Next Article in Journal
High-Growth Aspirations of Entrepreneurs in Latin America: Do Alliances Matter?
Next Article in Special Issue
The Influence of Salinity on the Removal of Ni and Zn by Sorption onto Iron Oxide- and Manganese Oxide-Coated Sand
Previous Article in Journal
On the Design of an Integrated System for Wave Energy Conversion Purpose with the Reaction Mass on Board
Previous Article in Special Issue
Water Quality Improvement and Pollutant Removal by Two Regional Detention Facilities with Constructed Wetlands in South Texas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hematite/Graphitic Carbon Nitride Nanofilm for Fenton and Photocatalytic Oxidation of Methylene Blue

Department of Civil and Environmental Engineering, Hanyang University, Seoul 04763, Korea
*
Author to whom correspondence should be addressed.
Sustainability 2020, 12(7), 2866; https://0-doi-org.brum.beds.ac.uk/10.3390/su12072866
Submission received: 18 March 2020 / Revised: 1 April 2020 / Accepted: 2 April 2020 / Published: 3 April 2020
(This article belongs to the Special Issue Green and Sustainable Solutions for the Environment)

Abstract

:
Hematite (α-Fe2O3)/graphitic carbon nitride (g-C3N4) nanofilm catalysts were synthesized on fluorine-doped tin oxide glass by hydrothermal and chemical vapor deposition. Scanning electron microscopy, energy-dispersive spectroscopy, X-ray diffraction, and X-ray photoelectron spectroscopy analyses of the synthesized catalyst showed that the nanoparticles of g-C3N4 were successfully deposited on α-Fe2O3 nanofilm. The methylene blue degradation efficiency of the α-Fe2O3/g-C3N4 composite catalyst was 2.6 times greater than that of the α-Fe2O3 single catalyst under ultraviolet (UV) irradiation. The methylene blue degradation rate by the α-Fe2O3/g-C3N4 catalyst increased by 6.5 times after 1 mM of hydrogen peroxide (H2O2) was added. The photo-Fenton reaction of the catalyst, UV, and H2O2 greatly increased the methylene blue degradation. The results from the scavenger experiment indicated that the main reactants in the methylene blue decomposition reaction are superoxide radicals photocatalytically generated by g-C3N4 and hydroxyl radicals generated by the photo-Fenton reaction. The α-Fe2O3/g-C3N4 nanofilm showed excellent reaction rate constants at pH 3 (Ka = 6.13 × 10−2 min−1), and still better efficiency at pH 7 (Ka = 3.67 × 10−2 min−1), compared to other methylene blue degradation catalysts. As an immobilized photo-Fenton catalyst without iron sludge formation, nanostructured α-Fe2O3/g-C3N4 are advantageous for process design compared to particle-type catalysts.

