Next Article in Journal
Nickel(II) N-Heterocyclic Carbene Complex for the Hydrogenation of 2-Acetylpyridine under Mild Conditions
Previous Article in Journal
Vanadium Oxide-Based Cathode Materials for Aqueous Zinc-Ion Batteries: Energy Storage Mechanism and Design Strategy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Z-Scheme CoFe2O4-Loaded g-C3N4 Photocatalyst with High Degradation Efficiency of Methylene Blue under Visible-Light Irradiation

1
Department of Industrial Chemistry, Addis Ababa Science and Technology University, Addis Ababa 16417, Ethiopia
2
Nanotechnology Center of Excellence, Addis Ababa Science and Technology University, Addis Ababa 16417, Ethiopia
3
School of Mechanical and Electrical Engineering, University of Electronic Science and Technology of China, Chengdu 611731, China
4
Institute of Applied Physics, Vienna University of Technology, Wiedner Hauptstraße 8-10, 1040 Vienna, Austria
*
Authors to whom correspondence should be addressed.
Inorganics 2023, 11(3), 119; https://doi.org/10.3390/inorganics11030119
Submission received: 15 February 2023 / Revised: 7 March 2023 / Accepted: 11 March 2023 / Published: 13 March 2023

Abstract

:
Magnetically recyclable direct Z-scheme CoFe2O4-loaded g-C3N4 photocatalyst material was fabricated using a facile hydrothermal technique and subsequently characterized by XRD, VSM, PL, FT-IR, EDX, DRS, SEM, and BET techniques. The characterization results confirmed that nanoparticles of CoFe2O4 are loaded on the surface of g-C3N4 sheets. The optical band gap of g-C3N4 has been decreased from 2.65 eV to 1.30 eV by means of the loading of CoFe2O4 nanoparticles onto the nanosheets of g-C3N4. This has enhanced the separation process of electron-hole. Under visible light irradiation, the photocatalytic activity of the developed direct Z-scheme CoFe2O4-loaded g-C3N4 photocatalyst was evaluated for the photodegradation of methylene blue (MB); during this process the MB decomposed by up to 98.86% in 140 min. Meanwhile, under the same irradiation and time conditions, the g-C3N4 and CoFe2O4 themselves degraded MB up to 74.92% and 51.53%, respectively. The direct Z-scheme CoFe2O4-loaded g-C3N4 material was recovered from the solution after the photocatalytic activity using an external magnet and studied to determine its stability. It was shown that the photoactivity did not change significantly after five consecutive cycles.

Graphical Abstract

1. Introduction

With the increasing expansion of global industry, the problem of access to clean water has risen to the forefront of discussion. Every year, a massive volume of industrial dye wastewater is dumped into natural water bodies, posing a risk to both the aquatic ecosystem and people [1]. Nowadays, removal of organic pollutants from wastewater/water through photocatalysis provides a viable method for solving environmental issues [2,3]. In order to eliminate environmental pollution and degrade these harmful organic pollutants, photocatalytic technology has received an increasing amount of interest. So far, TiO2 is the most widely studied photocatalyst; it was reported by Fujishima and Honda in 1972 for use in photoelectrochemical water splitting [4]. This study was a pioneer in photocatalysis technology. Nevertheless, TiO2 has high band gap (3.2 eV), which accounts for around 4% of sunlight; this severely restricts its ability to function as a photocatalyst and restricts the range of use of this material [5,6]. These limitations have inspired researchers to develop novel materials with narrow band gaps (Eg) to utilize solar energy in a more efficient way. Due to its chemical inertness, distinctive layered structure, nontoxic nature, a middle band (i.e., band gap between 2.4–2.9 eV), and capacity to absorb visible light, graphitic carbon nitride (g-C3N4) is now frequently employed for the photocatalytic degradation of organic contaminants [7,8,9,10,11,12,13,14]. However, inadequate absorption of visible light, small specific surface area, and fast electron-hole pair recombination still limits its photocatalytic activity [15,16].
Thus, numerous techniques were exploited to improve graphitic carbon nitride’s photocatalytic efficiency, including doping [12], morphological modification [17], copolymerization [18], and combining with one or more semiconductor materials [19]. Thanks to its appropriate valence band edge and conduction band edge positions, g-C3N4 can produce a heterojunction structure/nanocomposite with other semiconductor photocatalysts. This could improve its photocatalytic activity performance by improving the photogenerated electron-hole separation rate of the heterojunction photocatalyst [20,21,22]. Currently, spinel ferrites with the general formula of MFe2O4 (M = Zn, Co, Ni, Cu) have been widely used in photocatalytic applications due to their visible-light absorption, stability, environmentally friendly nature, magnetic properties, and low-cost [23]. Cobalt ferrite (CoFe2O4) is one of the inverse spinel structures that attracted significant interest because of its non-toxic nature and its abundant and narrow bandgap (1.2–2.7 eV) [24,25]. Even though pristine CoFe2O4 has a poor photocatalytic performance, its photocatalytic activity considerably increases when it is coupled with a π-conjugated semiconductor materials [26]. For instance, CoFe2O4 nanoparticles can be coupled with g-C3N4 to prepare heterojunction structures, direct Z-schemes, or all-solid-state Z-schemes due to the matching conduction band edge position and valence band edge position of CoFe2O4 and g-C3N4 semiconductors.
Huang et al. [27] reported on the CoFe2O4/g-C3N4 composites, which are synthesized using an easy calcination process. The composite containing 41.4 wt% CoFe2O4 showed the maximum catalytic activity for the MB degradation under irradiation of visible light in the presence of H2O2. However, the use of H2O2 may not be economically viable during practical applications. Inbaraj et al. [28] also synthesized CoFe2O4/g-C3N4 composite via the coupling of a honey-mediated green approach and hydrothermal technique. The MB degradation and the adsorption removal of lead (Pb2+) ion from water were both used to test the photocatalytic activity of the composite photocatalyst. However, the hydrothermal process used in the synthesis is time-consuming and labor-intensive; our experimental strategy seems to be simpler. Moreover, the photocatalytic mechanism was a type-II heterojunction, which means the redox reaction takes place in the less negative conduction band/reduction potential and in the less positive valence band/oxidation potential of the semiconductors. This limits the photocatalytic performance of the CoFe2O4/g-C3N4 composites. Recently, in the presence of peroxymonosulfate (PMS), which is utilized as a sulfate radical-based Fenton-like oxidation reaction, Guo et al. [1] reported that rhodamine B (RhB) was photodegraded by CoFe2O4@g-C3N4 photocatalyst material. The Fen+/PMS system must be run at an acidic pH because of the precipitation and hydrolysis of iron ions; doing so after the reaction results in additional operational costs [29,30]. These inescapable disadvantages limit the widespread application of the homogeneous iron/PMS approach in wastewater treatment technology.
Even though several research have been widely published about the photocatalytic activity of the CoFe2O4/g-C3N4 composite and its ability to photodegrade several organic dyes/pollutants, the literature currently available falls significantly short in addressing crucial variables, including (1) construction of the direct Z-scheme structure and (2) the photoactivity of the composite in the absence of a Fenton-like system. Hence, to overcome these challenges, direct Z-scheme CoFe2O4-loaded g-C3N4 photocatalyst material was successfully fabricated using the facile hydrothermal method.
The photodegradation activity of the obtained direct Z-scheme CoFe2O4-loaded g-C3N4 photocatalyst was assessed for methylene blue (MB) photodegradation under irradiation of visible light and shows degradation activity 6.6 and 3.6 times higher than pristine CoFe2O4 and pure g-C3N4 photocatalysts, respectively. Such an extraordinary enhancement of catalytic activity under irradiation of visible-light could be due to the significantly increased electron-hole separation in the direct Z-scheme g-CoFe2O4-loaded g-C3N4, which improves the oxidation/reduction ability of the photocatalytic reaction. Moreover, the direct Z-scheme g-CoFe2O4-loaded g-C3N4 photocatalyst was collected from the aqueous solution without significant loss and demonstrated a negligible decline in performance throughout five cycles.

