 M NaOH aqueous solution was added drop-by-drop, maintained at pH = 10, and stirred for the next 2 h to achieve a translucent milky solution. The solution was then added to a 50 mL autoclave heated in an oven at 180 °C for 12 h. After natural cooling to room temperature, the precipitates were collected by centrifugations and alternatively washed 5 times with DI H2O and ethanol. The precipitate was oven-dried at 80 °C and ground into powder to obtain SO.

3.3. Synthesis of SO/FO

For fabrication of 2D/2D (xSO/FO) nanocomposites, 1g of FO was dispersed in an equal volume ratio of aqueous ethanol mixture (50:50 mL) and alternatively sonicated and stirred for 30 min. The above-mentioned dispersion was then supplemented with SO in an optimized mass percent ratio (x = 1, 3, 5, and 7%) and the mixture was rapidly stirred for 2 h. The solution was dried in a water bath at 85 °C to completely evaporate all solvents and finally calcined at 350 °C for 2 h in a muffle furnace. The samples were marked as xSO/FO, where x% indicates the mass ratio of SO to FO.

3.4. Synthesis of Ag-SO/FO

For the photo-deposition of nanosized Ag on SO in 5SO/FO nanocomposite, 1 g of freshly prepared 5SO/FO was dispersed in an equal volume ratio of aqueous ethanol mixture (50: 50 mL), stirred, and sonicated for 30 min. Subsequently, the AgNO3 solution was added into the above dispersions to prepare different mass ratios of Ag (0.5, 1, 2, and 3%) loaded on the surface of 5SO/FO and stirred for 30 min to achieve homogeneous dispersion. The beaker was then sealed with sealing film and bubbled with pure N2 gas for 1 h to flush out adsorbed O2 and facilitate an inert atmosphere for the photo-deposition reaction. The mixture solution was irradiated under a 300 W Xe lamp with a UV-cut-off filter (λ > 420 nm) for 3 h, with continuous stirring and water cooling to avoid evaporation. Finally, the mixture solution was centrifuged, and the resulting solid was washed with ethanol and DI water several times. Then, the solid was dried in the vacuum oven at 80 °C for 12 h. The resulting nanocomposites were denoted as yAg-5SO/FO, where y% indicates the mass ratio of nanosized Ag relative to 5FO/SO.

3.5. Photocatalytic Activity for Pollutants Degradation

The as-fabricated nanocomposites were evaluated for the degradation of 2,4-dichlorophenol (2,4-DCP) pollutants. For the photocatalytic degradation under visible-light irradiation, 0.1 g of the photocatalyst was dispersed in 50 mL 2,4-DCP solution with a concentration of 20 mg L−1 and stirred in the dark for 30 min to achieve adsorption–desorption equilibrium. After that, the mixture was placed in an open glass reactor under continuous stirring and irradiation by a 150 W xenon lamp with a UV-cut-off filter (λ > 420 nm) for 1.5 h. Then, 5 mL of supernatant was centrifuged and analyzed for the remaining concentration of 2,4-DCP with a spectrophotometer. In addition, the same experimental protocols were also followed for the photocatalytic aerobic degradation of bisphenol A (BPA) and methyl orange (MO).

3.6. Photocatalytic Activity for CO2 Conversion

In the following experimental procedure, 20 mg of photocatalyst and 5 mL of DI water were added to a 100 mL steel cylinder reactor with an exposed irradiation surface of 3.5 cm2. For the next 20 min, pure CO2 gas was bubbled into the mixed solution in the cylindrical reactor, then the CO2 gas supply was switched off and the system was continuously stirred for the next 10 min to achieve equilibrium. The cylindrical photocatalytic reactor was irradiated with a 300 W xenon lamp (radiation source) for the next 4 h, with a UV-cut-off filter (λ > 420 nm) under a continuous flow of water to maintain photocatalytic reactor temperature and avoid solvent evaporation. Moreover, the incident light intensity was 380 mW/cm2. After a certain period, 0.25 mL of the gaseous mixture was taken from the top head of the reactor using a syringe and tested for the production of carbon-containing products, such as CO and CH4, using a gas chromatography model (GC-7920, Beijing).

