1. Introduction
Environmental contamination and energy demand represent a worldwide concern. To overcome these problems, nanomaterials are an alternative to address environmental aspects of wastewater treatment contamination and energy generation by organic compounds [
1,
2,
3]. The immoderate emptying of waste toxic organic dyes, such as methyl orange (MO), into soils and water pools counteracts environmental protection and safety. Consequently, it is necessary to treat the dyeing wastewater with practical and harmless methods [
4,
5]. It is well-known that solar light is a favorite choice of clean energy for decomposing pollutants due to its great redox activity and energy density only makes H
2O [
6]. The utilization of visible light is a rather effective process as it adopts abundantly available solar light. To date, the photocatalytic system, as a green and direct reach, has been found to be effective in environmental treatment, especially in the removal of organic pollutants. Photocatalytic degradation with visible light activity is employed to excite the semiconductor material creating electron (
e−) and hole (
h+) pairs for the degradation of organic compounds. To date, different semiconductors have been adopted as photocatalysts in environmental applications [
7]. Photocatalytic removal pollutants have attracted sufficient exploration attention and are considered a hopeful system to solve the energy crisis and environmental issues. However, the low absorption of sunlight, high electron–hole recombination, and photo-decay effects are still limiting to the acquisition of an efficient photocatalyst with visible light activity. [
8]. Accordingly, it is necessary to fabricate and acquire nano-heterojunction photocatalysts with great solar-to-redox reaction conversion performance.
Recently, heterostructured photocatalysts have attracted researchers’ attention due to superior features such as improved electron–hole pair separation performance [
9]. The transition-metal dichalcogenides (TMDs) have obtained large attention owing to a series of interesting excellent properties [
10]. Between these materials, molybdenum disulfide (MoS
2) has great charge carrier density, high conductivity, and a controllable energy band gap, which is valuable and effective for optimizing environmental remediation and improving photocatalytic water splitting capability with visible light irradiation [
11]. As a kind of normal TMDs, MoS
2 has a bandgap energy of about 1.7~2.7 eV and proper valence conduction band potentials that own great photocatalytic efficiency and can employ an immense section of solar light [
12]. Unfortunately, the poor light harvesting ability, inactive removal pollutants performance, and quick recombination rate of photo-excited charge carriers are the natural disadvantages that inhibit the useful utilization of MoS
2 [
13]. In addition, Co
3O
4 also exhibits obvious stability under the photocatalytic system. In addition, Co
3O
4 possesses the matched band gap potential with MoS
2 to develop an efficient step-scheme (S-scheme) heterostructure because the valence band (VB) of Co
3O
4 is close to the conduction band (CB) of MoS
2 based on previous works [
12]. However, the Co
3O
4 catalyst absorbs primarily visible light and possesses a wide band gap energy, which prohibits their photocatalytic efficiency. Among these heterostructures, MoS
2 and Cobalt oxide (Co
3O
4) photocatalysts exhibit various superior properties for instance remarkable direct bandgaps, excellent mechanical properties, outstanding optical property as well as distinguished photocatalytic efficiency [
14,
15,
16]. On the other hand, bare MoS
2 and Co
3O
4 for direct photocatalytic water splitting or removal of contaminants are quite challenging because of their poor operation and high electron and hole recombination rate [
17,
18]. In addition, this heterojunction not only improved H
+ adsorption and promoted migration and transport of H
+ to the surface reactive sites, but further suppress the stacking or aggregation of materials [
19]. To solve this issue, a common method is to develop nano-heterostructure between Co
3O
4 and MoS
2 under visible light irradiation to improve their visible light activity. To overcome these barriers, substantial promotions have been made in the reasonable formation and design of MoS
2/Co
3O
4-based nanohybrids with well-established interfaces and nanostructures [
20]. Constructing heterojunction photocatalysts can assist the charge transfer between the diverse photocatalysts and enhance the separation of charge carriers. Therefore, the introduction of new-form nano-heterojunction photocatalysts with structural variety, band gap energy, and better visible-light utilization is superior in this investigate area [
21,
22].
