Thi Viet Ha
Luu
*a,
Van Cuong
Nguyen
a,
Thi Dieu
Thuy Tran
a,
Van Dat
Doan
a,
Thi Lieu
Nguyen
a,
Nguyen Xuan
Dung
b and
Huu Phuc
Dang
*c
aFaculty of Chemical Engineering, Industrial University of Ho Chi Minh City, No. 12 Nguyen Van Bao, Ward 1, Go Vap District, Ho Chi Minh City, 700000, Vietnam. E-mail: luuthivietha@iuh.edu.vn
bVinh University, 182 Le Duan, Vinh City, Nghe An 460000, Vietnam
cFaculty of Fundamental Science, Industrial University of Ho Chi Minh City, No. 12 Nguyen Van Bao, Ward 1, Go Vap District, Ho Chi Minh City, 700000, Vietnam. E-mail: danghuuphuc@iuh.edu.vn
First published on 28th February 2025
A novel n–p Fe2O3@ZnBi2O4 (FZB) heterojunction with a unique 3D structure was fabricated in two simple steps to break down MB under visible light. First, the polymer gel combustion technique was employed to fabricate a 3D Fe2O3 framework. Next, a microwave-assisted precipitation approach was used to incorporate 3D ZnBi2O4 flakes onto the framework surface. FZB can effectively collect a broad range of UV-Vis light and sunlight. Surprisingly, the core–shell p–n heterojunction structure makes it easier for photogenerated charges to move and separate. This is because the semiconductor parts are better connected and the electric field is inside the two junctions. UV-Vis-DRS, EIS, PL, and XPS analyses confirmed this phenomenon. As a result, n–p FZB, with a Fe2O3/ZnBi2O4 molar ratio of 2:
1, showed the highest photocatalytic activity, increasing the reaction rate by 4.2 times compared to Fe2O3 and 2.8 times compared to ZnBi2O4. Exposure to light for 100 min at a concentration of 0.7 g L−1 the FZB catalyst and pH = 9 led to the breakdown of more than 95% of the 50 ppm MB solution. In addition, the photodegradation of MB by n–p FZB increased the reaction rate by 2.15 times by adding hydrogen peroxide (H2O2), which is known as the photo-Fenton reaction. The efficacy showed remarkable photocatalytic activity, which increased the reaction rate by 5.95 times when persulfate was used. Finally, after examination of the energy band structure of the materials and the findings regarding the function of the oxidizing sites in the photocatalytic process, a reaction mechanism was proposed.
Advanced oxidation uses highly reactive oxidizing agents to break down and remove organic pollutants from water or air. Studies have demonstrated that Advanced Oxidation Processes (AOPs) are efficient techniques for converting harmful organic pollutants into minerals and for eliminating contaminants from wastewater. Advanced oxidation processes (AOPs) that use ˙OH radicals (which have a standard electrode potential of 2.80 V/SHE) are very strong oxidizing agents that do not care what they are attacking. They can efficiently convert organic contaminants present in water into harmless smaller molecules such as CO2 and H2O.3–6 Photocatalysis, typically using UV-Vis light, generates ˙OH radicals. Advanced oxidation techniques have been used in the textile, pharmaceutical, printing, and pulp industries to treat wastewater.3,7,8 However, this technology is still in use. In practice, photocatalytic technology still has limitations in terms of quantum efficiency, durability, and catalyst cost. In the majority of individual semiconductor photocatalysts, the pace at which the photoelectron–hole pairs recombine is rapid, resulting in low quantum efficiency and performance.9,10 Several widely recognized semiconductors exhibit excellent photocatalytic activities owing to their large bandgap energies. UV radiation specifically activates TiO2, ZnO, CeO2, and SnO2. Consequently, their spectral response range is limited, preventing the effective absorption of sunlight, particularly visible light.11–13 Sulfide and nitride semiconductors have band gap energies in the visible range; however, they are not very stable and do not last long in wet environments.14 Hence, it is imperative to address the aforementioned constraints by creating novel semiconductor materials derived from conventional semiconductors.