2. Results and Discussion

2.1. Characterization of the Photocatalyst

The FT-IR spectra of CoFe2O4, g-C3N4, and CoFe2O4-loaded g-C3N4 are displayed in Figure 1. As shown in Figure 1a, the broad bands in the 3086–3252 cm−1 matches to the -NH stretching mode at the uncondensed sites of the aromatic structures [31]. The peaks at 1572 and 1639 cm−1, and the other four peaks at 1407, 1461, 1316, and 1236 cm−1, represented typical C=N stretching vibration modes and C-N stretching of g-C3N4, respectively. The intense peak at 812 cm−1 can be attributed to the tri-s-triazine units’ breathing mode [7,32,33]. In spectrum of CoFe2O4, in the range of 587 and 419 cm−1, there were two prominent absorption bands that are ascribed to both the tetrahedral and octahedral metal-oxygen (M-O) vibrational modes links in the spinel lattice of CoFe2O4 nanoparticles, as shown in Figure 1b [34]. The H-O-H bending vibration of absorbed or free water molecules is responsible for the peak seen at 3408 cm−1. All the important characteristic peaks of CoFe2O4 and g-C3N4 appeared in the CoFe2O4-loaded g-C3N4 material; this confirmed the formation of CoFe2O4-loaded g-C3N4 photocatalyst material (Figure 1c).
XRD spectra of all the photocatalysts were taken for phase identification. Figure 2 displays the XRD spectra of CoFe2O4, g-C3N4, and CoFe2O4-loaded g-C3N4 photocatalysts. The diffraction peaks at 30°, 35.62°, 43.28°, 53.78°, 57°, and 62.27° are willingly attributed to the face centered cubic inverse spinel crystal structure of CoFe2O4 (JCPDS file no: 22-1086) [35]. The XRD spectrum from pure g-C3N4 exhibited peaks at 27.52° and 12.74°; these can be assigned to the (002) and (100) crystal planes of polymeric g-C3N4, respectively. In the spectrum of the CoFe2O4-loaded g-C3N4, all diffraction peaks of g-C3N4 and CoFe2O4 are observed, indicating that the CoFe2O4-loaded g-C3N4 photocatalyst materials were successfully prepared. Moreover, the Scherrer’s formula was used to determine the average crystallite size of CoFe2O4-loaded g-C3N4 photocatalyst and [36,37]; this was determined to be 13.43 nm.
The surface morphology and texture of the photocatalysts were investigated using SEM; the image is presented Figure 3a–c. The cubic structure of CoFe2O4 and crumpled sheets g-C3N4 are observed in Figure 3a,b, respectively. It has also been observed that numerous nanoparticles of CoFe2O4 were loaded on the g-C3N4 sheet surface (Figure 3c), suggesting the formation of CoFe2O4-loaded g-C3N4 photocatalyst. The elemental composition of the fabricated samples was examined using EDX; the spectra is displayed in Figure 3d–f. As shown in Figure 3a, the EDX peaks of the pristine CoFe2O4 sample are accredited to the elements Fe, Co, and O. The peaks in the EDX of neat g-C3N4 are linked to the elements C and N, as shown in Figure 3e. Also, the equivalent EDX spectrum of g-C3N4 that has CoFe2O4 loaded on it exhibits the presence of C, N, Co, Fe, and O, as shown in Figure 3f. This confirms the CoFe2O4-loaded g-C3N4 photocatalyst was synthesized without any additional impurities.
The magnetic possessions of the CoFe2O4-loaded g-C3N4 material and pristine CoFe2O4 photocatalysts were studied; the magnetization hysteresis curves are shown in Figure 4. The saturation magnetization (Ms) of pristine CoFe2O4 and CoFe2O4-loaded g-C3N4 photocatalyst was determined to be 43.4 emu/g and 41.9 emu/g, respectively. As a result, the CoFe2O4-loaded g-C3N4 photocatalyst could be easily collected using a magnet.
The pore size distribution, surface area, and pore volume of the CoFe2O4-loaded g-C3N4 photocatalyst were studied using N2 adsorption and desorption isotherms. The type IV isotherm and H3 hysteresis loop of the N2 adsorption-desorption curve are clearly visible in Figure 5a, confirming the mesoporous nature of the CoFe2O4-loaded gC3N4 photocatalyst [38]. The Barrett–Joyner–Halenda equation and Brunauer–Emmett–Teller (BET) method were used to determine the photocatalyst’s pore size distribution and surface area. The BET specific surface area of CoFe2O4-loaded g-C3N4 photocatalyst was determined to be 63.632 m2/g. Furthermore, the pore volume and pore size of the CoFe2O4-loaded g-C3N4 photocatalyst was 4.29 nm and 0.247 cm3/g, respectively, as presented as Figure 5b.
The UV−vis DRS spectra of g-C3N4, CoFe2O4, and CoFe2O4-loaded g-C3N4 samples are shown in Figure 6a. The pristine g-C3N4 photocatalyst possessed an absorption edge at about 500 nm. The CoFe2O4 nanoparticle showed a wide absorption range (200 to 900 nm). The absorption efficiency of the g-C3N4 photocatalyst was significantly extended in the visible-light region after loading nanoparticles of CoFe2O4 onto the g-C3N4 (Figure 6a). As a result, the boosted visible light absorption of the CoFe2O4-loaded g-C3N4 led to the production of extra electron−hole pairs under irradiation of visible-light, which boosted its photocatalytic activity. Furthermore, the band gap energy (Eg) of the photocatalysts were assessed using Tauc’s equation and the results are displayed in Figure 6b,c. The obtained band gap (Eg) values for pristine CoFe2O4 and g-C3N4 are 1.30 eV and 2.65 eV, respectively. However, after loading the CoFe2O4 on the surface of g-C3N4 nanosheets, its band gap energy (Eg) decreased to 1.42 eV. This result confirmed that the visible light absorption efficiency of g-C3N4 was significantly improved after modification by CoFe2O4 nanoparticles.
The photogenerated produced electron-hole pairs separation efficiency of the g-C3N4 and CoFe2O4-loaded g-C3N4 photocatalysts was examined using PL; the result is presented in Figure 6d. The PL strength of the CoFe2O4-loaded g-C3N4 is enormously diminished when compared to that of the pristine g-C3N4. This result shows that the rate of electron-hole separation efficiency in the CoFe2O4-loaded g-C3N4 was better than that in the g-C3N4 material. This could be because of the transfer of the electron-hole pairs between CoFe2O4 and g-C3N4 in the CoFe2O4-loaded g-C3N4.