3.7. Scavenger Experiments

Trapping experiments in the presence of three different types of scavengers, including isopropyl alcohol (IPA), ethylenediamine tetra-acetic acid disodium salt (EDTA-2Na), and benzoquinone (BQ), were performed to determine the reactive species involved in the degradation of aerobic organic pollutants 2,4-DCP under visible-light irradiation. In the typical experimental procedure, 0.1 g of the photocatalyst was dispersed in 50 mL of 20 mg/L solutions of 2,4-DCP and stirred in the dark for 10 min, then 0.8 mL of 1 mM of each of the following scavengers (i.e., IPA, EDTA-2Na, and BQ solution) were added separately to the above-mentioned solution and further stirred in the dark for 20 min to achieve adsorption–desorption equilibrium. After that, the mixture solution was exposed for 1.5 h in an open glass reactor with continuous stirring to a 150 W xenon lamp with a UV-cut-off filter (λ > 420 nm). Then, 5 mL of supernatant was centrifuged and analyzed for the remaining concentration of 2,4-DCP with a spectrophotometer and check the effect of added scavengers on the activity.

3.8. Recycle Experiment

For recycling testing, the photocatalyst was separated from the mixed solution by centrifugation, washed with acetone, followed by DI water, and then dried in the vacuum oven overnight at 70 °C, ready to be utilized for the following cycle.

3.9. Fluorescence Spectra Measurement

Hydroxyl radicals are essential in the photocatalytic process. To determine the number of hydroxyl radicals generated, a 50 mg photocatalyst was dispersed in 40 mL of 0.001 M coumarin solution and stirred in the dark for 30 min to attain adsorption–desorption equilibrium. The solution was then exposed for 1 h to a 150 W xenon lamp with a UV-cut-off filter (>420 nm). Finally, 5 mL of supernatant was centrifuged and analyzed for the production of the hydroxyl radicals using a spectrofluorometer F-7000 FL at excitation wavelengths of 332 nm.

3.10. Characterization

The as-prepared nanocomposite materials were analyzed by different characterization techniques. The crystalline structures were studied by Rigaku D/max-rB diffractometer (Rigaku, Japan) with Cu Kα radiation. The surface morphologies were checked by transmission electron microscopy (TEM) on an FEI Tecnai G2 S-Tw (USA) with an acceleration voltage of 200 kV. The optical bandgap was measured by realizing the UV–visible diffused reflectance spectra (UV–vis DRS) with a Model Shimadzu UV2550 spectrophotometer (Japan), with BaSO4 as a reference. The surface composition and elemental chemical state of the samples were determined using an X-ray photoelectron spectroscopy (XPS) technique with a Model VG ESCALAB apparatus (Thermo Fisher Scientific, Waltham, MA, USA) and an Mg K X-ray source, and the binding energies were calibrated concerning the signal for adventitious carbon (BE = 284.6 eV). The charge separation property was tested using steady-state surface photovoltage spectroscopy (SS-SPS) using home-built equipment with a lock-in amplifier (SR830, USA) synchronized with a Xenon lamp (CHF XQ500 W) at room temperature. The photoelectrochemical and electrochemical investigation was carried out in a three-electrode configuration with a working electrode (catalyst pasted on FTO glass), a counter electrode (Pt foil), and a reference electrode (Ag/AgCl electrode) in an electrolyte of 0.2 M Na2SO4. To prepare the electrode, 20 mg of the catalyst was dispersed in a mixture of 20 μL (5 wt% Nafion ionomers) and 0.15 mL ethanol. This mixture was then sonicated for 30 min and then stirred continuously for 72 h to obtain a homogenous slurry. Cleaned and dried FTO glasses were cut into 3 × 4 cm. The samples were pasted by the doctor blade method and then calcined for 30 min at 300 °C in a muffle furnace. The FTO glass with coated samples was submerged in a quartz cell containing 60 mL of 0.2 M Na2SO4 solution, coupled with electrodes, and purged with highly pure O2 gas for 20 min. The IVIUM V13806 electrochemical workstation (Netherland) was then used to measure electrochemical curves in the range of (0.2 to -2 V). Gamry’s electrochemical workstation was used to record electrochemical impedance spectra at an AC voltage magnitude of 0.05 V and a frequency range of 103–10 Hz. A Gamry electrochemical workstation was also utilized to record the linear sweep voltammetry (LSV) under light irradiation in a quartz cell. A 500 W xenon lamp with a UV-cut-off filter (λ > 420 nm) was employed as the irradiation source. All electrochemical studies were conducted at standard room temperature.