In the present study, a new MoS
2/Co
3O
4 nanohybrid photocatalyst was fabricated by a facile hydrothermal route for powerful photocatalytic activity. The green–hydrothermal technique is energy-saving, cost-effective, environmentally friendly, and a green process compared with other synthesis methods [
23,
24]. To our knowledge, there is no report on the photocatalytic degradation of MoS
2/Co
3O
4 nanohybrids with visible light irradiation. A potential S-scheme heterojunction can be constructed between MoS
2 and Co
3O
4 while Schottky junction developed at the MoS
2/Co
3O
4 interface, respectively, which connected to enhance the charge carries separation ability. With the optimized amount of Co
3O
4 loading in the composite, the MoS
2/Co
3O
4 nano-heterojunction displayed a great removal MO rate of 95.6%, which was approximately 2.1 and 2.3 folds higher than that of bulk MoS
2 and Co
3O
4 samples, respectively. MoS
2/Co
3O
4 nanohybrids demonstrated higher photocatalytic efficiency than MoS
2 and Co
3O
4 on the removal of MO dye with visible light irradiation. This system opened up a new route for the controlled fabrication of MoS
2-based heterojunction and also improved their organic compound photodegradation activity. The further the surface reactive positions created by the huger specific surface area, the higher the redox efficiency due to the farther negative conduction capacity, and decreased recombination degree of photogenerated electrons and holes eventually resulted in improved photocatalytic performance.
2. Results and Discussion
XRD analysis was further collected to check the crystal phase form and structural composition of the as-fabricated photocatalysts, as exhibited in
Figure 1. During the hydrothermal process, the apparent feature peaks of MoS
2 were weaker while several new feature peaks appeared at 31.2°, 36.8°, 44.9° and 59.4° of the MoS
2/Co
3O
4 photocatalysts, which are related with the (220), (311), (400) and (511) diffraction face of Co
3O
4 (JCPDS No. 42–1467) [
15], respectively. The MoS
2 materials exhibits a series of primary feature peaks at 14.3°, 32.9°, 39.8°, 44.5°, 50.2°, 50.7°, and 60.7°, which assigned to the (002), (100), (103), (006), (105), and (110) crystal planes (JCPDS: 37-1492), respectively, and no other impurity peaks were perceived [
11]. Its value indicating that there is no altered signal in the XRD analysis of the MoS
2/Co
3O
4 nanocomposites compared to the MoS
2 and Co
3O
4 samples, suggesting that the presented Co
3O
4 content is fewer, and the dispersion is greater, and there are no other impurity peaks or feature peak shift. In addition, when the reaction solutions concentration of Co
3O
4 was increased from 5 wt% to 20 wt%, the (311) feature peak intensity of Co
3O
4 became higher, which indicated an enlarge in the dosage of Co
3O
4 in the heterojunction. In particular, contrast with the XRD analysis of Co
3O
4, the overall diffraction intensities of MoS
2/Co
3O
4 heterojunctions were altered at 36.8° with (311) diffraction face (see
Figure 1 inset). In addition, the feature peaks of MC-X (X = 5 wt%, 10 wt%, 15 wt%, and 20 wt%) are comparable to the feature XRD analysis of MoS
2 without additional impurity peaks, which suggests that the nanohybrids photocatalyst further possess coherent crystal structure [
25,
26].
In order to explore the development step of morphologies after hydrothermal process, SEM and TEM were characterized. As exhibited in
Figure 2a,b, apparently large-sized bulk MoS
2 and Co
3O
4 can be found, respectively. It is obtained that the microstructure of the photocatalyst is sphere-like and plate-like with approximately 5 μm equivalent size. However, the MoS
2 and Co
3O
4 possess a high grade of collection, which intensely suppress the surface of the active location of the photocatalysts. As displayed in
Figure 2c, the MoS
2/Co
3O
4 heterojunction presents a nanomaterial with an irregular surface, which gives it a greater specific surface area. Therefore, this reflection was verified in the specific surface area examination test. It is pronounced that despite the size of MC-X being immensely decreased, the crystals of MoS
2 still retain great crystallinity, as presented by the (002) major peak observed in the XRD patterns (
Figure 1). This nanohybrid of MoS
2/Co
3O
4 serves a sufficient area for the efficient load of Co
3O
4, which can mainly disperse MoS
2/Co
3O
4 nanohybrids and decrease aggregation. As displayed in
Figure 2d, the MoS
2/Co
3O
4-20% heterojunction displays an aggregated form with a dimension of approximately 4–10 nm. From
Figure 2e, the lattice fringe of 0.26 nm could be apparently noticed, which corresponds to the (002) plane of hexagonal MoS
2. The fringe spacing of 0.32 nm is assigned to the (311) plane of Co
3O
4. During further testing, it could be found that the MoS
2/Co
3O
4-20% nanohybrids were eventually constructed. The element mappings exhibit that the Mo, S, and Co elements are consistently distributed in the MoS
2/Co
3O
4-20% nanohybrids (
Figure 2f–i). The construction of the heterostructure of MoS
2/Co
3O
4 is broadly supposed to possess a definite effect on promoting the charge carrier’s separation ability of photocatalytic. The consistent distribution of Mo, S, and Co elements can confirm the effective formation of MoS
2/Co
3O
4 nanohybrids.