2.2. Photocatalytic Activity

Figure 7a depicts the MB degradation under irradiation of visible light without photocatalyst as well as with CoFe2O4, g-C3N4, and CoFe2O4-loaded g-C3N4 photocatalysts. Figure 7a demonstrates that, without a photocatalyst, the MB is practically stable for 140 min when irradiated using visible light. This demonstrates that the presence of a photocatalyst caused MB degradation. The CoFe2O4-loaded g-C3N4 photocatalyst degraded 98.86% of the MB within 140 min under irradiation of visible light. However, under the same irradiation and time conditions, the g-C3N4 and CoFe2O4 degraded MB up to 74.92% and 51.53%, respectively. Therefore, the photodegradation rates of MB by the CoFe2O4-loaded g-C3N4 photocatalyst was better than that of pure CoFe2O4 and pristine g-C3N4 photocatalysts. In addition, the rate constant (kap) of degradation of MB by the CoFe2O4, g-C3N4, and CoFe2O4-loaded g-C3N4 photocatalysts can be expressed by a pseudo-first-order kinetics equation: ln(Co/C) = kapt, where Co is initial concentration, C is concentration after irradiation at a time t, and kap is apparent pseudo-first-order rate constant (Figure 7b). As shown in Figure 7b, the kap of g-C3N4 is higher than that of CoFe2O4. However, after loading CoFe2O4 on the g-C3N4 sheets surface, its kap increased significantly. As a result, CoFe2O4-loaded g-C3N4 photocatalysts showed photodegradation activity with 6.6 and 3.6 times higher kap than pristine CoFe2O4 and g-C3N4, respectively. This might be a result of the components’ synergistic interaction. Table 1 compares the degradation efficiency of different materials from the literature with the fabricated direct Z-scheme CoFe2O4-loaded g-C3N4 photocatalyst material for the photodegradation of organic dyes in water.
The stability of CoFe2O4-loaded g-C3N4 photocatalyst was examined by recycling and using it for repeated photocatalytic reactions. The CoFe2O4-loaded g-C3N4 photocatalyst exhibits nearly the same photocatalytic activity after five consecutive cycles of the degradation reaction, as displayed in Figure 8, demonstrating the produced photocatalyst’s exceptional stability. This could be due to the good adherence of the CoFe2O4 nanoparticles on the g-C3N4 sheets.
In order to explore the principal active species in the photocatalytic process, reactive species detection tests were also conducted. As superoxide radical (•O2), hydroxyl radical (•OH), and hole (h+) scavengers, p-benzoquinone (BQ), isopropanol (IPA), and formic acid (FA) were introduced at the same concentration (0.5 mmol/L), respectively [39,40]. As shown in Figure 9, the addition of IPA has an insignificant effect on the degradation efficiency, while the addition of BQ and FA diminished the degradation efficiency to 37.8% and 62.54%, respectively. Therefore, •O2 and h+ are the primary reactive species in the photocatalytic process, and •O2 has a greater influence on the photodegradation process than h+.
The boosted photocatalytic efficiency of the synthesized CoFe2O4-loaded g-C3N4 photocatalyst can be credited to the visible light sensitization of the g-C3N4 by the CoFe2O4 nanoparticles. To clarify the separation of electron-hole in the CoFe2O4-loaded g-C3N4 photocatalyst, the valence band (EVB) edge and conduction band (ECB) edge positions of the photocatalysts were calculated by Equations (1) and (2) [40].
E C B = X E 1 2 E g
E V B = E C B + E g
where ECB is conduction band potential, EVB is valence band potential, χ is the absolute electronegativity of the photocatalyst (the χ values of g-C3N4 is 4.73 eV and χ values CoFe2O4 is 5.81 eV [3,41]), E is the hydrogen scale’s free electron energy, which is 4.5 eV, and Eg is band-gap energy. Hence, it was discovered that CoFe2O4 and gC3N4 have band-gap energies of 1.30 eV and 2.65 eV, respectively (Figure 6b,c). Then, the values of EVB and ECB for CoFe2O4 and g-C3N4 were calculated using Equations (1) and (2), respectively. The obtained results are shown in Table 2.
Photogeneration electron-hole (e-h+) pairs are formed by both g-C3N4 and CoFe2O4 of the CoFe2O4-loaded g-C3N4 after irradiation by sufficient photon energy. The ECB of CoFe2O4 (+0.66 eV) is less negative than that of the g-C3N4 (−1.09 eV), and the EVB of CoFe2O4 (+1.96 eV) is more positive than the EVB of g-C3N4 (+1.56 eV). In the case of type II heterojunction phase way, the e in the ECB of g-C3N4 is transferred to the ECB of CoFe2O4 and the h+ in the EVB of CoFe2O4 are transferred to EVB of the g-C3N4. These interfacial e-h+ transfers suppress the charge carrier recombination and improve its photocatalytic activity. But, the e at ECB of CoFe2O4 (+0.66 eV) cannot reduce O2 to •O2 (E° = −0.33 eV) [16]. Alternatively, the h+ at EVB of g-C3N4 (+1.56 eV) has insufficient oxidation capacity to oxidize H2O or OH into •OH (•OH/H2O, E° = +1.99 eV) [16]. This contradicts the experiments on active species trapping. Clearly, the type-II photogenerated electron-hole (e-h+) mechanism was not appropriate for the CoFe2O4-loaded g-C3N4 photocatalytic system. Hence, the e-h+ mechanism of the CoFe2O4-loaded g-C3N4 may be illustrated by the direct Z-scheme system (Figure 10). In the direct Z-scheme system, the e in the ECB of CoFe2O4 are tending to transfer and recombine with the h+ in the EVB of g-C3N4 by driving by the internal electric field at the intersection of surfaces, resulting in the gathering of e and h+ in the ECB of g-C3N4 and the EVB of the CoFe2O4, respectively [42,43]. Consequently, the e in the ECB of g-C3N4 can easily reduce the O2 into •O2 as shown in Figure 10. Even though these h+ in the EVB of the CoFe2O4 cannot oxidize H2O or OH into •OH, they can directly participate in MB pollutants adsorbed onto the surface of the CoFe2O4-loaded g-C3N4 photocatalyst [38]. The experiments on active species trapping are compatible with this strategy. As such, the CoFe2O4-loaded gC3N4 photocatalyst in the present direct Z-scheme photocatalytic system not only improves the transfer and separation efficiency of photogenerated charge carriers but also retains a high degree of redox ability during the photodegradation of MB pollutant.

