1. Introduction
Spinel-structured photocatalysts are gaining attention for their cost-effectiveness, durability, and strong photoelectrochemical response, making them prime for enhancing solar energy capture [
1,
2,
3]. Various AB
2O
4 spinels, such as ZnFe
2O
4 and ZnCo
2O
4, have been explored for applications ranging from gas sensing to energy storage and degradation of pollutants under visible light [
4,
5,
6]. ZnCo
2O
4 is classified as a p-type semiconductor due to its spinel crystal structure and is notable for its versatility, including roles in Li-ion batteries, catalysis, and supercapacitors [
7,
8]. ZnCo
2O
4-based sensors have shown exceptional sensitivity to various gases, likely due to their high surface area [
9]. The nanoparticles' morphology influences gas sensing performance, indicating a significant research interest in optimizing ZnCo
2O
4 sensors for higher sensitivity and lower operating temperatures [
10,
11]. Diverse synthesis methods like hydrothermal and microwave-assisted heating techniques have been developed, offering efficiency and cost-effectiveness for oxide material structures [
12,
13,
14,
15]. ZnCo
2O
4's effective use in breaking down organic pollutants showcases its potential for environmental cleanup [
16,
17,
18]. The evolution from bulk to porous structures, offering more active sites and better light absorption, marks a significant advancement [
19]. The annealing-based self-sacrificial templating emerges as a cost-effective, high-performance method for fabricating these porous photocatalysts, emphasizing the importance of precursor selection in achieving superior product quality [
16,
20,
21]. Previous studies have demonstrated the successful preparation of ZnCo
2O
4 nanostructures on carbon cloth for electrodes for supercapacitors and lithium-ion batteries [
22,
23,
24]. However, their applications on carbon cloth for photocatalytic degradation remain infrequent.
Carbon cloth, a carbon filament textile, offers high conductivity, mechanical strength, and flexibility, making it ideal for flexible energy storage systems [
25,
26,
27]. Despite a low surface area and few electroactive sites, carbon cloth is a flexible substrate for electrode materials [
28,
29,
30]. Notably, Wang et al.'s oxidative method with hydrogen peroxide and sulfuric acid introduces oxygen-containing groups onto carbon cloth, enhancing its function [
31]. Kordek et al., Liu et al., and Zhao et al. employed various complex activation methods to enhance the oxygen electrocatalytic activities of carbon cloth, each achieving improved performance through surface modification techniques, such as etching, calcining, plasma treatment, and doping with heteroatoms [
32,
33]. Therefore, surface modification technology can further improve carbon cloth's specific surface area and active sites, improving its application in various fields. Past research has shown that combining carbon cloth and ZnCo
2O
4 nanostructures creates a high-capacity, flexible anode with excellent cycle stability and rate performance, thus forming a highly flexible lithium-ion battery with excellent electrochemical properties [
23,
34]. In this study, the combination of carbon cloth and ZnCo
2O
4-Zn(OH)
2 microspheres is expected to be further used in the photocatalytic degradation of methyl violet and tetracycline, thereby simplifying the subsequent recycling process.
ZnCo2O4-Zn(OH)2 microspheres fabricated on carbon cloth via a simplified hydrothermal process exhibited enhanced photocatalytic degradation of pollutants like methyl violet, thanks to optimized porosity from varied thermal annealing temperatures. Extensive characterization revealed that annealed microspheres offer superior surface area, light absorption, and charge carrier separation, making them highly effective under UVC light. These microspheres proved efficient and reusable photocatalysts, with superoxide and hydroxyl radicals identified as the major reactive species in pollutant degradation.