Hematite, also known as Fe2O3, is a semiconductor that exhibits n-type conductivity and has a band gap energy ranging from approximately 2.0 to 2.2 electron volts (eV). Hematite can capture light ranging from yellow to ultraviolet, specifically at a wavelength of λ ≤ 600 nm. It can absorb up to 40% of the energy in the solar spectrum. In addition, hematite exhibits favorable environmental characteristics, demonstrates excellent stability in aqueous solutions with pH greater than 3, and is among the most cost-effective options for semiconductor materials.14–16 Thus, hematite is a highly intriguing choice for photocatalytic applications that specifically aim to harness visible light, particularly sunlight.17 Nevertheless, certain characteristics, such as the rapid recombination rate of electrons and holes, the limited diffusion length of holes (2–4 nm), and inadequate electrical conductivity, constrain the photocatalytic efficiency of Fe2O3, leading to diminished quantum optical efficiency.14 By combining with a second semiconductor, we can overcome these constraints and create a novel heterogeneous junction catalyst that offers exceptional benefits.18–23 ZnBi2O4, a bismuthate compound, is considered a promising candidate owing to its exceptional characteristics. ZnBi2O4 is a well-established p-type semiconductor with a direct bandgap ranging from 2.2 to 3.0 eV,24,25 which makes it highly appropriate for applications that utilize visible light or sun sources. ZnBi2O4 has a distinctive crystal structure, characterized by zigzag chains of metal–oxygen bonds formed by Zn and Bi atoms at the corners.26 Moreover, ZnBi2O4 demonstrates chemical stability, a high optical current density, and the ability to undergo oxidation by hydroxyl (˙OH) ions.27,28
Fe2O3 and ZnBi2O4 combinations can form n–p heterojunction semiconductors that are photocatalyst-driven under visible light or sun sources. Fe2O3 is an n-type semiconductor with the Fermi level located near the ECB, whereas ZnBi2O4 is a p-type semiconductor with the Fermi level located near the EVB. At the junction between the two semiconductors, electrons are transferred from Fe2O3 to ZnBi2O4 until their Fermi levels align, creating an electric field that can boost the transport and separation of photogenerated charges over ZnBi2O4.
In this study, a novel 3D Fe2O3@ZnBi2O4 n–p heterojunction was fabricated in two steps. Initially, the gel–polymer combustion technique produced Fe2O3 using tartaric acid and polyvinyl alcohol as the combustion agents. Next, ZnBi2O4 was introduced onto the Fe2O3 surface via microwave-assisted precipitation. In addition, the physicochemical properties and photocatalytic activities of the heterojunctions were studied and discussed in detail. Furthermore, the physicochemical characteristics and photocatalytic activity of the heterojunctions were comprehensively examined and analyzed. Additionally, experimental investigations into the function of oxidizing agents in the breakdown of MB have been conducted. A reaction mechanism for dye degradation was proposed.
To establish equilibrium between the adsorption and desorption of MB on the surface of the catalyst, the reaction mixture was stirred in the absence of light for 1 h as a step in the adsorption process.
Once the catalyst surface reaches the point of MB adsorption–desorption equilibrium, the breakdown process can be initiated by exposing it to light irradiation. To carry out the photocatalytic reaction, the light source was activated until most of the color was eliminated from the MB solution. A magnetic stirrer and circulating water were used to agitate the reaction mixture until it reached a steady state at ambient temperature. The catalyst was separated from the suspension by subjecting 3 mL of the mixture to centrifugation every 20 min. The MB concentration was determined using an ultraviolet-visible spectrophotometer.
Furthermore, SEM and TEM analysis were used to study the morphology and structure of the samples. The SEM image in Fig. 2A shows that the Fe2O3 material produced using the polymer gel combustion technique possessed a three-dimensional structure characterized by flat branches. These branches are formed from the amalgamation of several tiny planar particles during gel combustion. Besides, the FZB material (Fig. 2B) exhibited a 3D structure composed of two components. This structure consisted of flower-like clusters composed of ZnBi2O4, which were accompanied by long slender blades that covered the surface of the Fe2O3 branches. The Fe2O3 flat branches were seen to undergo rounder and larger size transformation after being subjected to microwave heating and calcination during the synthesis of FZB. In addition, the high-resolution TEM (HR-TEM) results confirmed the formation of an FZB heterojunction. The lattice fringes were approximately ∼0.370 nm which is ascribed to the (012) plane of α-Fe2O3,32,33 in which ∼0.292 nm belongs to the (102) plane of ZnBi2O4. Together with the XRD findings mentioned earlier, this provides evidence that the FZB material was effectively synthesized using a two-step process.