The UV-vis absorption spectra are generally adopted to check the visible light absorption capability of the photocatalyst, and the analysis of the as-prepared MoS
2/Co
3O
4 heterojunction is exhibited in
Figure 3. As can be observed, the two optical band gaps of MoS
2/Co
3O
4-X nanocomposites are 1.56~1.69 eV (
Eg1) and 2.31~2.61 eV (
Eg2), which are in good agreement with the MoS
2 and Co
3O
4 energy band gap. As MoS
2/Co
3O
4 photocatalyst is a photo-response sample, it possessed a wide wavelength absorption ability.
Eg1 is joined to the onset of O(II)-Co(III) excitations [
13,
14,
15]. The maximum response wavelength of MoS
2/Co
3O
4 is extended to 550 nm. Then, it presents a narrower band gap (
Eg2) of 2.31~2.61 eV after adding Co
3O
4, which is favorable to improve visible light application and photoexcited charge carriers. In addition, the UV-vis spectra of the MoS
2/Co
3O
4 nanohybrids are relativity close. The construction for this case is that the minor added dosage of Co
3O
4 does not generate an apparent alteration to the light absorption efficiency of the MoS
2/Co
3O
4 nanohybrids. The MC-20% nanohybrids display better light absorption and the absorption range has a minor red shift in the range of 550–680 nm. It could be assigned to the cooperation between MoS
2 and Co
3O
4. Hence, the Co
3O
4 materials may be a significant reason to enlarge the light absorption of the MoS
2/Co
3O
4 nanohybrids. The band gap energy (
Eg) of the corresponding photocatalyst was received via the transformed Kubelka–Munk function [
27,
28]. As exhibited in
Figure 3b, the
Eg2 (
Eg1) values of MC-5%, MC-10%, MC-15%, and MC-20% are about 2.61 (1.56) eV, 2.51 (1.61) eV, 2.45 (1.64) eV, and 2.31 (1.69) eV.
Then, X-ray photoelectron spectroscopy (XPS) is validated to recognize surface information of the fabricated samples.
Figure 4a displays the survey XPS spectra. As further observed, MC-20% is constructed of Mo, S, Co and O elements, which involve the integral elements of MoS
2/Co
3O
4 nanohybrids. There are no other impurity peaks emerging in survey spectra, suggesting that there are no other contaminate photocatalysts during hydrothermal route. The high-resolution spectrum of Mo 3d (
Figure 4b) could be divided into two peaks, which centered are at ∼229.6 eV and ∼232.8 eV, referring to Mo 3d 5/2 and Mo 3d 3/2 [
10]. Correspondingly, two feature peaks in the high-resolution S 2p XPS spectra of MC-20% centered at nearly 162.6 eV and nearly 163.6 eV (
Figure 4c) are referring to S 2p3/2 and S 2p1/2, respectively [
11]. As shown in
Figure 4d, binding energies of ∼780.1 eV and ∼795.5 eV referring to Co 2p 3/2 and Co 2p 1/2 for MC-20% nanohybrids, respectively [
14]. The O 1s spectra of MC-20% (
Figure 4e) binding energies at 529.8 eV and 531.8 eV, referring to the O
2− in the Co
3O
4 and the adsorbed oxygen [
15]. These results suggested that the MoS
2/Co
3O
4 nanohybrids were well fabricated with a strong interaction between MoS
2 and Co
3O
4. The elements concerned with the formation of MoS
2/Co
3O
4 nanohybrids were confirmed by the EDS spectra. Four noticeable peaks were obtained at nearly 0.8 keV, which was owing to the presence of oxygen atom, and other three peaks arising at 0.8 keV and 6.9 keV related to Co, and 2.3 keV, are related to Mo and S as exhibited in
Figure 4f. The results confirmed the successful fabrication of MoS
2/Co
3O
4 nanohybrids. The Brunauer-Emmett-Teller (BET) specific surface areas of MoS
2, Co
3O
4, and MC-20% are evaluated to be ~35.2, 24.6, and 68.7 m
2 g
−1, respectively (
Table 1). The main aperture size of MoS
2, Co
3O
4, and MC-20% are ~38.4, 48.5 and 8.9 nm, respectively. The pore volumes of MoS
2, Co
3O
4, and MC-20% are ~0.193 cm
3 g
−1, 0.124 cm
3 g
−1and 0.372 cm
3 g
−1, respectively. BET of MC-20% was more than 2 times than that of MoS
2 and Co
3O
4, demonstrating the introduced Co
3O
4 had a critical influence on raising the specific surface area of MoS
2.