2. Results and Discussion
Figure 1 is a detailed schematic diagram illustrating the growing ZnCo
2O
4-Zn(OH)
2 microspheres on a carbon cloth substrate. Initially, carbon cloth undergoes a meticulous etching process using a mixture of sulfuric acid and hydrogen peroxide. This critical step creates a series of micropores on the surface, enhancing its texture and providing anchoring points that facilitate the subsequent nucleation and growth of ZnCo
2O
4-Zn(OH)
2 microspheres. After surface treatment, the carbon cloth was subjected to hydrothermal treatment, during which ZnCo
2O
4-Zn(OH)
2 microspheres began to form and adhere to the etched surface. The reaction process was performed at a controlled temperature of 120 °C and maintained for 2 h, ensuring uniform growth of microspheres. This carefully planned process resulted in the successful integration of ZnCo
2O
4-Zn(OH)
2 microspheres onto carbon cloth, laying the foundation for its potential application in photocatalytic degradation due to enhanced surface properties and synergistic effects between carbon cloth and ZnCo
2O
4-Zn(OH)
2 microspheres.
Figure 2a and b show FESEM images of carbon cloth before and after soaking in a solution containing hydrogen peroxide and sulfuric acid. Before shaking, the carbon cloth surface is quite smooth. However, after soaking, the surface of the carbon cloth displays holes of various sizes. Subsequently, the two types of carbon are placed in a ZnCo
2O
4-Zn(OH)
2 reaction precursor at a reaction temperature of 120 °C for 2 h by a facile hydrothermal process, as shown in
Figure 2 c and d. From these images, it can be observed that after soaking in the sulfuric acid and hydrogen peroxide mixed solution, the ZnCo
2O
4-Zn(OH)
2 microspheres that grow on the surface of the carbon cloth have a higher density. This result proves that soaking the carbon cloth in a solution containing hydrogen peroxide and sulfuric acid creates more pores on its surface and aids in the subsequent growth of ZnCo
2O
4-Zn(OH)
2 microspheres.
Additionally, high-resolution FESEM images (as seen in
Figure 3a) depict the presence of ZnCo
2O
4-Zn(OH)
2 microspheres. The distribution of specific elements within these microspheres is further elucidated via FESEM-EDS elemental mapping, as showcased in
Figure 3b-d. Examination of these images reveals that the microspheres are composed of zinc (Zn), cobalt (Co), and oxygen (O), with these components being uniformly distributed throughout. Consequently, this highlights the specific elemental composition of the ZnCo
2O
4-Zn(OH)
2 microspheres.
Figure 4 shows scanning electron microscope images of ZnCo
2O
4-Zn(OH)
2 microspheres subjected to different thermal annealing temperatures for 2h. The thermal annealing temperatures are (a) without, (b) 450 °C, (c) 550 °C, and (d) 650 °C, respectively. As the annealing temperature rises, there is a progressive emergence of porosity on the surface. The surface morphology of the ZnCo
2O
4-Zn(OH)
2 microspheres remains unchanged compared to the unannealed samples, with the only noticeable difference being the appearance of porosity at an annealing temperature of 550°C for 2 h. When the annealing temperature rises to 650°C, it becomes apparent that some microspheres have collapsed. This occurrence drastically decreases the reactive surface area, negatively impacting the efficiency of subsequent photocatalytic reactions. The structural collapse observed at higher temperatures underscores the material's thermal instability, highlighting the necessity for meticulous optimization of the thermal annealing process to preserve the desired functional properties of the microspheres. Hence, the BET analyzer can assess the specific surface area of ZnCo
2O
4-Zn(OH)
2 microspheres before and after thermal annealing at 550 °C for 2 h. The surface area of the microspheres was measured at 20.78 m
2g
−1 before thermal annealing, and following the annealing process, this value rose to 31.29 m
2g
−1. This result indicates that thermal annealing effectively enhances the specific surface area of the microspheres.
X-ray diffraction (XRD) was employed for the examination of the crystalline structure of the ZnCo
2O
4-Zn(OH)
2 microspheres (without thermal annealing) and ZnCo
2O
4-Zn(OH)
2 microspheres (thermal annealing at 550 °C for 2h), as shown in
Figure 5. The diffraction angles of 29.6°, 35.5°, 43.3°, 47.5°, 57.2°, 58.0°, 60.9°, and 64.5° can be observed in Zn(OH)
2, corresponding to the (031), (211), (350), (181), (0120), (2111), (510), (152) and (1141) planes of orthorhombic Zn(OH)
2 (PDF No. 00-020-1437), respectively. The diffraction angles of 31.2°, 36.8°, 48.9°, and 65.2° can be observed in ZnCo
2O
4, corresponding to the (220), (311), (331), and (440) planes of cubic ZnCo
2O
4 (PDF No. 00-023-1390), respectively. It is evident that the peak intensity of ZnCo
2O
4 in ZnCo
2O
4-Zn(OH)
2 microspheres is weakened, indicating that the content of ZnCo
2O
4 is lower and the crystallinity is not as strong as that of Zn(OH)
2 when Zn(OH)
2 and ZnCo
2O
4 are combined. This result confirms the successful generation of ZnCo
2O
4-Zn(OH)
2 microspheres.