4. Conclusions

In conclusion, a rationally designed ternary Ag-SO/FO heterojunction photocatalyst was synthesized and applied for efficient aerobic degradation of 2,4-DCP under visible-light irradiation. Compared with pristine FO, the optimal Ag-SO/FO photocatalyst exhibited nearly six-fold photoactivity enhancement, mainly due to the improved charge separation. Specifically, the HELEs of FO would transfer to the CB of SO and then to as-loaded Ag NPs. As the cocatalyst, Ag NPs could catalyze the O2 reduction to effectively consume the electrons, which also benefits the charge separation. The photocatalyst is also applicable to the visible-light degradation of other organic pollutants (such as BPA and MO) and even CO2 reduction. This work highlights the importance of modulating photoelectrons for the rational design of heterojunction photocatalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13040696/s1, Figure S1: Photocatalytic aerobic degradation for 2,4-DCP over (a) FO, xSO/FO, and (b) yAg-5SO/FO samples under visible-light irradiation.; Figure S2: (a) XRD patterns and (b) DRS spectra of FO, 5SO/FO, and 2Ag-5SO/FO, respectively.; Figure S3: TEM images of (a) FO, (b) SO, and (c) 5SO/FO, respectively.; Figure S4: FS spectra related to OH radical of (a) FO, xSO/FO, and (b) yAg-5SO/FO.; Figure S5: EIS Nyquist plots related to charge separation of (a) FO, xSO/FO, and (b) yAg-5SO/FO.; Figure S6: Photocatalytic activities for 2,4-DCP degradation of FO, 5SO/FO, and 2Ag-5SO/FO, respectively, under visible-light irradiation in the presence of different scavengers. Ethylenediamine tetra-acetic acid disodium salt (EDTA-2Na) is the scavenger for the photogenerated (h+), isopropyl alcohol (IPA) for (OH) and benzoquinone (BQ) for O2-, respectively.; Figure S7: Photocatalytic aerobic degradation for BPA over (a) FO, xSO/FO, and (b) yAg-5SO/FO samples under visible-light irradiation.; Figure S8: Photocatalytic aerobic degradation for MO over (a) FO, xSO/FO, and (b) yAg-5SO/FO samples under visible-light irradiation.

Author Contributions

W.A.: Formal analysis, Data curation, Writing—original draft; Z.L.: Methodology, Validation, Investigation, Visualization, Funding acquisition; L.B.: Conceptualization, Writing—review and editing, Visualization, Project administration, Funding acquisition; M.Z.A.: Methodology, Resources; A.Z.: Methodology, Resources; Y.Q.: Validation, Formal analysis; S.S.: Validation; L.J.: Conceptualization, Supervision, Project administration, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

We are grateful for financial support from the National Natural Science Foundation of China (U2102211), the Fundamental Research Foundation for Universities of Heilongjiang Province (2021-KYYWF-0004), and the Innovative talent training plan in colleges and universities of Heilongjiang province (UNPYSCT-2020003) Thanks to the Science Fund for Distinguished Young Scholars of Heilongjiang University (JCL202102).

Data Availability Statement

Not applicable.