3. Materials and Methods

3.1. Chemicals

Iron(III) chloride hexahydrate (FeCl3.6H2O, 97%), sodium hydroxide (NaOH, 98%), Cobalt(II) chloride hexahydrate (CoCl2.6H2O, 98%), Urea (NH2CONH2, 99.5%), and nitric acid (HNO3, 70%) were obtained from Merck (India). All chemicals were utilized directly without further purification, and all experimental work was conducted with distilled water.

3.2. Synthesis of Photocatalysts

Graphitic carbon nitride (g-C3N4) was prepared by thermal polycondensation of urea; the detailed procedure has been described in our previously published article [7]. The CoFe2O4 loaded g-C3N4 photocatalyst was fabricated using the hydrothermal technique. Typically, 0.5 g of g-C3N4 was put in 50 mL of distilled water and sonicated for 40 min to form a uniform suspension (solution A). Similarly, 0.491g of FeCl3·6H2O (1.82 × 10−3 mol) and 21.62 mg of NiCl2·6H2O (9.1 × 10−5 mol) were put in 50 mL of distilled water and sonicated for 20 min (solution B). After that, the two solutions (solution A and solution B) were combined under sonication for 60 min. The resulting mixture was put in the Teflon-lined autoclave at 150 °C for 8 h. Then, the mixture was centrifuged and washed several times using ethanol and distilled water separately. The obtained sample was dried at 60 °C in a hot air oven to obtain the CoFe2O4 loaded g-C3N4 photocatalyst. Likewise, we prepared CoFe2O4-loaded g-C3N4 photocatalysts using the same procedure for comparison study.

3.3. Characterization Techniques

The crystallinity of the prepared samples was performed using X-ray Diffraction (D8 XRD, Bruker AXS) with Cu Kα radiation (λ = 0.154060 nm) over 2θ range of 7° to 70°. The molecular structure of the samples was examined using Fourier transform infrared spectroscopy (FT-IR, PerkinElmer–Frontier MIR/FIR) scanned in the range of 4000 to 400 cm−1. The morphology and elemental composition were determined using scanning electron microscopy (SEM, JEOL- JSM 6390LV) with energy-dispersive X-ray spectroscopy (EDX, Oxford Instruments). The UV–vis diffuse reflectance (DRS) was carried out on a UV–VIS spectrophotometer (Shimadzu–UV-2450) in the range of 200–900 nm by using BaSO4 as the reference. The Kubelka–Munk function (F(R) was then used to determine the optical bandgap energy of the photocatalysts. The recombination rate photoluminescence (PL) spectra were obtained by Fluorescence spectrophotometer (PerkinElmer–LS 55). By employing a BET surface area analyzer (NOVA 1000E, Quantachrom) to conduct a N2 adsorption-desorption investigation, the BET and pore size distribution of the sample were examined. A vibrating sample magnetometer (VSM, Lakeshore-7410) was used to examine the magnetic characteristics.