![]() | ||
Fig. 2 SEM images of (A) Fe2O3 and (B) FZB; (C) TEM, (D) HR-TEM images of FZB, (E–I) EDS mapping of FZB, and (J) EDS spectra of FZB. |
Moreover, EDS spectra and elemental mapping of FZB further confirmed the effective synthesis of FZB. Fig. 2E–J illustrate that the sample contained Zn, Fe, Bi, and O and exhibited a very regular distribution. No extraneous foreign components were identified in any of the samples.
The chemical bonds of the catalysts were studied by FT-IR spectroscopy (Fig. 3). Fig. 3A shows that the absorption peaks at approximately 462 and 555 cm−1 of the Fe2O3 material can be attributed to the stretching vibrations of Fe–O.20,34 For ZnBi2O4, the absorption peaks at 845 and 1390 cm−1 correspond to the Bi–O and Bi–O–Bi vibrations, respectively. Additionally, the peaks at 480, 528, and 575 cm−1 indicate the vibration of Zn–O bonds.25,26,31,35 It is worth mentioning that the absorption peaks in the FT-IR spectrum of the FZB material are very strong and show bond metal oxygens, such as Zn–O, Fe–O, Bi–O, and Bi–O–Bi. Additionally, vibration peaks were observed in the range of 3300–3500 cm−1 attributed to O–H from absorbed H2O. The FT-IR results further support the successful synthesis of n–p FZB heterojunctions.
The UV-Vis-DRS spectrum is associated with the optical absorption range and bandgap energy of the material. Fig. 3B demonstrates that in the ultraviolet light spectrum, the synthetic materials exhibited excellent absorption capabilities, with the order of absorption being ZBO < FZB < Fe2O3. In the visible light spectrum, the FZB and Fe2O3 materials exhibited significant absorption, whereas the ZBO material exhibited significantly weaker absorption. The optical absorption characteristics can be verified by estimating the position of the optical absorption edge and bandgap of the material using the Tauc diagram (Fig. 3B and C).36,37 The bandgaps of ZBO, Fe2O3, and FZB were measured at 2.84, 1.95, and 1.98 eV, respectively. The absorption wavelength of each material was determined using the formula λ = hc/Eg, resulting in values of 436, 635 nm for Fe2O3, and 626 nm for FZB. Thus, it can be deduced that ZBO exhibits a high capacity for absorbing ultraviolet rays and only partially absorbs visible light with a wavelength λ ≤ 436 nm. However, both the FZB and Fe2O3 materials possess a strong ability to absorb light across the ultraviolet and visible regions of the light spectrum. The advantageous attribute of FZB is highly advantageous for the practical implementation of the FZB material as a solar photocatalyst, as it enables the utilization of sunlight as a light source to initiate the photocatalytic process. The new catalyst (FZB) has notable advantages in this regard.
In addition, the optical characteristics of the FZB material were studied using photoluminescence (PL) spectroscopy, which enables the investigation of both electron–hole recombination and charge separation in materials. The photoluminescence (PL) spectra of ZnBi2O4, Fe2O3, and FZB were recorded at an excitation wavelength of 400 nm. Fig. 3D shows that all emission peaks are situated exclusively within the visible light range, specifically in the deep-level emission (DLE) region. The strongest emission peak at 436 nm is related to the rapid recombination probability of electron–hole pairs in ZnBi2O4, and the peak position agrees well with the bandgap (2.84 eV) of ZnBi2O4. Meanwhile, the broad peak with the highest intensity at 440 nm was associated with the atomic transitions of iron vacancies in Fe2O3.38 The weak shoulder peaks at 458 and 488 nm and a broad peak at 583 nm in the ZnBi2O4 PL spectra correspond to the 3P0,1 → 1S0 transitions of the Bi3+ ions.39
The formation of complexes on the heterogeneous surface reduced the PL intensity and peak position shift. The peak observed at 438 nm for the FZB composite exhibited a sharp decrease in intensity, and there was a slight red shift compared to that of pristine ZnBi2O4 (436 nm) and a slight blue shift compared to that of pure Fe2O3 (440 nm). The significant decrease in intensity indicates that after the formation of the heterojunction between Fe2O3 and ZnBi2O4, the rapid recombination of photogenerated free radicals was significantly inhibited. The broad monotonically shifted emission peak of the FZB nanocomposite was attributed to the presence of defect sites, leading to the generation of localized energy levels in the forbidden energy gap.