Acknowledgments

We thank Muhammad Maqsood (Director General of NINVAST-NCP Pakistan), Muhammad Ahmad Yar (Director of NINVAST-NCP Pakistan) Jawad Ali (SSO), and Muhammad Sajid (SSO) for providing the use of lab facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Van, K.N.; Huu, H.T.; Nguyen Thi, V.N.; Le Thi, T.L.; Truong, D.H.; Truong, T.T.; Dao, N.N.; Vo, V.; Tran, D.L.; Vasseghian, Y. Facile construction of S-scheme SnO2/g-C3N4 photocatalyst for improved photoactivity. Chemosphere 2022, 289, 133120. [Google Scholar] [CrossRef] [PubMed]
  2. Chu, M.; Hu, K.; Wang, J.; Liu, Y.; Ali, S.; Qin, C.; Jing, L. Synthesis of g-C3N4-based photocatalysts with recyclable feature for efficient 2,4-dichlorophenol degradation and mechanisms. Appl. Catal. B Environ. 2019, 243, 57–65. [Google Scholar] [CrossRef]
  3. Yu, X.; Cabooter, D.; Dewil, R. Effects of process variables and kinetics on the degradation of 2,4-dichlorophenol using advanced reduction processes (ARP). J. Hazard. Mater. 2018, 357, 81–88. [Google Scholar] [CrossRef] [PubMed]
  4. Ali, W.; Ullah, H.; Zada, A.; Muhammad, W.; Ali, S.; Shaheen, S.; Alamgir, M.K.; Ansar, M.Z.; Khan, Z.U.; Bilal, H.; et al. Synthesis of TiO2 modified self-assembled honeycomb ZnO/SnO2 nanocomposites for exceptional photocatalytic degradation of 2,4-dichlorophenol and bisphenol A. Sci. Total Environ. 2020, 746, 141291. [Google Scholar] [CrossRef] [PubMed]
  5. Humayun, M.; Hu, Z.; Khan, A.; Cheng, W.; Yuan, Y.; Zheng, Z.; Fu, Q.; Luo, W. Highly efficient degradation of 2,4-dichlorophenol over CeO2/g-C3N4 Composites under visible-light irradiation: Detailed reaction pathway and mechanism. J. Hazard. Mater. 2019, 364, 635–644. [Google Scholar] [CrossRef]
  6. Guo, L.; Yang, J.; Zhang, H.; Wang, R.; Xu, J.; Wang, J. Highly Enhanced Visible-Light Photocatalytic Activity via a Novel Surface Structure of CeO2/g−C3N4 toward Removal of 2,4-Dichlorophenol and Cr(VI). ChemCatChem 2021, 13, 2034–2044. [Google Scholar] [CrossRef]
  7. Huang, Z.; Chen, G.; Zeng, G.; Guo, Z.; He, K.; Hu, L.; Wu, J.; Zhang, L.; Zhu, Y.; Song, Z. Toxicity mechanisms and synergies of silver nanoparticles in 2,4-dichlorophenol degradation by phanerochaete chrysosporium. J. Hazard. Mater. 2017, 321, 37–46. [Google Scholar] [CrossRef]
  8. Raziq, F.; Humayun, M.; Ali, A.; Wang, T.; Khan, A.; Fu, Q.; Luo, W.; Zeng, H.; Zheng, Z.; Khan, B.; et al. Synthesis of S-doped porous g-C3N4 by using ionic liquids and subsequently coupled with au-tio2 for Exceptional Cocatalyst-Free Visible-Light Catalytic Activities. Appl. Catal. B Environ. 2018, 237, 1082–1090. [Google Scholar] [CrossRef]
  9. Li, X.; Anwer, S.; Guan, Q.; Anjum, D.H.; Palmisano, G.; Zheng, L. Coupling Long-Range Facet Junction and Interfacial Heterojunction via Edge-Selective Deposition for High-Performance Z-Scheme Photocatalyst. Adv. Sci. 2022, 9, 2200346. [Google Scholar] [CrossRef]
  10. Gong, Y.; Quan, X.; Yu, H.; Chen, S.; Zhao, H. Enhanced photocatalytic performance of a two-dimensional BiOIO3/g-C3N4 heterostructured composite with a Z-scheme configuration. Appl. Catal. B Environ. 2018, 237, 947–956. [Google Scholar] [CrossRef]