3.4. Photocatalytic Experiments

To study the photocatalytic activity of the photocatalysts in the breakdown of methylene blue (MB) under irradiation of visible light, several photocatalytic experiments were carried out. To attain an equilibrium between the photocatalyst and the dye, 30 mg of photocatalyst was suspended in 100 mL of MB aqueous solution (10 mg/L) and swirled in the dark for 60 min. A 10 W LED lamp (Havells, India) was used to irradiate the photocatalytic system and a 3 mL aliquot part; these were withdrawn from the reaction solution at 20 min intervals. Then, the magnetic photocatalysts was separated using a magnet and nonmagnetic photocatalysts were separated using a centrifuge (Universal 320 Hettich). The residual concentration of MB was analyzed using a UV-visible spectrophotometer (Agilent Cary 60) at a 664 nm. The primary active species were trapped using scavengers including formic acid (FA, h+ scavenger), benzoquinone (BQ, •O2 radical scavenger), and 2-propanol (2-PA, •OH radical scavenger).

4. Conclusions

In conclusion, we fabricated magnetically separable direct Z-scheme CoFe2O4-loaded g-C3N4 using a facile and simple hydrothermal technique. Characterization results confirmed that the CoFe2O4-loaded g-C3N4 was prepared successfully. The DRS study also showed that the absorption efficiency of the g-C3N4 photocatalyst was significantly extended in the visible-light region after loading nanoparticles of CoFe2O4 onto the g-C3N4 nanosheets. The photocatalytic performance of CoFe2O4, g-C3N4, and CoFe2O4-loaded g-C3N4 photocatalysts were examined for the MB degradation under irradiation of visible-light. The direct Z-scheme CoFe2O4-loaded g-C3N4 shows supreme degradation efficiency, and its activity is about 6.6 and 3.6 times higher than the CoFe2O4 and g-C3N4 photocatalysts, respectively. Such an extraordinary activity enhancement under irradiation of visible-light was possibly due to the significantly promoted electron-hole separation in the direct Z-scheme g-CoFe2O4-loaded g-C3N4; this improves the reduction/oxidation ability of the photocatalytic reaction. Mechanisms elucidated by scavenger studies revealed that •O2 and holes were the primary reactive radicals responsible for the degradation of MB. Moreover, the direct Z-scheme g-CoFe2O4-loaded g-C3N4 photocatalyst was collected from the aqueous solution without significant loss and demonstrated a negligible catalytic performance decline over five consecutive cycles. Therefore, this study can deliver novel insights for design and preparation of magnetic direct Z-scheme photocatalysts for dye degradation applications.