The surface chemistry of the synthesized ZnBi2O4 and FZB was analyzed using XPS. Fig. 4 shows the binding energies of the Zn 2p, O 1s, Fe 2p, and Bi 4f orbitals.
![]() | ||
Fig. 4 XPS spectra of (A) FZB heterojunction and high-resolution (B) O 1s, (C) Fe 2p, and (D) Bi 4f XPS spectra. |
The Zn 2p orbitals (Fig. 4A) of the FZB sample exhibit a complex electronic structure, as evidenced by the presence of two distinct peaks in the X-ray photoelectron spectroscopy (XPS) analysis. The peaks observed at 1021.54 eV and 1044.64 eV correspond to the energy levels of Zn 2p3/2 and Zn 2p1/2,40 respectively. These energy levels are characteristic of spin–orbit splitting in the Zn2p orbitals, providing valuable information about the electronic configuration of zinc atoms within the composite material. The slight redshift observed in the binding energy peaks of the FZB sample compared to those of ZnBi2O4 (ZBO) is a significant indicator of the successful formation of the composite material. This shift in the binding energies suggests a change in the chemical environment surrounding the zinc atoms, likely due to the interaction between the Fe2O3 and ZnBi2O4 components.41 The redshift may be attributed to factors such as charge transfer, changes in coordination, or alterations in the local electronic structure resulting from the integration of iron oxide into the zinc–bismuth oxide matrix.
The XPS spectrum of O 1s, as depicted in Fig. 4B, reveals two distinct peaks that provide valuable information regarding the oxygen species present in ZBO. The lattice oxygen (O2−) at 530.33 eV represents the oxygen ions incorporated into the crystal lattice structure of the material. The surface-adsorbed oxygen (O−) at 532.02 eV corresponds to the oxygen species absorbed on the surface of the material.42 The XPS analysis of the FZB sample revealed a notable shift in the two peaks towards higher energy levels, indicating significant changes in the electronic structure of the material. The most striking difference is observed in the relative intensity of the 532.65 eV peak, which is markedly higher in FZB than in ZBO. This increase in the peak intensity is a strong indicator of a higher concentration of oxygen vacancies within the FZB sample. Oxygen vacancies play a crucial role in the electronic and optical properties of metal oxides and often serve as active sites for various chemical reactions.
Fig. 4C shows two prominent peaks of the Fe 2p core level at 710.7 eV, and 724.4 eV attributed to Fe 2p3/2 and Fe 2p1/2 states.43,44 The reported binding energy values are consistent with the typical values for iron oxides, suggesting that the iron in the sample is likely to be in an oxidized state.
Fig. 4D shows that the two distinct peaks of 158.4 and 163.6 eV are related to the Bi 4f7/2 and Bi 4f5/2 spin-orbital pairs of Bi3+, for ZnBi2O4.45 The binding energies of Bi 4f7/2 (158.6 eV) and Bi 4f5/2 (163.8 eV) shifted to higher binding energies for FZB. This shift in binding energies suggests a change in the chemical environment of bismuth atoms when ZnBi2O4 is combined with Fe2O3.46 The lower binding energies observed in FZB may indicate a reduction in the oxidation state of Bi or alterations in its electronic structure.
Electrochemical impedance spectroscopy (EIS) and transient photocurrent measurements were performed to investigate the charge-transfer behavior. As illustrated in Fig. 5A, the Nyquist plot reveals that the arc radius of the two-component nanocomposite was smaller than that of the individual components. This indicates that the FZB heterostructure encounters fewer obstacles in the charge transfer across the interface, resulting in a higher charge separation efficiency. The reduced charge transfer resistance in the FZB heterostructure can be attributed to the synergistic effect between the two components, which facilitates electron–hole pair separation and enhances overall photocatalytic performance. This improvement in charge transfer dynamics is further supported by transient photocurrent measurements, which show a higher and more stable photocurrent response for the nanocomposite compared to the individual components. To validate these findings, the photocurrent responses of the samples were recorded over four on/off cycles (Fig. 5B). FZB exhibited significantly higher photocurrent responses than the Fe2O3 and ZnBi2O4 photocatalysts. These results confirm that the internal electric field generated at the n–p heterojunction interface between ZnO and ZnBi2O4 enhances charge separation and transport efficiency, consistent with the EIS and PL measurement results.