Graphical Abstract

1. Introduction

Advanced oxidation processes (AOPs) include various methods of decomposing organics through the generation of reactive oxygen species (ROS), such as hydroxyl radicals (•OH) or superoxide radicals (•O2) [1], which are intermediate reactants with strong oxidizing power. AOPs include the Fenton reaction, ozone (O3)/hydrogen peroxide (H2O2), ultraviolet (UV)/H2O2, UV/O3, and photocatalysis. Photocatalysis is promising because of its environmental stability, low processing costs, and wide applications [2]. Photocatalysts are semiconductors that can excite electrons with light energy and form hole–electron pairs. The excited electrons and generated holes are decomposed directly or indirectly by ROS generation through redox reactions. The energy levels in the valence and conduction bands where holes and electrons are present are important factors in photocatalytic activity. They determine the oxidizable organics and the ROS that can be produced depending on the redox potential. Another important factor in the development of photocatalytic processes is the bandgap energy, which is the difference between energy levels of the valence and conduction bands. Photocatalysts are activated by light with energy above the bandgap, and the energy of light is related to the wavelength of light available. Lower bandgaps have a wider spectrum of available light.
The Fenton process is a widely used AOP for water treatment, as it is inexpensive. It can be carried out at room temperature and low atmospheric pressure. In the Fenton reaction, ferric ions react with H2O2to form ROS species that oxidize organic matter as shown in Equation (1).
Fe2+ + H2O2 → Fe3+ + OH + •OH.
One disadvantage of the Fenton process is that H2O2 consumes radicals and wastes oxidizing agents, reducing oxidation efficiency. The other disadvantage is the loss of Fe ions and the formation of solid sludge. The photo-Fenton process, which combines UV with traditional Fenton reactions, is an alternative to the conventional Fenton process that increases oxidation efficiency and reduces sludge. The combination of UV with the Fenton process was reported to significantly reduce iron loss and sludge production [3]. The additional reaction by UV, as shown in Equations (2) and (3), allows the photo-Fenton process to generate more ROS than the simple Fenton process.
Fe(OH)2+ + hv → Fe2+ + 2OH,
H2O2 + hv → 2•OH.
Hematite (α-Fe2O3) can be used as a photocatalyst with a bandgap of 2.2 eV and also can be used as a catalyst for the Fenton reaction by the change of Fe ions on its surface. By combining α-Fe2O3 with H2O2 and UV, photocatalytic reactions can be carried out simultaneously. However, existing hematite-based photocatalysts exhibit high hole–electron recombination due to the narrow bandgap and limited ROS production. This is because of the low redox potential of the conduction band; as a result, it is not practical to apply to water treatment. To overcome these shortcomings, various modification studies have been carried out, testing non-metal doping [4,5], metal doping [6,7,8], shape control [9,10,11], and heterojunction [12,13,14]. A heterojunction with other catalysts accumulates charge and reduces hole–electron recombination by moving excited electrons to the energy levels of other catalysts through the interaction of two semiconductors with different band structures. Graphitic carbon nitride (g-C3N4) has a relatively low bandgap energy and enough conduction band potential for •O2 generation, making it suitable as a junction catalyst for α-Fe2O3 modification.
Nanostructured α-Fe2O3 has been studied to be used as a catalyst for the Fenton reaction. Guo et al. [15] confirmed that S-doped α-Fe2O3 exhibited excellent heterogeneous Fenton efficiency under UV/visible light, while little Fenton reactivity was observed in the dark. Liao et al. [16] reported that phenol and phenolic contaminants were decomposed using an α-Fe2O3 catalyst supported on multiwall carbon nanotubes, showing 20% higher degradation efficiency than bare α-Fe2O3. Chan et al. [17] reported that exposed surface facets play an important role in determining catalytic performance in methylene blue degradation with nanostructured α-Fe2O3. It was confirmed that the photocatalytic performance of three types of α-Fe2O3 nanoparticles was in the order of {113}> {104}> {001}. Liu et al. [18] reported that the increase in sintering temperature of α-Fe2O3 nanoparticles increased in methylene decomposition. They used α-Fe2O3 nanoparticles synthesized by a simple sodium dodecyl sulfate-assisted grinding reaction and subsequent heating treatment process. The α-Fe2O3 particles obtained at 500 °C showed the highest decomposition efficiency. Liu et al. [19] synthesized nanoring-type α-Fe2O3 by adding H2PO4, SO42−, and citric acid in a hydrothermal reaction and reported increased photo-Fenton decomposition of p-nitrophenol by activating photoabsorption and increasing surface-coupled Fe2+ content. Jiang et al. [20] reported that α-Fe2O3/BiOI catalysts showed 3 and 10 times higher decomposition efficiency than BiOI and α-Fe2O3 catalysts in methyl orange, phenol, and tetracycline hydrochloride decomposition. The interaction of the two catalysts increases the separation and transfer efficiency of photogenerated charges and results in an optimized Fenton reaction at the solid–liquid interface. Ren et al. [21] reported that Cu-doping interacted with α-Fe2O3 to have higher Fenton activity than undoped samples in methylene blue degradation. The previous studies show that nanostructured α-Fe2O3 can be used in the Fenton reaction and can increase the activity of the catalyst through shape control and doping. However, there are inherent limitations in the process design of nanoparticle-type catalysts. They require a separate catalyst recovery process after contaminant treatment. This may reduce the merits of decreased sludge generation by using α-Fe2O3, which is advantageous compared to the Fenton reaction using iron salt. The α-Fe2O3 catalyst with nanostructures in an immobilized form rather than in particle form can be easier to use and decrease the costs of process design and operation.
The α-Fe2O3 also is being actively studied as a photocatalyst. However, due to its low photoactivity when used alone, it is often used in heterojunction with other catalysts. The g-C3N4 is frequently used with α-Fe2O3 because of their band structures. It acts as a Z-scheme photocatalyst with the band of α-Fe2O3 and can produce •O2, that cannot be produced by α-Fe2O3. Bai et al. [22] reported that α-Fe2O3 and porous g-C3N4 composites (Ka = 1.28×10−1 min−1) showed 1.43 times better efficiency than pure g-C3N4 (ka = 5.64 × 10−2 min−1) in the degradation of Rhodamine B. Sun et al. [23] found that the α-Fe2O3/g-C3N4 heterojunction photocatalyst (Ka = 4.70 × 10−3 min−1) showed increases of 11.8 and 5.9 times in arsenic removal efficiency relative to pure α-Fe2O3 and g-C3N4, respectively. Hao et al. [24] reported that 55% of methylene blue was removed in a 180 min reaction with g-C3N4 alone, while 75% of methylene blue was removed after 180 min with the α-Fe2O3/g-C3N4 complex photocatalyst. In the charge transfer characterization using electrochemical impedance spectroscopy and photocurrent, α-Fe2O3/g-C3N4 has a smaller charge transfer resistance and higher photocurrent response than g-C3N4. Zhou et al. [25] confirmed that the photo-Fenton reaction using the α-Fe2O3/g-C3N4 catalyst was 45.4 times and 8.4 times more efficient than those of pure α-Fe2O3 and g-C3N4, respectively. It also was 7.2 times more efficient than a mixture of the two catalysts. However, its potential as a Fenton catalyst was seldom utilized, as most studies focused on the use of α-Fe2O3 as a photocatalyst.
In this study, we synthesized a nanostructured α-Fe2O3/g-C3N4 catalyst on a fluorine-doped tin oxide (FTO) glass. The α-Fe2O3/g-C3N4 catalyst simultaneously acts as a photocatalyst and photo-Fenton catalyst. Catalysts immobilized on FTO glass have advantages in process design because they reduce the loss of iron ions and do not require a separate recovery process. A junction with g-C3N4 enables the use of more ROS species while improving the photoelectrochemical properties of the catalyst.

2. Experimental Section

2.1. Materials Synthesis

Nanostructured α-Fe2O3 films were synthesized on FTO glass simply by a modified hydrothermal method [26]. As a precursor for synthesizing α-Fe2O3 nanofilms, 10 mM FeCl3•6H2O and NaNO3 were dissolved in 100 mL of 0.05 M HCl solution. The solution was put in a 150 mL Teflon liner. A piece of FTO glass ultrasonicated in acetone, ethanol, and distilled water was placed on the Teflon liner. Next, the Teflon liner was placed in a stainless-steel autoclave and heated at 95 °C for 4 h in an electric furnace. Akaganeite(β-FeOOH) films were successfully synthesized on the FTO glass. The obtained β-FeOOH films were washed with distilled water and sintered in an electric furnace at 520 °C for 2 h, resulting in the synthesis of red-colored α-Fe2O3 films.
The g-C3N4 was synthesized on α-Fe2O3 nanofilms by directly heating melamine [27]. First, 5 g of melamine was placed in a ceramic crucible with a cover. Then, the β-FeOOH films were placed on the melamine with β-FeOOH facing down. The crucible was partially covered and heated at 520 °C for 2 h in an electric furnace. The sample was then cooled to room temperature in the atmosphere. After the reaction, brick-colored α-Fe2O3/g-C3N4 films were obtained with yellow g-C3N4 powders on the bottom of the crucible. It was confirmed that 280 mg of catalyst was synthesized on one FTO substrate by comparing the weight before and after synthesis.