  11. Di, J.; Xiong, J.; Li, H.; Liu, Z. Ultrathin 2D Photocatalysts: Electronic-Structure Tailoring, Hybridization, and Applications. Adv. Mater. 2018, 30, 1704548. [Google Scholar] [CrossRef]
  12. Jiang, Z.; Wan, W.; Li, H.; Yuan, S.; Zhao, H.; Wong, P.K. A Hierarchical Z-Scheme α-Fe2O3/g-C3N4 Hybrid for Enhanced Photocatalytic CO2 Reduction. Adv. Mater. 2018, 30, 1706108. [Google Scholar] [CrossRef]
  13. Geng, Y.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. Z-Scheme 2D/2D α-Fe2O3/g-C3N4 heterojunction for photocatalytic oxidation of nitric oxide. Appl. Catal. B Environ. 2021, 280, 119409. [Google Scholar] [CrossRef]
  14. Yan, S.; Shi, Y.; Tao, Y.; Zhang, H. Enhanced persulfate-mediated photocatalytic oxidation of bisphenol A using bioelectricity and a g-C3N4/Fe2O3 heterojunction. Chem. Eng. J. 2019, 359, 933–943. [Google Scholar] [CrossRef]
  15. She, X.; Wu, J.; Xu, H.; Zhong, J.; Wang, Y.; Song, Y.; Nie, K.; Liu, Y.; Yang, Y.; Rodrigues, M.T.F.; et al. High Efficiency Photocatalytic Water Splitting Using 2D α-Fe2O3/g-C3N4 Z-Scheme Catalysts. Adv. Energy Mater. 2017, 7, 1700025. [Google Scholar] [CrossRef]
  16. Mu, Z.; Chen, S.; Wang, Y.; Zhang, Z.; Li, Z.; Xin, B.; Jing, L. Controlled Construction of Copper Phthalocyanine/α-Fe2O3 Ultrathin S-Scheme Heterojunctions for Efficient Photocatalytic CO2 Reduction under Wide Visible-Light Irradiation. Small Sci. 2021, 1, 2100050. [Google Scholar] [CrossRef]
  17. Jin, Y.; Dang, L.; Zhang, H.; Song, C.; Lu, Q.; Gao, F. Synthesis of unit-cell-thick α-Fe2O3 nanosheets and their transformation to γ-Fe2O3 nanosheets with enhanced LIB performances. Chem. Eng. J. 2017, 326, 292–297. [Google Scholar] [CrossRef]
  18. Bian, J.; Qu, Y.; Fazal, R.; Li, X.; Sun, N.; Jing, L. Accepting Excited High-Energy-Level Electrons and Catalyzing H2 Evolution of Dual-Functional Ag-TiO2 Modifier for Promoting Visible-Light Photocatalytic Activities of Nanosized Oxides. J. Phys. Chem. C 2016, 120, 11831–11836. [Google Scholar] [CrossRef]
  19. Li, Z.; Luan, P.; Zhang, X.; Qu, Y.; Raziq, F.; Wang, J.; Jing, L. Prolonged lifetime and enhanced separation of photogenerated charges of nanosized α-Fe2O3 by coupling SnO2 for efficient visible-light photocatalysis to convert CO2 and degrade acetaldehyde. Nano Res. 2017, 10, 2321–2331. [Google Scholar] [CrossRef]
  20. Qu, Y.; Li, Z.; Sun, N.; Zhang, X.; Chen, S.; Jing, L. Visible-light induced electron modulation to improve photoactivities of coral-like Bi2WO6 by coupling SnO2 as a Proper energy platform. Catal. Today 2019, 327, 288–294. [Google Scholar] [CrossRef]
  21. Chen, S.; Yan, R.; Zhang, X.; Hu, K.; Li, Z.; Humayun, M.; Qu, Y.; Jing, L. Photogenerated electron modulation to dominantly induce efficient 2,4-dichlorophenol degradation on BiOBr nanoplates with different phosphate modification. Appl. Catal. B Environ. 2017, 209, 320–328. [Google Scholar] [CrossRef]
  22. Humayun, M.; Sun, N.; Raziq, F.; Zhang, X.; Yan, R.; Li, Z.; Qu, Y.; Jing, L. Synthesis of ZnO/Bi-doped porous LaFeO3 nanocomposites as highly efficient nano-photocatalysts dependent on the enhanced utilization of visible-light-excited electrons. Appl. Catal. B Environ. 2018, 231, 23–33. [Google Scholar] [CrossRef]