Figure 5a exhibits the decay in absorbance peak intensity of 447 nm with extending irradiation time. It displays that the absorbance ability of MO was continuously removed with visible light activity in the presence of the as-fabricated photocatalysts. In addition, the absorbance of MO decreased significantly with the increase of irradiation time by using MC-20% nanohybrids as primarily photocatalyst under solar irradiation was recorded by UV–vis absorption spectroscopy, which proved the destruction of MO molecules. The photocatalytic activity of MO by as-fabricated MoS
2, Co
3O
4 and MC-X materials was evaluated with and without visible light activities (
Figure 5b). The control experiments displayed the capability values for the self-photocatalytic of MO pollutants under same light condition. This result suggests that the degradation of MO pollutants is owing to the photodegradation activity in the as-fabricated samples. On the other hand, the individual MoS
2 and Co
3O
4 samples could degrade about 45.8 and 41.4% of MO under 140 min of visible light radiation. The photodegradation capability of the photocatalysts followed the sequence: MC-20% nanohybrids > MC-15% nanohybrids > MC-10% nanohybrids > MC-5% nanohybrids > MoS
2 > Co
3O
4. This poor photodegradation performance could be assigned to the high recombination of electrons and holes in the bulk component [
29]. Noticeably, MC-20% nanohybrids described excellent photodegradation performance of MO pollutants (~95.6%) than bare MoS
2 or Co
3O
4, which was about 2.3 folds than bare Co
3O
4. The major process of photocatalytic reaction for the composite magnetic photocatalyst MoS
2/Co
3O
4 based on the basic principles and studies reported in the literature can be summarized as follows:
In addition, the photocatalytic reaction kinetics of an MO removal test was calculated to a pseudo first order equation model in order to check the photocatalytic activity. The linear relationship between ln (Co/C) vs. irradiation time (t) is exhibited in
Figure 5c and the removal rate constant (K
app) is displayed in
Figure 5d. The MC-20% nanohybrids were measured with a higher K
app value compared to the other synthesized samples. For effective utility purposes, the stability and reusability of MC nanohybrids were estimated by cycling runs. After every test, the sample was centrifuged, washed and dried before operating the next run. As shown in
Figure 5e, the photodegradation capability of the MC-20% sample was 95.6%, 92.3%, 90.5%, 88.3%, and 86.8% for first, second, third, fourth, and fifth test, respectively. The noticed reduction of the photocatalytic activity of MC-20% can be referred to an adsorbed MO and its evolution products on the surface of MC-20% nanohybrids, thus prohibiting the active sites and visible light from approaching the photocatalyst [
30]. The exploration of the photodegradation mechanism of the nanohybrids acted as a critical factor in the effective utilization of the heterojunction photocatalyst. Accordingly, the free radical trapping tests were performed, in which p-BQ (1,4-benzoquinone), IPA (2-propanol) and EDTA-2Na were adopted as scavenging free radical (
•O
2−,
•OH and h
+) agents [
31]. It could be observed from the analysis results in
Figure 5f that when various scavenging free radical agents were further joined, the photocatalytic efficiency of the MC-20% sample was influenced to altering degrees. The MC-20% sample was highly responsive to p-BQ and EDTA-2Na, implying that
•O
2− and h
+ were the major active species during the photodegradation activity, and
•OH only acted as an assisting active species. Consequently, these results demonstrate that the formation of MoS
2/Co
3O
4 with a proper ratio (20 wt%) is able to receive the better photocatalytic efficiency of MoS
2/Co
3O
4 in the nano-heterostructure. The stability and repeatability of the sample are important indexes to evaluate the performance of photocatalysts. XRD analysis was performed after the fourth recycling experiment. Taking MC-20% photocatalyst as an example, its activity did not change significantly after five cycles, which can still degrade 86.8% of MO (
Figure 5g). This result clearly suggests the stability of the fabricated MoS
2/Co
3O
4 NC photocatalyst. In addition,
Table 2 displays the photocatalytic performance of relevant heterojunction in previous reports [
31,
32,
33,
34,
35,
36]. By contrast, it is observed that the MoS
2/Co
3O
4 nanomaterial we prepared has remarkable benefits for organic compounds of high concentration in the photocatalytic process.