2.2. Materials Characterization

The morphology and nanostructures of the samples were analyzed by field-emission scanning electron microscopy (FE-SEM; JSM-6701F and FEI Nova NanoSEM 450, JEOL, Japan). Energy-dispersive spectroscopy (EDS) was used for surface elemental analysis using the FE-SEM. To analyze the crystal phase, X-ray diffraction (XRD; SmartLab, Rigaku, Japan) analysis was carried out with Cu kα radiation. XRD data were obtained at conditions of 45 kV, 200 mA, from 2θ = 20° to 80°, with a step size of 0.04° at 4°/min. X-ray photoelectron spectroscopy (XPS; K-alpha plus, Thermo Scientific, Massachusetts, US) was taken with Al kα (1486.6 eV).

2.3. Aqueous Organic Oxidation

The organic oxidation efficiency of the sample was confirmed by methylene blue decomposition. The sample was placed in a 250 mL quartz beaker and 200 mL of 10 mg/L methylene blue solution was added. The beaker was placed in a reactor with three UV lamps (254 nm, 6 W) and irradiated with light for the reaction for 90 min. The light intensity applied to the center of the catalyst was 2.88 mW/cm2. It was measured using a lux meter (UVC-254SD, Lutron Electronics, Pennsylvania, US). During the reaction, the solution was stirred with an electric stirrer. Experiments were also carried out under the same conditions with addition of various concentrations of H2O2 (0.1, 0.2, 0.5, 1, and 2 mM) and pH (3, 5, 7, 9, and 11). Scavenger experiments were performed to measure the main reactants of the methylene blue degradation. The methylene blue degradations after addition of 1 mM benzoquinone (BQ) or tert-butyl alcohol (TBA) were compared after 90 min of reactions at the initial pH 7. A total of 4 mL of the solution was collected at regular intervals and filtered using a syringe filter (pore size: 0.45 µm, Whatman, UK). The filtered samples were analyzed by examining the absorbance of 665 nm light by a UV-vis spectrometer (DR 2800, HACH, Colorado, US) to determine the methylene blue degradation rate. All experimental data were duplicated and averaged.

3. Results and Discussion

3.1. Structural Analysis

Figure 1 shows the morphology and nanostructures of (a) β-FeOOH and (b) α-Fe2O3. The FeOOH has nanorods arranged like flowers. The nanorods are about 50 nm wide and angled. The nanostructure of α-Fe2O3 changed to a more rounded form after sintering, and both the width and length were smaller than those of the β-FeOOH nanorods. This change is due to the escape of H2O when β-FeOOH is sintered and α-Fe2O3 is formed, as shown in Equation (4) [28].
2FeOOH → Fe2O3 + H2O
Figure 2 shows the FE-SEM image and EDS analysis of the α-Fe2O3/g-C3N4 films. The particles of g-C3N4 were deposited on the α-Fe2O3 nanostructures. The g-C3N4 particles are less than 35 nm in size, and the shape of Fe2O3 is rounder due to further sintering. The C, N, Fe, and O peaks were identified in EDS analysis. The peaks of elements other than O were measured slightly lower because the analysis depth of EDS is greater than the thickness of α-Fe2O3/g-C3N4 films. As a result, the elements of the FTO under the catalyst are also measured. The peaks between 3 and 4 keV correspond to Sn, which is a component of FTO.
Figure 3 shows the XRD patterns of β-FeOOH, α-Fe2O3, and α-Fe2O3/g-C3N4 samples and bare FTO. The peak of FeOOH was observed on the FTO substrate after the hydrothermal synthesis. After sintering, FeOOH was successfully converted to α-Fe2O3, and the eight characteristic peaks of α-Fe2O3 ({012}, {104}, {110}, {113}, {024}, {116}, {214}, and {300}) were observed [29]. The peaks of the FTO glass were also observed in the samples due to the thinness of the α-Fe2O3 film layer. No characteristic peaks of g-C3N4 were observed in the α-Fe2O3/g-C3N4 samples, and only the peaks of α-Fe2O3 and FTO were identified. This may be because the amount of g-C3N4 deposited on the surface is relatively small, and it is dispersed well on the surface without affecting the structure of α-Fe2O3.
XPS analysis was performed to confirm the synthesis of g-C3N4. Figure 4 shows (a) the XPS spectrum of α-Fe2O3/g-C3N4 and the (b) C 1s and (c) N 1s regions. Four elements (Fe, O, C, and N) were detected in the XPS spectra. The most prominent peak in the C 1s region is attributed to the carbon–nitrogen bond structure (N-C = N) of the aromatic systems. The peak at 398.3 eV seen in the N 1s region is the peak attributed to graphene N and generally indicates the case where the nitrogen atom is incorporated into the graphene layer to replace the carbon atom. The peak at 401.0 eV is attributed to nitrogen spiked with three sp2 carbon atoms [30,31]. The peaks in the C 1s and N 1s regions show that the melamine produced g-C3N4 with a graphitic structure through heating and was successfully deposited on the α-Fe2O3 surface.