The FETEM image in
Figure 6a reveals a microspherical structure of ZnCo
2O
4-Zn(OH)
2, which aligns with the SEM results. This configuration is characterized by the arrangement of numerous sheets stacked on each other. The typical SAED pattern (
Figure 6b) further confirms the polycrystalline nature of the ZnCo
2O
4-Zn(OH)
2 microsphere. The major diffraction ring closely matches the orthorhombic Zn(OH)
2 (PDF No. 00-020-1437) and cubic ZnCo
2O
4 (PDF No. 00-023-1390) crystal structures. The HRTEM image of the ZnCo
2O
4-Zn(OH)
2 microsphere (
Figure 6c) displays crystal lattice fringes characterized by two discernible interplanar spacings: 0.302 nm and 0.244 nm. These can be attributed to the (031) crystallographic plane of the orthorhombic phase of Zn(OH)
2 and the (311) crystallographic plane of the cubic phase of ZnCo
2O
4.
Figure 6d reveals the corresponding elemental mapping images of the ZnCo
2O
4-Zn(OH)
2 microsphere, showing the distribution of Zn, Co, and O elements. This result indicates that Zn, Co, and O define the ZnCo
2O
4-Zn(OH)
2 microsphere composition.
To examine the elemental composition and valence distribution of the surface of ZnCo
2O
4-Zn(OH)
2 microspheres, an analysis utilizing X-ray photoelectron spectroscopy (XPS) was conducted.
Figure 7a displays the full range XPS spectrum of ZnCo
2O
4-Zn(OH)
2 microspheres annealed at 550 °C, exhibiting distinct peaks corresponding to C, Zn, Co, and O. These peaks are consistent with the TEM-EDS observations, further confirming the presence of these elements. The carbon element is believed to have its source in the pump oil present in the vacuum system of the XPS equipment, carbon cloth, or an organic layer that has been applied to the surface of the sample. The high-resolution XPS spectrum (
Figure 7b) of Zn 2p, showing peaks at 1020.9 eV and 1044.1 eV for Zn 2p
3/2 and Zn 2p
1/2, respectively, confirms the presence of Zn
2+ in the ZnCo
2O
4-Zn(OH)
2 structure [
35,
36].
Figure 7c reveals the high-resolution XPS spectrum of Co 2p of ZnCo
2O
4, which can be deconvolved into four different states: Co
3+ at 779.5 eV (2p
3/2) and 794.5 eV (2p
1/2), and Co
2+ at 780.6 eV (2p
3/2) and 795.7 eV (2p
1/2). Additionally, two vibrational satellite peaks for Co
2+ are located at 789.6 eV near the Co 2p
3/2 band and 804.8 eV near the Co 2p
1/2 band, consistent with previous literature [
37,
38]. The O 1s spectrum of the synthesized ZnCo
2O
4-Zn(OH)
2 microspheres (
Figure 7d) reveals a primary peak at 529.5 eV corresponding to lattice oxygen (O
L), along with shoulder peaks at 530.8 eV and 532.1 eV attributed to surface hydroxyl groups (O
OH) and chemisorbed oxygen (O
C) [
37,
38].