Author Contributions

The major work for this article, including conceptualization, methodology, software, validation, and writing the original draft, was performed by the first author (G.G.). The second author (M.G.) participated in conceptualization, software, and drawing. The final three authors (M.T., B.L. and W.L.) were responsible for conceptualization, supervising, and editing the article. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the internal grant of Addis Ababa Science and Technology University Ref. No. EG-51/11-1/20.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge Tezpur University, Assam, India, for some characterization facility support. We are also thankful to Addis Ababa Science and Technology University, Ethiopia, for the financial support from the Internal grant project (Ref. No. EG-51/11-1/20).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guo, X.; Ai, S.; Yang, D.; Zhao, L.; Ding, H. Synergistic photocatalytic and Fenton-like degradation of organic contaminants using peroxymonosulfate activated by CoFe2O4@g-C3N4 composite. Environ. Technol. 2021, 42, 2240–2253. [Google Scholar] [CrossRef] [PubMed]
  2. Hassani, A.; Eghbali, P.; Ekicibil, A.; Metin, Ö. Monodisperse cobalt ferrite nanoparticles assembled on mesoporous graphitic carbon nitride (CoFe2O4/mpg-C3N4): A magnetically recoverable nanocomposite for the photocatalytic degradation of organic dyes. J. Magn. Magn. Mater. 2018, 456, 400–412. [Google Scholar] [CrossRef]
  3. Mousavi, M.; Habibi-Yangjeh, A.; Abitorabi, M. Fabrication of novel magnetically separable nanocomposites using graphitic carbon nitride, silver phosphate and silver chloride and their applications in photocatalytic removal of different pollutants using visible-light irradiation. J. Colloid Interface Sci. 2016, 480, 218–231. [Google Scholar] [CrossRef]
  4. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, Y.; Luan, J. Synthesis, Property Characterization and Photocatalytic Activity of the Novel Composite Polymer Polyaniline/Bi2SnTiO7. Molecules 2012, 17, 2752–2772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Gebrezgiabher, M.; Gebreslassie, G.; Gebretsadik, T.; Yeabyo, G.; Elemo, F.; Bayeh, Y.; Thomas, M.; Linert, W. A C-Doped TiO2/Fe3O4 Nanocomposite for Photocatalytic Dye Degradation under Natural Sunlight Irradiation. J. Compos. Sci. 2019, 3, 75. [Google Scholar] [CrossRef] [Green Version]
  7. Gebreslassie, G.; Bharali, P.; Gebremariam, G.; Sergawie, A.; Alemayehu, E. Graphitic Carbon Nitride with Extraordinary Photocatalytic Activity under Visible Light Irradiation; Springer International Publishing: Cham, Switzerland, 2021; Volume 385. [Google Scholar] [CrossRef]
  8. Cheng, L.; Zhang, H.; Li, X.; Fan, J.; Xiang, Q. Carbon—Graphitic Carbon Nitride Hybrids for Heterogeneous Photocatalysis. Small 2020, 17, 2005231. [Google Scholar] [CrossRef] [PubMed]
  9. Wen, J.; Xie, J.; Chen, X.; Li, X. A review on g-C3N4-based photocatalysts. Appl. Surf. Sci. 2017, 391, 72–123. [Google Scholar] [CrossRef]
  10. Zhang, M.; Yang, Y.; An, X.; Hou, L. A critical review of g-C3N4-based photocatalytic membrane for water purification. Chem. Eng. J. 2021, 412, 128663. [Google Scholar] [CrossRef]
  11. Mishra, A.; Mehta, A.; Basu, S.; Shetti, N.P.; Reddy, K.R.; Aminabhavi, T.M. Graphitic carbon nitride (g-C3N4)–based metal-free photocatalysts for water splitting: A review. Carbon 2019, 149, 693–721. [Google Scholar] [CrossRef]
  12. Jiang, L.; Yuan, X.; Pan, Y.; Liang, J.; Zeng, G. Doping of graphitic carbon nitride for photocatalysis: A reveiw. Appl. Catal. B Environ. 2017, 217, 388–406. [Google Scholar] [CrossRef]
  13. Chebanenko, M.I.; Tikhanova, S.M.; Nevedomskiy, V.N.; Popkov, V.I. Synthesis and Structure of ZnO-Decorated Graphitic Carbon Nitride (g-C3N4) with Improved Photocatalytic Activity under Visible Light. Inorganics 2022, 10, 249. [Google Scholar] [CrossRef]
  14. Farooq, N.; Luque, R.; Hessien, M.M.; Qureshi, A.M.; Sahiba, F.; Nazir, M.A.; ur Rehman, A. A comparative study of cerium- and ytterbium-based GO/g-C3N4/Fe2O3 composites for electrochemical and photocatalytic applications. Appl. Sci. 2021, 11, 9000. [Google Scholar] [CrossRef]
  15. Ismael, M.; Wark, M. Photocatalytic activity of CoFe2O4/g-C3N4 nanocomposite toward degradation of different organic pollutants and their inactivity toward hydrogen production: The role of the conduction band position. FlatChem 2022, 32, 100337. [Google Scholar] [CrossRef]
  16. Paul, A.; Dhar, S.S. Designing Cu2V2O7/CoFe2O4/g-C3N4 ternary nanocomposite: A high performance magnetically recyclable photocatalyst in the reduction of 4-nitrophenol to 4-aminophenol. J. Solid State Chem. 2020, 290, 121563. [Google Scholar] [CrossRef]
  17. He, W.; Liu, L.; Ma, T.; Han, H.; Zhu, J.; Liu, Y.; Fang, Z.; Yang, Z.; Guo, K. Controllable morphology CoFe2O4/g-C3N4 p-n heterojunction photocatalysts with built-in electric field enhance photocatalytic performance. Appl. Catal. B Environ. 2022, 306, 121107. [Google Scholar] [CrossRef]
  18. Zhang, J.; Chen, X.; Takanabe, K.; Maeda, K.; Domen, K.; Epping, J.D.; Fu, X.; Antonietti, M.; Wang, X. Synthesis of a Carbon Nitride Structure for Visible-Light Catalysis by Copolymerization. Angew. Chem. 2010, 49, 441–444. [Google Scholar] [CrossRef]
  19. Gebreslassie, G.; Bharali, P.; Chandra, U.; Sergawie, A.; Boruah, P.K.; Das, M.R.; Alemayehu, E. Novel g-C3N4/graphene/NiFe2O4 nanocomposites as magnetically separable visible light driven photocatalysts. J. Photochem. Photobiol. A Chem. 2019, 382, 111960. [Google Scholar] [CrossRef]