3.2. Methylene Blue Degradation

Methylene blue degradation results are shown in Figure 5. No significant changes were observed when the experiment was performed using bare FTO glass. Only 12% of methylene blue was degraded with pure α-Fe2O3 after the 90 min reaction even under UV irradiation, whereas 31% of methylene blue was degraded with the α-Fe2O3/g-C3N4 heterojunction catalyst. This is because the band structure of α-Fe2O3, which cannot generate •O2, is not suitable for methylene blue decomposition, whereas the conduction band of g-C3N4 has enough redox potential for •O2 production. In the decomposition of photocatalytic organic matter of α-Fe2O3/g-C3N4, g-C3N4 can be said to be the main catalyst, as opposed to α-Fe2O3.
After addition of 1 mM H2O2, the methylene blue degradation rate increased significantly in all samples. The increase in methylene blue degradation rate in bare FTO samples was due to the UV/H2O2 reaction, and the increases in degradation efficiency due to H2O2 addition in α-Fe2O3 and α-Fe2O3/g-C3N4 samples were greater than that by the UV/H2O2 reaction. This means that there is an interaction between the α-Fe2O3 and α-Fe2O3/g-C3N4 catalyst and H2O2 in addition to the UV/H2O2 reaction. The methylene blue degradation rate of the α-Fe2O3 catalyst under UV irradiation increased from 12% to 77% with H2O2 addition after 90 min of reaction. Methylene blue degradation of α-Fe2O3/g-C3N4 showed 31% degradation efficiency after 90 min of reaction under UV irradiation, but after H2O2 addition, the complete decomposition of methylene blue was achieved in only 75 min of reaction. This is due to the Fenton reaction, in which the Fe3+ on the surface of α-Fe2O3 becomes Fe2+ due to the excited α-Fe2O3 electrons under UV light. They, in turn, reacted with H2O2 to form •OH [17,32].
The change of methylene blue decomposition efficiency caused by the concentration of H2O2 added is shown in Figure 6. For H2O2 concentrations of up to 1 mM, the higher the concentration of the H2O2 added, the higher the decomposition efficiency will be. However, no noticeable change in efficiency was seen at 1 and 2 mM. This may be due to the saturation of the number of surface Fe ions reacting with H2O2 at 1 mM H2O2. The effect of the initial pH is presented in Figure 7. At low initial pH concentrations, the methylene blue decomposition efficiency was increased. This optimal pH is similar to those of the Fenton process. This shows that the α-Fe2O3/g-C3N4 catalyst performs both Fenton and photocatalytic oxidation simultaneously.

3.3. Detection of Reactive Species

ROS scavengers were added to the methylene blue oxidation to identify the main reactants. BQ was introduced as an •O2 scavenger, and TBA was used as an •OH scavenger [33]. As shown in Figure 8, the effect of the •OH scavenger was low without H2O2 addition. On the other hand, when the •O2 scavenger was added, the reaction was significantly inhibited. This means that, in the absence of H2O2, the main reactant of methylene blue oxidation is the •O2 produced in the conduction band of g-C3N4. However, when H2O2 assists in oxidation, the degradation rate was significantly lowered when •OH scavenger was added. This shows that the •OH generated by the Fenton reaction is one of the major reactants in the methylene blue decomposition reaction.

3.4. Degradation Kinetics

The methylene blue degradation during the first 45 min of this study can be explained by the pseudo-first-order reaction kinetics of Equation (5).
ln (C0/C) = Ka × t + constant,
where C0 is the initial concentration of methylene blue, C is the methylene blue concentration at time t, and Kapp is the first-order reaction constant. Ka was chosen as a kinetic parameter to compare the catalytic activity of different systems independently of concentration. Kinetic parameters for each methylene blue degradation process are listed in Table 1 and Table 2. All results followed a linear regression and Ka effectively expressed the methylene blue decomposition efficiency. The reaction rate of α-Fe2O3/g-C3N4 nanofilm was 2.4 times higher than that of theα-Fe2O3 film under conditions of pH 7 and 1 mM H2O2 concentration, and the methylene blue decomposition rate of α-Fe2O3/g-C3N4 nanofilms at the initial pH 3 increased by 67% over that at pH 7. Previous studies on methylene blue degradation are listed in Table 3, demonstrating that α-Fe2O3/g-C3N4 nanofilms are excellent even in neutral conditions.

3.5. Methylene Blue Degradation Mechanism

The degradation mechanism of the catalyst based on these results is shown in Figure 9. Traditional dual-carrier delivery systems, shown in Figure 9a, cannot account for the high activity of •O2 in reactions without H2O2. When the redox potential for generating the •O2 is −0.046 (eV vs. normal hydrogen electrode) and the electrons move to the conduction band of α-Fe2O3, the photocatalytic activity is somewhat lower [43]. The holes and electrons accumulated due to the Z-scheme process of α-Fe2O3/g-C3N4 can be used for photocatalytic oxidation, as shown in Figure 9b. The Z-scheme action of the α-Fe2O3/g-C3N4 heterojunction catalyst in the field of electrolysis and CO2 reduction has been reported [44,45,46]. However, the activity of low •OH in the absence of H2O2 cannot be explained with the Z-scheme system. If the •OH is produced in the valence band of α-Fe2O3, there should be enough •OH activity even without H2O2. Thus, the combination of photocatalytic and Fenton processes is presented in Figure 9c. The high activity of •OH in the presence of H2O2 means that •OH involved in photodegradation is generated in H2O2. The electrons excited by UV can change Fe3+ to Fe2+ on the α-Fe2O3 surface, and Fe2+ reacts with H2O2 to generate •OH [32,47]. The difference in the energy levels of g-C3N4 and α-Fe2O3 induces charge transfer. The junctions of two semiconductors with different band structures can accumulate charge by shifting the photoexcited electrons to the energy levels of the other catalysts. The coupled g-C3N4 valence band acts as an electron trap for electrons excited by the conduction band of α-Fe2O3. Catalysts composed of highly efficient nanostructures have low electron resistance, and the heterojunction structures of both catalysts make charge separation more efficient and reduce hole–electron recombination caused by instability of excited electrons.