The efficacy of photodegradation is frequently affected by the solution pH, partly because of its association with the surface charge of the catalyst, which subsequently affects the adsorption of organic compounds onto the catalyst surface. In this experiment, the pH of the solution was calibrated within an interval of 4–10. Fig. 6C and D illustrate that the decolorization of MB by FZB21 was slow at solution pH values ranging from 4 to 7; however, its rate greatly increased at solution pH values ranging from 8 to 10. After 45 min of exposure to light, the reaction rate constants varied between 0.017 min−1 and 0.020 min−1 within the pH range of 4–7. Within the pH range of 8–10, the values varied from 0.0572 to 0.0853 min−1. The breakdown of MB was observed to be especially slow at pH values below 7, mostly because of the positive charge present on both the catalyst surface (pHz of FZB21 = 7.8) and MB+ ions. Thus, only a restricted amount of MB was absorbed onto the catalyst surface and underwent breakdown. At pH values greater than 7, the FZB21 surface had a negative charge, which facilitated the effective adsorption of MB onto its surface. Hence, the rate constant of the reaction increases and attains its maximum value at pH 10. Nevertheless, because of the highly alkaline pH of 10, this study employed a pH of 9.
Experimental investigations of the impact of the initial concentration of the MB solution on the effectiveness of MB decolorization are shown in Fig. 6E and F. The results indicated that the decline in the MB concentration exhibited a progressive reduction as the initial MB concentration increased from 30 to 60 ppm. At a concentration of 30 ppm, MB exhibited a reaction rate constant of 0.0698 min−1, 40 ppm at 0.0536 min−1, 50 ppm at 0.0294 min−1, and 60 ppm at 0.0206 min−1. The reduction in efficiency with an increasing initial MB concentration is attributed to the light-blocking phenomenon. The deeper hue of the solution obstructs the ability of the catalyst to absorb light, resulting in a reduction in the concentration of photoexcitation charge carriers, which is the process responsible for organic matter biodegradation.
Furthermore, the effect of the catalyst concentration was also investigated. Fig. 6G and H indicate that the efficiency of MB degradation approached 0.0310 min−1 when the catalyst concentration was 0.5 g L−1 and significantly increased to 0.0322 min−1 when the catalyst concentration was 0.7 g L−1. Nevertheless, the efficiency of MB degradation declined to 0.0283 and 0.0259 when the catalyst concentration was raised to 1.0 and 1.2 g L−1, respectively. The reduction in efficiency can be attributed to two factors: first, the surplus catalyst induces a light-blocking phenomenon, and second, it facilitates competition for the active sites on the catalyst surface.
To study the parameters influencing the efficiency of MB degradation, the optimal conditions for the photocatalytic reaction were determined as follows: FZB21 material, pH 9, MB concentration 60 ppm, catalyst content 0.7 g L−1.
Additionally, we performed a series of studies to evaluate the influence of H2O2 and K2S2O8 on the MB degradation efficiency under optimal conditions. Fig. 7A and B indicate that in the absence of light, the degradation of MB was significantly challenging in the presence of a catalyst and H2O2, yielding a rate constant of k = 0.00173 min−1 and a degradation efficiency of only 13.9%. In contrast, exposure to light resulted in a substantial reduction in the MB concentration. The reaction rate constants were k = 0.0322 min−1 for 0 mM H2O2, k = 0.0338 min−1 for 2 mM H2O2, k = 0.0645 min−1 for 4 mM H2O2, and k = 0.0695 min−1 for 8 mM H2O2. The addition of H2O2 increased the reaction rate by a factor of 2.0 with 4 mM H2O2 and by a factor of 2.15 with 8 mM H2O2. Doubling the concentration of H2O2 from 4 to 8 mM resulted in a small increase in the reaction rate constant, indicating that the 4 mM concentration of H2O2 was nearly optimal. The introduction of H2O2 initiated the photo-Fenton reaction (H2O2/Vis, Fe2O3/H2O2), augmenting the concentration of ˙OH and ˙HO2 in the solution (reactions (1)–(3)). This facilitated the expedited breakdown of MB47–49
Fe3+ + e− → Fe2+ | (1) |
Fe2+ + H2O2 → Fe3+ + ˙HO + OH− | (2) |
Fe3+ + H2O2 → H+ + Fe2+ + ˙HO2− | (3) |
![]() | ||
Fig. 7 Effect of H2O2 (A) and (B) and K2S2O8 (C) and (D) on MB degradation using the FZB heterojunction under visible light. |
Furthermore, the effect of S2O82− ions on the photodegradation efficiency of FZB was also studied (Fig. 7C and D). The reaction rate constant augmented by 1.5 times (k = 0.0486 min−1) for 1 mM K2S2O8, 5.7 times (k = 0.1836 min−1) for 2.0 mM K2S2O8, and 5.95 times (k = 0.1917 min−1) for 3 mM K2S2O8. Electrons generated using light interact with Na2S2O8 on the surface of the catalyst to create ROS radicals ˙SO4− and ˙OH (reactions (1), (4)–(7))50,51 which accelerates the reaction. The reaction slows the recombination of photogenerated electrons and holes, and K2S2O8, acts as an electron trapping agent. At the same time, it creates reactive oxygen species (ROS), which make the strong oxidizing agent more effective at breaking down methylene blue (MB).