To comprehend the correlation between the photocatalytic efficiency of ZnCo
2O
4-Zn(OH)
2 microspheres under different annealing temperatures, the photocatalytic activity in degrading methyl violet (MV), an organic pollutant commonly found in the textile industry was evaluated. ZnCo
2O
4-Zn(OH)
2 microspheres were grown on a 2.5 cm × 1.5 cm carbon cloth substrate as photocatalytic samples. The photocatalytic efficiency of ZnCo
2O
4-Zn(OH)
2 microspheres at different annealing temperatures was evaluated by the degradation of MV by UVC light (253 nm, 10 W), as shown in
Figure 8a. The time variation of MV concentration was monitored by examining the change in maximum absorbance at 587 nm in UV-vis spectroscopy. The photodegradation percentages of MV were 58.0% (without annealing), 89.8% (450 °C), 91.7% (550 °C), and 78.6% (650 °C). When the annealing temperature was below 550 °C, a decrease in maximum absorbance was observed with increasing irradiation time and annealing temperature. ZnCo
2O
4-Zn(OH)
2 microspheres (550 °C) exhibited the highest photocatalytic activity in MV decomposition. The photocatalytic degradation process conformed to pseudo-first-order kinetics, and the plot of −ln(C/C
0) versus irradiation time (t) showed a pseudo-first-order linear relationship (
Figure 8b), where C
0 is the initial concentration of MV and C is the actual concentration of MV at time t. The slope of the pseudo-first-order linear line is the apparent rate constant (k, min
–1) of the photocatalytic reaction. The rate constants of ZnCo
2O
4-Zn(OH)
2 microspheres at different annealing temperatures were calculated to be 0.03298 (without annealing), 0.05002 (450 °C), 0.08519 (550 °C), and 0.07241 (650 °C), respectively, in min
–1. ZnCo
2O
4-Zn(OH)
2 microspheres (550 °C) displayed the highest photocatalytic efficiency in MV photodegradation under UVC light irradiation. The rate constant (k) of ZnCo
2O
4-Zn(OH)
2 microspheres (550 °C) was about 2.58 times higher than that of the non-annealed ones. This phenomenon is attributed to the formation of porosity on the surface of ZnCo
2O
4-Zn(OH)
2 microspheres after annealing, which increases the active sites. However, when the annealing temperature is too high, it can cause the structure to collapse, leading to a decrease in active sites, which is consistent with the observations from SEM results.
We chose tetracycline (TC) as an antibiotic to illustrate that ZnCo
2O
4-Zn(OH)
2 microspheres can also be used for photocatalytic antibiotic degradation. Tetracycline (TC), a widely used antibiotic effective against various infections, is prevalent in water bodies due to its use as a growth promoter in aquaculture and insufficient removal by traditional wastewater treatments [
39,
40]. As demonstrated in
Figure 9a, we observe the degradation rates of ZnCo
2O
4-Zn(OH)
2 microspheres before and after annealing when subjected to UVC light. The findings indicated that the microspheres that did not go through the annealing process had a degradation rate of 75.6%, whereas those that were annealed at a temperature of 550 °C exhibited a rate of 83.3%.
Figure 9b represents the pseudo-first-order linear relationship of the ZnCo
2O
4-Zn(OH)
2 microspheres in the pre and annealing process. The reaction constants, corresponding to TC degradation over the non-annealed microspheres and those annealed at 550 °C, were determined to be 0.00794 and 0.00986 min
−1, respectively. Notably, the microspheres that underwent annealing at 550 °C showed superior photocatalytic activity, with their reaction constant being 1.24 times greater than their non-annealed counterparts when exposed to UVC light. This suggests an enhancement in photocatalytic efficiency due to the annealing process.