  20. Li, Y.; Zhou, M.; Cheng, B.; Shao, Y. Recent advances in g-C3N4-based heterojunction photocatalysts. J. Mater. Sci. Technol. 2020, 56, 1–17. [Google Scholar] [CrossRef]
  21. Kumar, O.P.; Shahzad, K.; Nazir, M.A.; Farooq, N.; Malik, M.; Ahmad Shah, S.S.; ur Rehman, A. Photo-Fenton activated C3N4x/AgOy@Co1-xBi0.1-yO7 dual s-scheme heterojunction towards degradation of organic pollutants. Opt. Mater. 2022, 126, 112199. [Google Scholar] [CrossRef]
  22. Wang, Q.; Li, Y.; Huang, F.; Song, S.; Ai, G.; Xin, X.; Zhao, B.; Zheng, Y.; Zhang, Z. Recent Advances in g-C3N4-Based Materials and Their Application in Energy and Environmental Sustainability. Molecules 2023, 28, 432. [Google Scholar] [CrossRef]
  23. Gupta, N.K.; Gha, Y.; Kim, S.; Bae, J.; Kim, K.S. Photocatalytic Degradation of Organic Pollutants over MFe2O4 (M = Co, Ni, Cu, Zn) Nanoparticles at Neutral pH. Sci. Rep. 2020, 10, 4942. [Google Scholar] [CrossRef] [Green Version]
  24. Nife, O.; Holinsworth, B.S.; Mazumdar, D.; Sims, H.; Sun, Q.; Yurtisigi, M.K.; Sarker, S.K.; Gupta, A.; Butler, W.H.; Musfeldt, J.L. Chemical tuning of the optical band gap in spinel ferrites: CoFe2O4 vs. NiFe2O4. Appl. Phys. Lett. 2013, 103, 082406. [Google Scholar]
  25. He, H.; Lu, J. Highly photocatalytic activities of magnetically separable reduced graphene oxide-CoFe 2 O 4 hybrid nanostructures in dye photodegradation. Sep. Purif. Technol. 2017, 172, 374–381. [Google Scholar] [CrossRef]
  26. Fu, Y.; Chen, H.; Sun, X.; Wang, X. Combination of cobalt ferrite and graphene: High-performance and recyclable visible-light photocatalysis. Appl. Catal. B Environ. 2012, 111, 280–287. [Google Scholar] [CrossRef]
  27. Huang, S.; Xu, Y.; Xie, M.; Xu, H.; He, M.; Xia, J.; Huang, L.; Li, H. Synthesis of magnetic CoFe2O4/g-C3N4 composite and its enhancement of photocatalytic ability under visible-light. Colloids Surf. A Physicochem. Eng. Asp. 2015, 478, 71–80. [Google Scholar] [CrossRef]
  28. Inbaraj, D.J.; Chandran, B.; Mangalaraj, C. Synthesis of CoFe2O4 and CoFe2O4/g-C3N4 nanocomposite via honey mediated sol-gel auto combustion method and hydrothermal method with enhanced photocatalytic and efficient Pb+2 adsorption property. Mater. Res. Express 2019, 6, 055501. [Google Scholar] [CrossRef]
  29. Yuan, C.; Cao, H.; Zhu, S.; Hua, H.; Hou, L. Core-shell ZnO/ZnFe2O4@C mesoporous nanospheres with enhanced lithium storage properties towards high-performance Li-ion batteries. J. Mater. Chem. A 2015, 3, 20389–20398. [Google Scholar] [CrossRef]
  30. Hou, L.; Lian, L.; Zhang, L.; Pang, G.; Yuan, C.; Zhang, X. Self-sacrifice template fabrication of hierarchical mesoporous bi-component-active ZnO/ZnFe2O4 sub-microcubes as superior anode towards high-performance lithium-ion battery. Adv. Funct. Mater. 2015, 25, 238–246. [Google Scholar] [CrossRef]
  31. Dong, F.; Wu, L.; Sun, Y.; Fu, M.; Wu, Z.; Lee, S.C. Efficient synthesis of polymeric g-C 3 N 4 layered materials as novel efficient visible light driven photocatalysts. J. Mater. Chem. 2011, 21, 15171–15174. [Google Scholar] [CrossRef]
  32. Liu, J.; Zhang, T.; Wang, Z.; Dawson, G.; Chen, W. Simple pyrolysis of urea into graphitic carbon nitride with recyclable adsorption and photocatalytic activity. J. Mater. Chem. 2011, 21, 14398–14401. [Google Scholar] [CrossRef]
  33. Zhang, S.; Li, J.; Zeng, M.; Zhao, G.; Xu, J.; Hu, W.; Wang, X. In Situ Synthesis of Water-Soluble Magnetic Graphitic Carbon Nitride Photocatalyst and Its Synergistic Catalytic Performance. ACS Appl. Mater. Interfaces 2013, 5, 12735–12743. [Google Scholar] [CrossRef]
  34. Anand, S.; Amaliya, A.P.; Janifer, M.A.; Pauline, S. Structural, morphological and dielectric studies of zirconium substituted CoFe2O4 nanoparticles. Mod. Electron. Mater. 2017, 3, 168–173. [Google Scholar] [CrossRef]
  35. Sabale, S.R. Studies on catalytic activity of MnFe2O4 and CoFe2O4MNPsas mediators in hemoglobin based biosensor. Mater. Today Proc. 2020, 23, 139–146. [Google Scholar] [CrossRef]
  36. Khan, M.; Pawar, H.; Kumari, M.; Patra, C.; Patel, G.; Dwivedi, U.K.; Rathore, D. Effect of concentration of SiC on physicochemical properties of CoFe2O4/SiC nanocomposites. J. Alloys Compd. 2020, 840, 155596. [Google Scholar] [CrossRef]
  37. Kumar, O.P.; Ahmad, M.; Nazir, M.A.; Anum, A.; Jamshaid, M.; Shah, S.S.A.; Rehman, A. Strategic combination of metal–organic frameworks and C3N4 for expeditious photocatalytic degradation of dye pollutants. Environ. Sci. Pollut. Res. 2022, 29, 35300–35313. [Google Scholar] [CrossRef] [PubMed]
  38. Renukadevi, S.; Jeyakumari, A.P. A one-pot microwave irradiation route to synthesis of CoFe2O4-g-C3N4 heterojunction catalysts for high visible light photocatalytic activity: Exploration of efficiency and stability. Diam. Relat. Mater. 2020, 109, 108012. [Google Scholar] [CrossRef]
  39. Hassani, A.; Çelikdağ, G.; Eghbali, P.; Sevim, M.; Karaca, S.; Metin, Ö. Heterogeneous sono-Fenton-like process using magnetic cobalt ferrite-reduced graphene oxide (CoFe2O4-rGO) nanocomposite for the removal of organic dyes from aqueous solution. Ultrason. Sonochem. 2018, 40, 841–852. [Google Scholar] [CrossRef] [PubMed]
  40. Gebreslassie, G.; Bharali, P.; Chandra, U.; Sergawie, A.; Baruah, P.K.; Das, M.R.; Alemayehu, E. Hydrothermal Synthesis of g-C3N4/NiFe2O4 Nanocomposite and Its Enhanced Photocatalytic Activity. Appl. Organomet. Chem. 2019, 33, e5002. [Google Scholar] [CrossRef]
  41. Hafeez, H.Y.; Lakhera, S.K.; Narayanan, N.; Harish, S.; Hayakawa, Y.; Lee, B.K.; Neppolian, B. Environmentally Sustainable Synthesis of a CoFe2O4-TiO2/rGO Ternary Photocatalyst: A Highly Efficient and Stable Photocatalyst for High Production of Hydrogen (Solar Fuel). ACS Omega 2019, 4, 880–891. [Google Scholar] [CrossRef] [Green Version]