Fe3+ + S2O82− → ˙S2O82− + Fe2+ | (4) |
S2O82− + Fe2+ → ˙SO4− + SO42− + Fe3+ | (5) |
S2O82− + e− → ˙SO4− + SO42− | (6) |
˙SO4− + H2O → ˙HO + SO42− + H+ | (7) |
The use of K2S2O8 significantly improved the effectiveness of MB degradation relative to that of H2O2. After 80 minutes of irradiation, the MB decomposition efficiency of FZB/H2O2 exceeded 99.4%, but the MB decomposition efficiency of FZB/K2S2O8 attained 99.9% after merely 40 minutes of irradiation. The reaction rate constant for FZB/K2S2O8 was 0.1917 min−1, which is 2.8 times greater than that of FZB/H2O2 at 0.695 min−1. There are several reasons why the MB decomposition reaction with K2S2O8 was much faster than that with H2O2. (i) The first is that the O–O bond energy in K2S2O8 is lower than that in H2O2 (213.3 kJ mol−1), which makes S2O82− more reactive than O22−.52 (ii) The second step involves the addition of K2S2O8, which generates radicals ˙SO4−, ˙S2O82−, and ˙HO (reactions (1), (2) and (4)–(7)), while the addition of H2O2 produces ˙HO and ˙HO2− (reactions (1)–(3)). The standard reduction potential of ˙SO4− (2.5–3.1 V) exceeds that of ˙HO (1.7–2.7 V) and ˙HO2− (1.8 V). At neutral pH, the ˙SO4− radical has a higher standard oxidation potential than the hydroxyl radical. It can successfully interact with organic molecules across a wide pH range of 2–8. (iii) The third is the sulfate radical, which is more selective and effective than ˙HO radicals and can transfer electrons to organic molecules with aromatic or unsaturated bonds.53 (iv) Finally, the half-life of ˙SO4− is 30–40 μs, exceeding that of ˙HO at 1 μs, and it can oxidize contaminants at an exceptionally high rate of 106–109 M s−1.54,55 The prolonged half-life enhances substantial interactions and mass transfer between ˙SO4− radical and reactants.
These results demonstrate that the FZB catalyst effectively degrades MB under visible light, demonstrating its superior photocatalytic capabilities. The incorporation of minor quantities of oxidants (hydrogen peroxide and sodium persulfate), especially persulfate, significantly improved this characteristic.
![]() | ||
Fig. 8 Effect of radical scavengers (A and B) on MB degradation using FZB. (C) Mott–Schottky plots of Fe2O3, ZBO, and FZB. |
Fig. 8C illustrates that the Mott–Schottky (M–S) plots exhibit a positive slope for n-type semiconductors and a negative slope for p-type semiconductors. The M–S plot of Fe2O3 exhibited a positive slope, signifying n-type semiconductor characteristics. The M–S curve of ZnBi2O4 has a negative slope, confirming its p-type characteristics. The flat band potentials of Fe2O3 and ZnBi2O4 were ascertained to be +0.28 V and +2.20 V vs. Ag/AgCl at pH 7 (corresponding to +0.485 V and +2.425 V vs. NHE), respectively. In n-type semiconductors, the conduction band (ECB) is generally positioned 0.1 V above the flat potential, whereas in p-type semiconductors, the valence band (EVB) is typically situated 0.1 V below the flat potential. The ECB of Fe2O3 was established at +0.385 V, whereas the EVB of ZnBi2O4 was computed to be +2.525 V.