  23. Sun, N.; Qu, Y.; Yang, C.; Yang, Z.; Yan, R.E.W.; Zhang, Z.; Li, Z.; Li, H.; Khan, I.S.R.; Jing, L.; et al. Efficiently photocatalytic degradation of monochlorophenol on In-Situ fabricated BiPO4/β-Bi2O3 heterojunction microspheres and O2-Free Hole-induced selective dechloridation conversion with H2 Evolution. Appl. Catal. B Environ. 2020, 263, 118313. [Google Scholar] [CrossRef]
  24. Yang, J.; Sun, N.; Zhang, Z.; Bian, J.; Qu, Y.; Li, Z.; Xie, M.; Han, W.; Jing, L. Ultrafine SnO2/010 facet-exposed BiVO4 nanocomposites as efficient photoanodes for controllable conversion of 2,4-dichlorophenol via a preferential dechlorination path. ACS Appl. Mater. Interfaces 2020, 12, 28264–28272. [Google Scholar] [CrossRef]
  25. Chu, X.; Qu, Y.; Zada, A.; Bai, L.; Li, Z.; Yang, F.; Zhao, L.; Zhang, G.; Sun, X.; Yang, Z.D.; et al. Ultrathin Phosphate-Modulated Co Phthalocyanine/g-C3N4 Heterojunction Photocatalysts with Single Co–N4 (II) Sites for Efficient O2 Activation. Adv. Sci. 2020, 7, 2001543. [Google Scholar] [CrossRef]
  26. Zhu, Z.; Lu, Z.; Wang, D.; Tang, X.; Yan, Y.; Shi, W.; Wang, Y.; Gao, N.; Yao, X.; Dong, H. Construction of high-dispersed Ag/Fe3O4/g-C3N4 photocatalyst by selective photo-deposition and improved photocatalytic activity. Appl. Catal. B Environ. 2016, 182, 115–122. [Google Scholar] [CrossRef]
  27. Raziq, F.; He, J.; Gan, J.; Humayun, M.; Faheem, M.B.; Iqbal, A.; Hayat, A.; Fazal, S.; Yi, J.; Zhao, Y.; et al. Promoting visible-light photocatalytic activities for carbon nitride based 0D/2D/2D hybrid system: Beyond the conventional 4-electron mechanism. Appl. Catal. B Environ. 2020, 270, 118870. [Google Scholar] [CrossRef]
  28. Em, S.; Yedigenov, M.; Khamkhash, L.; Atabaev, S.; Molkenova, A.; Poulopoulos, S.G.; Atabaev, T.S. Uncovering the role of surface-attached Ag nanoparticles in photodegradation improvement of rhodamine B by ZnO-Ag nanorods. Nanomaterials 2022, 12, 2882. [Google Scholar] [CrossRef]
  29. Bian, J.; Zhang, Z.; Feng, J.; Thangamuthu, M.; Yang, F.; Sun, L.; Li, Z.; Qu, Y.; Tang, D.; Lin, Z.; et al. Energy Platform for Directed Charge Transfer in the Cascade Z-Scheme Heterojunction: CO2 Photoreduction without a Cocatalyst. Angew. Chemie—Int. Ed. 2021, 60, 20906–20914. [Google Scholar] [CrossRef]
  30. Xiao, Y.; Yao, B.; Wang, Z.; Chen, T.; Xiao, X.; Wang, Y. Plasma Ag-modified α-Fe2O3/g-C3N4 self-assembled S-scheme heterojunctions with enhanced photothermal-photocatalytic-fenton performances. Nanomaterials 2022, 12, 4212. [Google Scholar] [CrossRef]
  31. Khan, I.; Chu, X.; Khan, I.; Liu, H.; Li, W.; Bai, L.; Jing, L. Synthesis of Nanosized Ag-Modified 2D/2D Hydroxylated g-C3N4/TS-1 Z-Scheme Nanocomposites for Efficient Photocatalytic CO2 Reduction. Mater. Res. Bull. 2020, 130, 110926. [Google Scholar] [CrossRef]
  32. Zha, R.; Shi, T.; He, L.; Zhang, M. Synergetic Excitonic and Defective Effects in Confined SnO2/α-Fe2O3 Nanoheterojunctions for Efficient Photocatalytic Molecular Oxygen Activation. Chem. Eng. J. 2021, 421, 129883. [Google Scholar] [CrossRef]
  33. Rej, S.; Hejazi, S.M.H.; Badura, Z.; Zoppellaro, G.; Kalytchuk, S.; Kment, Š.; Fornasiero, P.; Naldoni, A. Light-Induced Defect Formation and Pt Single Atoms Synergistically Boost Photocatalytic H2 Production in 2D TiO2-Bronze Nanosheets. ACS Sustain. Chem. Eng. 2022, 10, 17286–17296. [Google Scholar] [CrossRef]