  42. Xu, Q.; Zhang, L.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Jaroniec, M. Direct Z-scheme photocatalysts: Principles, synthesis, and applications. Mater. Today 2018, 21, 1042–1063. [Google Scholar] [CrossRef]
  43. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction Photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef] [PubMed]
Figure 1. FTIR spectra of (a) g-C3N4, (b) CoFe2O4, (c) CoFe2O4-loaded g-C3N4 photocatalysts.
Figure 1. FTIR spectra of (a) g-C3N4, (b) CoFe2O4, (c) CoFe2O4-loaded g-C3N4 photocatalysts.
Inorganics 11 00119 g001
Figure 2. XRD patterns of g-C3N4, CoFe2O4, and CoFe2O4-loaded g-C3N4 photocatalyst.
Figure 2. XRD patterns of g-C3N4, CoFe2O4, and CoFe2O4-loaded g-C3N4 photocatalyst.
Inorganics 11 00119 g002
Figure 3. SEM images of (a) pure CoFe2O4, (b) pure g-C3N4, and (c) CoFe2O4-loaded g-C3N4 photocatalyst, and EDX pattern of (d) pure CoFe2O4, (e) pure g-C3N4, and (f) CoFe2O4-loaded g-C3N4 photocatalyst.
Figure 3. SEM images of (a) pure CoFe2O4, (b) pure g-C3N4, and (c) CoFe2O4-loaded g-C3N4 photocatalyst, and EDX pattern of (d) pure CoFe2O4, (e) pure g-C3N4, and (f) CoFe2O4-loaded g-C3N4 photocatalyst.
Inorganics 11 00119 g003
Figure 4. The magnetic hysteresis loops of pure CoFe2O4 and CoFe2O4-loaded g-C3N4 photocatalyst material.
Figure 4. The magnetic hysteresis loops of pure CoFe2O4 and CoFe2O4-loaded g-C3N4 photocatalyst material.
Inorganics 11 00119 g004
Figure 5. (a) N2 adsorption–desorption isotherm of CoFe2O4-loaded g-C3N4 photocatalyst, and (b) the pore size distribution curve of CoFe2O4-loaded g-C3N4 photocatalyst.
Figure 5. (a) N2 adsorption–desorption isotherm of CoFe2O4-loaded g-C3N4 photocatalyst, and (b) the pore size distribution curve of CoFe2O4-loaded g-C3N4 photocatalyst.
Inorganics 11 00119 g005
Figure 6. (a) DRS spectra of g-C3N4, CoFe2O4, and CoFe2O4-loaded g-C3N4 photocatalyst. (b) Tauc plot for g-C3N4 photocatalyst. (c) Tauc plot for CoFe2O4 and CoFe2O4-loaded g-C3N4 photocatalyst. (d) PL spectra of g-C3N4, and CoFe2O4-loaded g-C3N4 photocatalyst.
Figure 6. (a) DRS spectra of g-C3N4, CoFe2O4, and CoFe2O4-loaded g-C3N4 photocatalyst. (b) Tauc plot for g-C3N4 photocatalyst. (c) Tauc plot for CoFe2O4 and CoFe2O4-loaded g-C3N4 photocatalyst. (d) PL spectra of g-C3N4, and CoFe2O4-loaded g-C3N4 photocatalyst.
Inorganics 11 00119 g006
Figure 7. (a) Photodegradation of MB by CoFe2O4, g-C3N4, and CoFe2O4-loaded g-C3N4 photocatalysts under visible light irradiation. (b) Photodegradation rate constant (kap) of MB over the CoFe2O4, g-C3N4, and CoFe2O4-loaded g-C3N4 photocatalysts.
Figure 7. (a) Photodegradation of MB by CoFe2O4, g-C3N4, and CoFe2O4-loaded g-C3N4 photocatalysts under visible light irradiation. (b) Photodegradation rate constant (kap) of MB over the CoFe2O4, g-C3N4, and CoFe2O4-loaded g-C3N4 photocatalysts.
Inorganics 11 00119 g007
Figure 8. Reusability of the CoFe2O4-loaded g-C3N4 photocatalyst after five successive runs.
Figure 8. Reusability of the CoFe2O4-loaded g-C3N4 photocatalyst after five successive runs.
Inorganics 11 00119 g008
Figure 9. Scavenger effects on the performance of MB degradation over CoFe2O4-loaded g-C3N4 photocatalyst.
Figure 9. Scavenger effects on the performance of MB degradation over CoFe2O4-loaded g-C3N4 photocatalyst.
Inorganics 11 00119 g009
Figure 10. The CoFe2O4-loaded g-C3N4 under irradiation of visible-light and the proposed direct Z-scheme photocatalytic process.
Figure 10. The CoFe2O4-loaded g-C3N4 under irradiation of visible-light and the proposed direct Z-scheme photocatalytic process.
Inorganics 11 00119 g010
Table 1. Comparison of the effectiveness of several magnetic photocatalysts in the degradation of organic dyes.
Table 1. Comparison of the effectiveness of several magnetic photocatalysts in the degradation of organic dyes.
No.PhotocatalystLight SourceTarget DyeDegradation EfficiencyRef.
1CoFe2O4/g-C3N4Sun lightMB98%/150 min[28]
241.4% CoFe2O4/g-C3N4-H2O2Xenon lampMB97.3%/180 min[27]
3CoFe2O4/g-C3N4-PMSHalogen tungsten lamp RhB96%/30 min[1]
4Direct Z-scheme CoFe2O4-loaded g-C3N4LED lampMB98.86%/140 minThis work
Table 2. Electronegativity (χ), band gap (Eg), and conduction band (ECB) position and valance band (EVB) of the photocatalysts on NHE.
Table 2. Electronegativity (χ), band gap (Eg), and conduction band (ECB) position and valance band (EVB) of the photocatalysts on NHE.
Photocatalystχ (eV)Eg (eV)ECB (eV)EVB (eV)
CoFe2O45.811.30+0.66+1.96
g-C3N44.732.65−1.09+1.56
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

Gebreslassie, G.; Gebrezgiabher, M.; Lin, B.; Thomas, M.; Linert, W. Direct Z-Scheme CoFe2O4-Loaded g-C3N4 Photocatalyst with High Degradation Efficiency of Methylene Blue under Visible-Light Irradiation. Inorganics 2023, 11, 119. https://doi.org/10.3390/inorganics11030119

AMA Style

Gebreslassie G, Gebrezgiabher M, Lin B, Thomas M, Linert W. Direct Z-Scheme CoFe2O4-Loaded g-C3N4 Photocatalyst with High Degradation Efficiency of Methylene Blue under Visible-Light Irradiation. Inorganics. 2023; 11(3):119. https://doi.org/10.3390/inorganics11030119

Chicago/Turabian Style

Gebreslassie, Gebrehiwot, Mamo Gebrezgiabher, Bin Lin, Madhu Thomas, and Wolfgang Linert. 2023. "Direct Z-Scheme CoFe2O4-Loaded g-C3N4 Photocatalyst with High Degradation Efficiency of Methylene Blue under Visible-Light Irradiation" Inorganics 11, no. 3: 119. https://doi.org/10.3390/inorganics11030119

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