4. Conclusions

Nano-structured α-Fe2O3 was successfully synthesized on FTO glass. Nanoparticles of g-C3N4 were successfully deposited on the synthesized α-Fe2O3 films. In the methylene blue oxidation using α-Fe2O3/g-C3N4 nanofilm, the activity of •OH under UV irradiation was low when H2O2 was not present. XPS analysis showed changes in Fe ions on the α-Fe2O3 surface by H2O2, which is similar to the change in iron ions in the Fenton reaction. The Fenton reaction of iron ions on the surface of α-Fe2O3 activated by H2O2 and the photocatalytic activity of g-C3N4 nitride enhanced through heterojunction with α-Fe2O3 are suggested as the main mechanisms of α-Fe2O3/g-C3N4 nanofilms. The α-Fe2O3/g-C3N4 nanofilms as immobilized Fenton-photocatalytic composite catalysts are advantageous for reactor design as an alternative to conventional treatment with iron salts.

Author Contributions

Conceptualization, S.L.; Data curation, S.L.; Formal analysis, S.L.; Methodology, S.L. and J.-W.P.; Supervision, J.-W.P.; Visualization, S.L.; Writing—original draft, S.L.; Writing—review & editing, J.-W.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Korea Environmental Industry and Technology Institute (KEITI) through The Chemical Accident Prevention Technology Development Project, funded by Korea Ministry of Environment (MOE) (2019001960005).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pera-Titus, M.; García-Molina, V.; Baños, M.A.; Giménez, J.; Esplugas, S. Degradation of chlorophenols by means of advanced oxidation processes: A general review. Appl. Catal. B Environ. 2004, 47, 219–256. [Google Scholar] [CrossRef]
  2. Chong, M.N.; Jin, B.; Chow, C.W.K.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef] [PubMed]
  3. Rahim Pouran, S.; Abdul Aziz, A.R.; Wan Daud, W.M.A. Review on the main advances in photo-Fenton oxidation system for recalcitrant wastewaters. J. Ind. Eng. Chem. 2015, 21, 53–69. [Google Scholar] [CrossRef]
  4. Pradhan, G.K.; Sahu, N.; Parida, K.M. Fabrication of S, N co-doped α-Fe2O3 nanostructures: Effect of doping, OH radical formation, surface area, [110] plane and particle size on the photocatalytic activity. RSC Adv. 2013, 3, 7912–7920. [Google Scholar] [CrossRef]
  5. He, L.; Jing, L.; Luan, Y.; Wang, L.; Fu, H. Enhanced Visible Activities of α-Fe2O3 by Coupling N-Doped Graphene and Mechanism Insight. ACS Catal. 2014, 4, 990–998. [Google Scholar] [CrossRef]
  6. Li, F.B.; Li, X.Z.; Liu, C.S.; Liu, T.X. Effect of alumina on photocatalytic activity of iron oxides for bisphenol A degradation. J. Hazard. Mater. 2007, 149, 199–207. [Google Scholar] [CrossRef] [Green Version]
  7. Chen, L.; Li, F.; Ni, B.; Xu, J.; Fu, Z.; Lu, Y. Enhanced visible photocatalytic activity of hybrid Pt/α-Fe2O3 nanorods. RSC Adv. 2012, 2, 10057–10063. [Google Scholar] [CrossRef]
  8. Cao, S.-W.; Fang, J.; Shahjamali, M.M.; Wang, Z.; Yin, Z.; Yang, Y.; Boey, F.Y.C.; Barber, J.; Loo, S.C.J.; Xue, C. In situ growth of Au nanoparticles on Fe2O3 nanocrystals for catalytic applications. CrystEngComm 2012, 14, 7229–7235. [Google Scholar] [CrossRef]
  9. Kawahara, T.; Yamada, K.; Tada, H. Visible light photocatalytic decomposition of 2-naphthol by anodic-biased α-Fe2O3 film. J. Colloid Interface Sci. 2006, 294, 504–507. [Google Scholar] [CrossRef]
  10. Tamboli, S.H.; Rahman, G.; Joo, O.-S. Influence of potential, deposition time and annealing temperature on photoelectrochemical properties of electrodeposited iron oxide thin films. J. Alloy. Compd. 2012, 520, 232–237. [Google Scholar] [CrossRef]
  11. Zhang, Z.; Hossain, M.F.; Takahashi, T. Self-assembled hematite (α-Fe2O3) nanotube arrays for photoelectrocatalytic degradation of azo dye under simulated solar light irradiation. Appl. Catal. B Environ. 2010, 95, 423–429. [Google Scholar] [CrossRef]
  12. Jaramillo-Páez, C.; Navío, J.A.; Hidalgo, M.C.; Bouziani, A.; Azzouzi, M.E. Mixed α-Fe2O3/Bi2WO6 oxides for photoassisted hetero-Fenton degradation of Methyl Orange and Phenol. J. Photochem. Photobiol. A Chem. 2017, 332, 521–533. [Google Scholar] [CrossRef]
  13. Sarkar, D.; Singh, A.K. Mechanism of Nonvolatile Resistive Switching in ZnO/α-Fe2O3 Core–Shell Heterojunction Nanorod Arrays. J. Phys. Chem. C 2017, 121, 12953–12958. [Google Scholar] [CrossRef]