Figure 1. Photocatalytic aerobic degradation for 2,4-DCP over (a) FO, 5SO/FO, and 2Ag-5SO/FO, respectively; (b) recycle test of 2Ag-5SO/FO for photocatalytic 2,4-DCP degradation.
Figure 1. Photocatalytic aerobic degradation for 2,4-DCP over (a) FO, 5SO/FO, and 2Ag-5SO/FO, respectively; (b) recycle test of 2Ag-5SO/FO for photocatalytic 2,4-DCP degradation.
Catalysts 13 00696 g001
Figure 2. (a) TEM image and (b) XPS Ag 3d spectrum of 2Ag-5SO/FO.
Figure 2. (a) TEM image and (b) XPS Ag 3d spectrum of 2Ag-5SO/FO.
Catalysts 13 00696 g002
Figure 3. (a) FS related to produced hydroxyl radicals under visible-light irradiation and (b) SS-SPS signals for FO, 5SO/FO, and 2Ag-5SO/FO, respectively.
Figure 3. (a) FS related to produced hydroxyl radicals under visible-light irradiation and (b) SS-SPS signals for FO, 5SO/FO, and 2Ag-5SO/FO, respectively.
Catalysts 13 00696 g003
Figure 4. (a) Linear sweep voltammetry (LSV) curves, and electrochemical impedance spectra (EIS) Nyquist plots, related to charging separation of FO, 5SO/FO, and 2Ag-5SO/FO (inset); (b) Electrochemical chemical (EC) reduction curves for O2-bubbled system of FO, 5SO/FO, and 2Ag-5SO/FO.
Figure 4. (a) Linear sweep voltammetry (LSV) curves, and electrochemical impedance spectra (EIS) Nyquist plots, related to charging separation of FO, 5SO/FO, and 2Ag-5SO/FO (inset); (b) Electrochemical chemical (EC) reduction curves for O2-bubbled system of FO, 5SO/FO, and 2Ag-5SO/FO.
Catalysts 13 00696 g004
Figure 5. Applicability of the FO-based photocatalysts (FO, 5SO/FO, and 2Ag-5SO/FO) for the visible-light aerobic degradation of (a) MO and BPA; (b) and CO2 reduction.
Figure 5. Applicability of the FO-based photocatalysts (FO, 5SO/FO, and 2Ag-5SO/FO) for the visible-light aerobic degradation of (a) MO and BPA; (b) and CO2 reduction.
Catalysts 13 00696 g005
Scheme 1. Schematic illustration of the photocatalytic mechanism over Ag-SO/FO photocatalyst for the degradation of organic pollutants under visible-light irradiation.
Scheme 1. Schematic illustration of the photocatalytic mechanism over Ag-SO/FO photocatalyst for the degradation of organic pollutants under visible-light irradiation.
Catalysts 13 00696 sch001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ali, W.; Li, Z.; Bai, L.; Ansar, M.Z.; Zada, A.; Qu, Y.; Shaheen, S.; Jing, L. Controlled Synthesis of Ag-SnO2/α-Fe2O3 Nanocomposites for Improving Visible-Light Catalytic Activities of Pollutant Degradation and CO2 Reduction. Catalysts 2023, 13, 696. https://doi.org/10.3390/catal13040696

AMA Style

Ali W, Li Z, Bai L, Ansar MZ, Zada A, Qu Y, Shaheen S, Jing L. Controlled Synthesis of Ag-SnO2/α-Fe2O3 Nanocomposites for Improving Visible-Light Catalytic Activities of Pollutant Degradation and CO2 Reduction. Catalysts. 2023; 13(4):696. https://doi.org/10.3390/catal13040696

Chicago/Turabian Style

Ali, Wajid, Zhijun Li, Linlu Bai, Muhammad Zaka Ansar, Amir Zada, Yang Qu, Shabana Shaheen, and Liqiang Jing. 2023. "Controlled Synthesis of Ag-SnO2/α-Fe2O3 Nanocomposites for Improving Visible-Light Catalytic Activities of Pollutant Degradation and CO2 Reduction" Catalysts 13, no. 4: 696. https://doi.org/10.3390/catal13040696

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