The recyclability of ZnCo
2O
4-Zn(OH)
2 microspheres (550 °C) was examined through repeated experiments involving the degradation of MV and TC solutions under UVC light irradiation, as shown in
Figure 10. In the case of the MV solution (
Figure 10a), the photocatalytic efficiency remained consistently high across four cycles, with efficiencies of 91.7%, 90.5%, 89.9%, and 87.4%, respectively. Similarly, the photocatalytic efficiency maintained a steady rate for the TC solution (
Figure 10b), with efficiencies of 82.8%, 82.4%, 81.7%, and 80.9% across the four cycles. Even after four cycles of use, the decline in the photocatalytic efficiency of the ZnCo
2O
4-Zn(OH)
2microspheres was insignificant, demonstrating their durability and consistent performance. This result suggests that the ZnCo
2O
4-Zn(OH)
2 microspheres, heated at 550 °C, possess a long lifespan as photocatalysts, maintaining high activity and reusability. The ZnCo
2O
4-Zn(OH)
2 microspheres were also directly cultivated on a carbon cloth. In the meantime, the tested ZnCo
2O
4-Zn(OH)
2 microspheres (
Figure 11) displayed remarkable consistency in the XRD patterns before and after degradation tests, validating their high resistance to photo-corrosion and stability. This result suggests that ZnCo
2O
4-Zn(OH)
2 microspheres have the promising potential for repeated use in practical applications, a highly desirable characteristic for sustainable and efficient photocatalysts. This resilience ensures their longevity and enhances their cost-effectiveness, making them a compelling choice for environmental applications. This unique growth method simplifies recycling and provides a stable and economical photocatalyst platform.
The optical properties of the ZnCo
2O
4-Zn(OH)
2 microspheres, both with and without thermal annealing, were examined using UV-visible spectroscopy. As shown in
Figure 11a, the ZnCo
2O
4-Zn(OH)
2 microspheres that underwent thermal annealing at 550 °C demonstrated a significantly superior light absorption capacity within the spectral range spanning from 250 to 800 nm compared to those that did not undergo annealing. The improved light absorption spectrum observed in the annealed microspheres offers advantages for maximizing solar energy utilization, thereby enhancing the photocatalytic degradation process. The energy band gaps (E
g) were established utilizing the Tauc relationship, represented by the equation: αhν = A(E
g − hν)
1/n [
41,
42]. In the provided equation, A, α, ν, E
g, and h represent constants: the constant, the absorption coefficient, the frequency of light, the band gap energy, and Planck’s constant, respectively. The variable "n" denotes a property of the semiconductor material, taking a value of 2 for indirect bandgap semiconductors and 1/2 for direct bandgap semiconductors, as illustrated in
Figure 11b. The energy band gap value of the ZnCo
2O
4-Zn(OH)
2 microspheres, both pre and post-annealing, was computed to be approximately 2.42 eV. This data validates that the energy gap remains relatively constant throughout the thermal annealing process.
Moreover, a straightforward thermal annealing technique can offer a high specific surface area and a more extensive optical absorption spectrum at a suitable temperature. This method significantly enhances the photocatalytic degradation of MV or TC solutions, improving their overall performance and effectiveness.
Four radical scavengers were introduced into the photocatalytic reaction to investigate the underlying mechanism of the photocatalysts of ZnCo
2O
4-Zn(OH)
2 microspheres during the photodegradation of MV or TC solution, as shown in
Figure 13a and b. Isopropyl alcohol (IPA), L-ascorbic acid (AA), triethanolamine (TEOA), and silver nitrate (AgNO
3) were utilized as scavengers to impede hydroxyl radicals (·OH), superoxide radical anions (·O
2–), holes (h
+), and electrons (e
–), respectively [
43,
44,
45,
46]. Adding IPA and AA scavengers to the photocatalytic reaction leads to a notable reduction in the photocatalytic efficiency. This outcome provides evidence that hydroxyl radicals (·OH) and superoxide radicals (·O
2−) are the primary active species involved in the photodegradation of TC. Potential reactions that may occur during the photocatalytic degradation of MV or TC solutions over ZnCo
2O
4-Zn(OH)
2 microspheres can be outlined as a schematic diagram, as shown in
Figure 13c. When the ZnCo
2O
4-Zn(OH)
2 microspheres are exposed to UVC light with photon energy (hv) exceeding their band gap, an electron (e
–) in the valence band (VB) can be promoted to the conduction band (CB), creating a hole in the VB and generating electron-hole pairs. Photogenerated electrons can interact with oxygen molecules on the surface, forming superoxide radical anions (·O
2–). These can then react with water molecules absorbed on the surface, producing hydroxyl radicals (·OH). Moreover, the photogenerated holes may combine with H
2O molecules, causing their dissociation into ·OH radicals. These superoxide radical anions and hydroxyl radicals are recognized as potent oxidants responsible for the decomposition of MV or TC molecules.