  14. Kang, J.; Kuang, Q.; Xie, Z.-X.; Zheng, L.-S. Fabrication of the SnO2/α-Fe2O3 Hierarchical Heterostructure and Its Enhanced Photocatalytic Property. J. Phys. Chem. C 2011, 115, 7874–7879. [Google Scholar] [CrossRef]
  15. Guo, L.; Chen, F.; Fan, X.; Cai, W.; Zhang, J. S-doped α-Fe2O3 as a highly active heterogeneous Fenton-like catalyst towards the degradation of acid orange 7 and phenol. Appl. Catal. B Environ. 2010, 96, 162–168. [Google Scholar] [CrossRef]
  16. Liao, Q.; Sun, J.; Gao, L. Degradation of phenol by heterogeneous Fenton reaction using multi-walled carbon nanotube supported Fe2O3 catalysts. Colloids Surf. A Physicochem. Eng. Asp. 2009, 345, 95–100. [Google Scholar] [CrossRef]
  17. Chan, J.Y.T.; Ang, S.Y.; Ye, E.Y.; Sullivan, M.; Zhang, J.; Lin, M. Heterogeneous photo-Fenton reaction on hematite (α-Fe2O3){104}, {113} and {001} surface facets. Phys. Chem. Chem. Phys. 2015, 17, 25333–25341. [Google Scholar] [CrossRef]
  18. Liu, J.; Wang, B.; Li, Z.; Wu, Z.; Zhu, K.; Zhuang, J.; Xi, Q.; Hou, Y.; Chen, J.; Cong, M.; et al. Photo-Fenton reaction and H2O2 enhanced photocatalytic activity of α-Fe2O3 nanoparticles obtained by a simple decomposition route. J. Alloy. Compd. 2019, 771, 398–405. [Google Scholar] [CrossRef]
  19. Liu, H.; Tong, M.; Zhu, K.; Liu, H.; Chen, R. Preparation and photo-fenton degradation activity of α-Fe2O3 nanorings obtained by adding H2PO4, SO42−, and citric acid. Chem. Eng. J. 2020, 382, 123010. [Google Scholar] [CrossRef]
  20. Jiang, J.; Gao, J.; Li, T.; Chen, Y.; Wu, Q.; Xie, T.; Lin, Y.; Dong, S. Visible-light-driven photo-Fenton reaction with α-Fe2O3/BiOI at near neutral pH: Boosted photogenerated charge separation, optimum operating parameters and mechanism insight. J. Colloid Interface Sci. 2019, 554, 531–543. [Google Scholar] [CrossRef]
  21. Ren, B.; Miao, J.; Xu, Y.; Zhai, Z.; Dong, X.; Wang, S.; Zhang, L.; Liu, Z. A grape-like N-doped carbon/CuO-Fe2O3 nanocomposite as a highly active heterogeneous Fenton-like catalyst in methylene blue degradation. J. Clean. Prod. 2019, 240, 118143. [Google Scholar] [CrossRef]
  22. Bai, J.; Xu, H.; Chen, G.; Lv, W.; Ni, Z.; Wang, Z.; Yang, J.; Qin, H.; Zheng, Z.; Li, X. Facile fabrication of α-Fe2O3/porous g-C3N4 heterojunction hybrids with enhanced visible-light photocatalytic activity. Mater. Chem. Phys. 2019, 234, 75–80. [Google Scholar] [CrossRef]
  23. Sun, S.; Ji, C.; Wu, L.; Chi, S.; Qu, R.; Li, Y.; Lu, Y.; Sun, C.; Xue, Z. Facile one-pot construction of α-Fe2O3/g-C3N4 heterojunction for arsenic removal by synchronous visible light catalysis oxidation and adsorption. Mater. Chem. Phys. 2017, 194, 1–8. [Google Scholar] [CrossRef]
  24. Hao, Q.; Mo, Z.; Chen, Z.; She, X.; Xu, Y.; Song, Y.; Ji, H.; Wu, X.; Yuan, S.; Xu, H.; et al. 0D/2D Fe2O3 quantum dots/g-C3N4 for enhanced visible-light-driven photocatalysis. Colloids Surf. A Physicochem. Eng. Asp. 2018, 541, 188–194. [Google Scholar] [CrossRef]
  25. Zhou, L.; Wang, L.; Zhang, J.; Lei, J.; Liu, Y. Well-Dispersed Fe2O3 Nanoparticles on g-C3N4 for Efficient and Stable Photo-Fenton Photocatalysis under Visible-Light Irradiation. Eur. J. Inorg. Chem. 2016, 2016, 5387–5392. [Google Scholar] [CrossRef]
  26. Vayssieres, L.; Beermann, N.; Lindquist, S.-E.; Hagfeldt, A. Controlled Aqueous Chemical Growth of Oriented Three-Dimensional Crystalline Nanorod Arrays: Application to Iron(III) Oxides. Chem. Mater. 2001, 13, 233–235. [Google Scholar] [CrossRef]
  27. Yan, S.C.; Li, Z.S.; Zou, Z.G. Photodegradation Performance of g-C3N4 Fabricated by Directly Heating Melamine. Langmuir 2009, 25, 10397–10401. [Google Scholar] [CrossRef]
  28. Liu, Q.; Chen, C.; Yuan, G.; Huang, X.; Lü, X.; Cao, Y.; Li, Y.; Hu, A.; Lu, X.; Zhu, P. Morphology-controlled α-Fe2O3 nanostructures on FTO substrates for photoelectrochemical water oxidation. J. Alloy. Compd. 2017, 715, 230–236. [Google Scholar] [CrossRef]
  29. Raja, K.; Mary Jaculine, M.; Jose, M.; Verma, S.; Prince, A.A.M.; Ilangovan, K.; Sethusankar, K.; Jerome Das, S. Sol–gel synthesis and characterization of α-Fe2O3 nanoparticles. Superlattices Microstruct. 2015, 86, 306–312. [Google Scholar] [CrossRef]
  30. Thomas, A.; Fischer, A.; Goettmann, F.; Antonietti, M.; Müller, J.-O.; Schlögl, R.; Carlsson, J.M. Graphitic carbon nitride materials: Variation of structure and morphology and their use as metal-free catalysts. J. Mater. Chem. 2008, 18, 4893–4908. [Google Scholar] [CrossRef] [Green Version]