The EVB of Fe2O3 and the ECB of ZnBi2O4 were computed using the formula EVB = ECB + Eg, where Eg represents the energy bandgap. The band gap values of Fe2O3 and ZnBi2O4 are 1.95 eV and 2.84 eV, respectively, with the EVB of Fe2O3 estimated at +2.355 V and the ECB of ZnBi2O4 estimated at −0.315 V. The photocatalytic mechanism for the Fe2O3@ZnBi2O4 nanocomposite is depicted in Fig. 9, based on these analytical and computational findings.
The introduction of p-type ZnBi2O4 onto the surface of n-type Fe2O3 resulted in the formation of an n–p heterojunction. This configuration induces fluctuations in the energy bands of ZnBi2O4 and Fe2O3 until achieving Fermi equilibrium. An internal electrostatic field is formed from Fe2O3 to ZnBi2O4, which facilitates charge separation. Upon exposure to visible light, ZnBi2O4 and Fe2O3 are stimulated to produce electron–hole pairs (e−/h+) (reaction (8)). The conduction band level of ZnBi2O4 was inferior to that of Fe2O3 following the formation of the n–p heterojunction. This indicates that the excited electrons of ZnBi2O4 can readily transition into the conduction band (CB) of Fe2O3. Conversely, the valence band (VB) level of Fe2O3 is more positive than that of ZnBi2O4, allowing the excited holes of Fe2O3 to transfer to the VB level of ZnBi2O4. The internal electric field at the interface between Fe2O3 and ZnBi2O4 propels these reactions. The internal electric field of the p–n heterojunction efficiently segregates the photogenerated charges.61 This significantly reduced the recombination of e−/h+ pairs in the Fe2O3/ZnBi2O4 heterojunction. Photoluminescence studies, photocurrent response, electrochemical impedance spectroscopy, and radical scavenger investigations validated the capacity for charge separation.
Moreover, the electrons at the heightened conduction band level of ZnBi2O4 can reduce the adsorbed O2 on the surface to form superoxide anion radicals (˙O2−) (reaction (8)), while the holes at the lowered valence band level of Fe2O3 can oxidize the adsorbed H2O on the surface to yield hydroxyl radicals (˙OH) (reaction (9)). Owing to the abundance of electrons at the conduction band level in Fe2O3, Fe3+ can undergo oxidation to Fe2+ (reaction (11)), resulting in the formation of ˙OH radicals (reaction (12)). The radicals exhibiting potent oxidizing characteristics, along with vacancies in the valence band of ZnBi2O4 and electrons in the conduction band of Fe2O3, contribute to the mineralization of the dye into CO2, H2O, and other inorganic byproducts (reaction (19)).
In the presence of H2O2, photo-Fenton processes occur, generating additional ˙OH and ˙HO2− radicals. This accelerated the reaction and enhanced its efficacy for the decomposition of MB. It was specifically the presence of K2S2O8 that sped up the breakdown of MB by creating ˙SO4− and ˙S2O8− radicals, which then increased the amount of ˙OH (reactions (15)–(18)). It is known that the radicals ˙SO4−, ˙S2O8−, and ˙HO are all strong oxidizing agents. The ˙SO4− radical strongly reacts with aromatic chemicals. The addition of persulfate significantly accelerated the MB breakdown reaction in the presence of the FZB catalyst and visible light (reaction (20)).
(p–n)FZB + hυ → e− + h+ | (8) |
e− + O2 → ˙O2− | (9) |
h+ + H2O → ˙OH + H+ | (10) |
e− + Fe3+ → Fe2+ | (11) |
Fe2+ + H2O + e− → ˙OH + Fe3+ + H+ | (12) |
Fe2+ + H2O2 → Fe3+ + ˙HO + OH− | (13) |
Fe3+ + H2O2 → H+ + Fe2+ + ˙HO2− | (14) |
Fe3+ + S2O82− → ˙S2O8− + Fe2+ | (15) |
S2O82− + Fe2+ → ˙SO4− + SO42− + Fe3+ | (16) |
S2O82− + e− → ˙SO4− + SO42− | (17) |
˙SO4− + H2O → ˙HO− + SO42− + H+ | (18) |
e−, h+, ˙OH, ˙O2− + MB → CO2 + H2O + inorganic products | (19) |
e−, h+, ˙OH, ˙O2−, ˙HO2− + MB → CO2 + H2O + inorganic products | (20) |
e−, h+, ˙OH, ˙O2−, ˙SO4−, ˙S2O8− + MB → CO2 + H2O + inorganic products | (21) |
This journal is © The Royal Society of Chemistry 2025 |