  31. Qiao, F.; Wang, J.; Ai, S.; Li, L. As a new peroxidase mimetics: The synthesis of selenium doped graphitic carbon nitride nanosheets and applications on colorimetric detection of H2O2 and xanthine. Sens. Actuators B Chem. 2015, 216, 418–427. [Google Scholar] [CrossRef]
  32. Su, S.; Liu, Y.; Liu, X.; Jin, W.; Zhao, Y. Transformation pathway and degradation mechanism of methylene blue through β-FeOOH@GO catalyzed photo-Fenton-like system. Chemosphere 2019, 218, 83–92. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, W.; Li, Y.; Liu, F.; Jiang, W.; Zhang, D.; Liang, J. Visible-light-driven photocatalytic degradation of diclofenac by carbon quantum dots modified porous g-C3N4: Mechanisms, degradation pathway and DFT calculation. Water Res. 2019, 151, 8–19. [Google Scholar] [CrossRef] [PubMed]
  34. Abou-Gamra, Z.M.; Ahmed, M.A. Synthesis of mesoporous TiO2–curcumin nanoparticles for photocatalytic degradation of methylene blue dye. J. Photochem. Photobiol. B Biol. 2016, 160, 134–141. [Google Scholar] [CrossRef]
  35. Chauhan, R.; Kumar, A.; Pal Chaudhary, R. Photocatalytic degradation of methylene blue with Cu doped ZnS nanoparticles. J. Lumin. 2014, 145, 6–12. [Google Scholar] [CrossRef]
  36. Jing, H.-P.; Wang, C.-C.; Zhang, Y.-W.; Wang, P.; Li, R. Photocatalytic degradation of methylene blue in ZIF-8. RSC Adv. 2014, 4, 54454–54462. [Google Scholar] [CrossRef]
  37. Atout, H.; Álvarez, M.G.; Chebli, D.; Bouguettoucha, A.; Tichit, D.; Llorca, J.; Medina, F. Enhanced photocatalytic degradation of methylene blue: Preparation of TiO2/reduced graphene oxide nanocomposites by direct sol-gel and hydrothermal methods. Mater. Res. Bull. 2017, 95, 578–587. [Google Scholar] [CrossRef] [Green Version]
  38. Gupta, V.K.; Saleh, T.A.; Pathania, D.; Rathore, B.S.; Sharma, G. A cellulose acetate based nanocomposite for photocatalytic degradation of methylene blue dye under solar light. Ionics 2015, 21, 1787–1793. [Google Scholar] [CrossRef]
  39. Yang, Y.; Xu, L.; Wang, H.; Wang, W.; Zhang, L. TiO2/graphene porous composite and its photocatalytic degradation of methylene blue. Mater. Des. 2016, 108, 632–639. [Google Scholar] [CrossRef]
  40. Wang, Z.; Luo, C.; Zhang, Y.; Gong, Y.; Wu, J.; Fu, Q.; Pan, C. Construction of hierarchical TiO2 nanorod array/graphene/ZnO nanocomposites for high-performance photocatalysis. J. Mater. Sci. 2018, 53, 15376–15389. [Google Scholar] [CrossRef]
  41. Niu, X.; Yan, W.; Shao, C.; Zhao, H.; Yang, J. Hydrothermal synthesis of Mo-C co-doped TiO2 and coupled with fluorine-doped tin oxide (FTO) for high-efficiency photodegradation of methylene blue and tetracycline: Effect of donor-acceptor passivated co-doping. Appl. Surf. Sci. 2019, 466, 882–892. [Google Scholar] [CrossRef]
  42. Wu, F.; Li, X.; Liu, W.; Zhang, S. Highly enhanced photocatalytic degradation of methylene blue over the indirect all-solid-state Z-scheme g-C3N4-RGO-TiO2 nanoheterojunctions. Appl. Surf. Sci. 2017, 405, 60–70. [Google Scholar] [CrossRef]
  43. Ma, D.; Wu, J.; Gao, M.; Xin, Y.; Sun, Y.; Ma, T. Hydrothermal synthesis of an artificial Z-scheme visible light photocatalytic system using reduced graphene oxide as the electron mediator. Chem. Eng. J. 2017, 313, 1567–1576. [Google Scholar] [CrossRef]
  44. 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]
  45. Xu, Q.; Zhu, B.; Jiang, C.; Cheng, B.; Yu, J. Constructing 2D/2D Fe2O3/g-C3N4 Direct Z-Scheme Photocatalysts with Enhanced H2 Generation Performance. Sol. RrL 2018, 2, 1800006. [Google Scholar] [CrossRef]
  46. 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]
  47. Shaban, M.; Abukhadra, M.R.; Ibrahim, S.S.; Shahien, M.G. Photocatalytic degradation and photo-Fenton oxidation of Congo red dye pollutants in water using natural chromite—Response surface optimization. Appl. Water Sci. 2017, 7, 4743–4756. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Scanning electron microscopy (SEM) image of nanostructure of (a) β-FeOOH and (b) α-Fe2O3.
Figure 1. Scanning electron microscopy (SEM) image of nanostructure of (a) β-FeOOH and (b) α-Fe2O3.
Sustainability 12 02866 g001
Figure 2. (a) SEM image and (b) energy-dispersive spectroscopy (EDS) spectrum of α-Fe2O3/g-C3N4 nanofilm.
Figure 2. (a) SEM image and (b) energy-dispersive spectroscopy (EDS) spectrum of α-Fe2O3/g-C3N4 nanofilm.
Sustainability 12 02866 g002
Figure 3. X-ray diffraction (XRD) pattern of bare fluorine-doped tin oxide (FTO), β-FeOOH, α-Fe2O3, and α-Fe2O3/g-C3N4.
Figure 3. X-ray diffraction (XRD) pattern of bare fluorine-doped tin oxide (FTO), β-FeOOH, α-Fe2O3, and α-Fe2O3/g-C3N4.
Sustainability 12 02866 g003
Figure 4. X-ray photoelectron spectroscopy (XPS) spectra of (a) whole surface, and (b) C1s and (c) N1s regions of the α-Fe2O3/g-C3N4.
Figure 4. X-ray photoelectron spectroscopy (XPS) spectra of (a) whole surface, and (b) C1s and (c) N1s regions of the α-Fe2O3/g-C3N4.
Sustainability 12 02866 g004aSustainability 12 02866 g004b
Figure 5. The degradation of methylene blue in different catalytic systems.
Figure 5. The degradation of methylene blue in different catalytic systems.
Sustainability 12 02866 g005
Figure 6. Methylene blue degradation of α-Fe2O/g-C3N4 nanofilm with addition of different concentrations of H2O2.
Figure 6. Methylene blue degradation of α-Fe2O/g-C3N4 nanofilm with addition of different concentrations of H2O2.
Sustainability 12 02866 g006
Figure 7. Methylene blue decomposition of α-Fe2O3/g-C3N4 nanofilm with different initial pH.
Figure 7. Methylene blue decomposition of α-Fe2O3/g-C3N4 nanofilm with different initial pH.
Sustainability 12 02866 g007
Figure 8. Methylene blue photodegradation efficiency of α-Fe2O3/g-C3N4 nanofilm with different radical scavengers.
Figure 8. Methylene blue photodegradation efficiency of α-Fe2O3/g-C3N4 nanofilm with different radical scavengers.
Sustainability 12 02866 g008
Figure 9. Schematic diagrams of charge transfer between g-C3N4 and α-Fe2O3 with the radical generation processes: (a) double transfer, (b) Z-scheme, and (c) combination of photo-Fenton and photocatalysis.
Figure 9. Schematic diagrams of charge transfer between g-C3N4 and α-Fe2O3 with the radical generation processes: (a) double transfer, (b) Z-scheme, and (c) combination of photo-Fenton and photocatalysis.
Sustainability 12 02866 g009
Table 1. Kinetic parameters of different methylene blue decomposition systems in this study.
Table 1. Kinetic parameters of different methylene blue decomposition systems in this study.
Type of Oxidation ProcessKa (min−1)R2
FTO + UV1.40 × 10−40.99
α-Fe2O3 + UV1.80 × 10−30.99
α-Fe2O3/g-C3N4 + UV5.65 × 10−30.99
FTO + UV + H2O24.11 × 10−30.99
α-Fe2O3 + UV + H2O21.52 × 10−20.99
α-Fe2O3/g-C3N4 + UV + H2O23.67 × 10−20.99
Table 2. Kinetic parameters of the methylene blue degradation of α-Fe2O3/g-C3N4 nanofilm according to concentration of added H2O2 and initial pH.
Table 2. Kinetic parameters of the methylene blue degradation of α-Fe2O3/g-C3N4 nanofilm according to concentration of added H2O2 and initial pH.
H2O2 (mM)pHKa (min−1)R2
0.171.10 × 10−20.99
0.271.50 × 10−20.99
0.572.73 × 10−20.99
173.67 × 10−20.99
273.72 × 10−20.99
136.13 × 10−20.99
155.15 × 10−20.99
191.07 × 10−20.99
1114.70 × 10−30.99
Table 3. Comparison of Kapp values of previous studies of methylene blue degradation.
Table 3. Comparison of Kapp values of previous studies of methylene blue degradation.
CatalystKa (/min)Light SourceReference
TiO2–curcumin3.05 × 10−2UV[34]
Cu doped ZnS1.45 × 10−2UV[35]
ZIF-81.70 × 10−2UV[36]
TiO2/rGO nanocomposites1.12 × 10−2UV[37]
CA/TPNC1.26 × 10−2Solar[38]
TiO2/graphene2.08 × 10−2Solar[39]
TiO2 NRAs/graphene/ZnO NPs1.65 × 10−2Visible[40]
(Mo,C)-TiO2/FTO2.95 × 10−2Visible[41]
g-C3N4-10RGO-TiO21.47 × 10−2Visible[42]
This study (1mM H2O2, pH 7)3.67 × 10−2UV-
This study (1mM H2O2, pH 3)6.13 × 10−2UV-

Share and Cite

MDPI and ACS Style

Lee, S.; Park, J.-W. Hematite/Graphitic Carbon Nitride Nanofilm for Fenton and Photocatalytic Oxidation of Methylene Blue. Sustainability 2020, 12, 2866. https://0-doi-org.brum.beds.ac.uk/10.3390/su12072866

AMA Style

Lee S, Park J-W. Hematite/Graphitic Carbon Nitride Nanofilm for Fenton and Photocatalytic Oxidation of Methylene Blue. Sustainability. 2020; 12(7):2866. https://0-doi-org.brum.beds.ac.uk/10.3390/su12072866

Chicago/Turabian Style

Lee, Sangbin, and Jae-Woo Park. 2020. "Hematite/Graphitic Carbon Nitride Nanofilm for Fenton and Photocatalytic Oxidation of Methylene Blue" Sustainability 12, no. 7: 2866. https://0-doi-org.brum.beds.ac.uk/10.3390/su12072866

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