For confirm the presented S-scheme heterojunction, electron spin resonance (ESR) is adopted to recognize the free radicals produced in the system of MoS
2, Co
3O
4 and MC-20% nanohybrids. Applying DMPO as a spin capture agent, DMPO-
•O
2− and DMPO-
•OH are observed with light conditions. It could be found in
Figure 6a that the MC-20% photocatalyst displays the higher feature signal of DMPO-
•O
2−, while the feature signal of DMPO-
•O
2−of MoS
2 and Co
3O
4 is nearly undetectable. It exhibits that the construction of heterojunction is more favorable to the production of
•O
2−, which could be referred to the rapider migration and separation of photoexcited charge carriers in heterojunction catalysts [
36]. Furthermore, the DMPO-
•OH feature signal peaks of MoS
2, Co
3O
4 and MC-20% could be found in
Figure 6b, of which MC-20% possess the highest signal peak. Hence, it could be resulted that MC-20% nanohybrids can generate more
•O
2− and
•OH. The coincidence of
•O
2− and
•OH radicals further verify that the charge carrier separation mechanism is S-scheme. The charge transfer routes are analyzed by photoelectrochemical measurements. Electrochemical impedance spectroscopy (EIS) with visible light activity is exhibited in
Figure 6c, and the smaller semi-circular curve in Nyquist plots expresses the lesser charge carrier transport resistance. Apparently, the lowest charge carrier transport resistance in MC-20% imply the greatest efficient charge carrier transport routes [
37], which corresponds with the outstanding PHE capability of MC-20% nanohybrids. As exhibited in
Figure 6d, all the MC-20% nanohybrids electrodes exhibit a great transient photocurrent response than that bare MoS
2 and Co
3O
4 behaves at the highest intensity, implying the great efficient charge separation. Room-temperate PL spectra are carried out to check the charge carrier dynamics behavior. As displayed in
Figure 6e, the apparently reduced PL intensity of MC-20% also proves the prohibited electrons and holes recombination, which is in good approval with the above results. Furthermore, the performance of the charge carrier is further explored under photocatalytic redox reactions. The efficient separation of photo-induced charge carriers is a required situation for enhancing the capability of photodegradation [
38]. It is also found that the poorer the peak PL intensity, the lower the possibility of electron and hole pairs recombination rate, which is more favorable to the development of the pollutant removal reaction. This is owing to the evidence that the formation of nano-heterojunction improves the quick separation and transport of photo-excited charge carriers, hence, greatly suppressing the recombination rate of photo-induced charge carriers and enhancing the performance of charge carrier separation.
Based on the above data analysis and results, a photocatalytic activity mechanism of MO removal on MC nanohybrids is recommended, as exhibited in
Scheme 1. Following the introduction of Co
3O
4 materials on the surface of MoS
2, a firm heterostructure is organized. During exposure to visible light, photocatalysis activity has various efficiencies to the incident visible light in varied wavelength ranges. In this work, MC nanohybrids can absorb a broad range of incident visible light with the same conditions, thus generating more photoexcited charge carriers. More significantly, the system of photocatalytic reaction of nanohybrids design is higher than that of single-component design, implying in an improved interaction among reactant and photocatalyst. Ultimately, the excellent hydrogen evolution and pollutants removal promote the nanohybrids to have a higher charge carrier separation behavior [
39,
40]. In the meantime, a built electric field from Co
3O
4 to MoS
2 induces the downward band curving of MoS
2 and upward band curving of Co
3O
4 at the heterojunction interface, respectively. During visible light irradiation, both MoS
2 and Co
3O
4 could be photo-induced to produce electrons and holes. The electric field, curved band edges and the coulomb interaction notably drive the CB electrons of MoS
2 to rebuild with the VB holes of Co
3O
4, leading to the practical conservation of the high-oxidative holes in the VB of MoS
2 and high-reductive electrons in the CB of Co
3O
4, which significantly promotes the photo-response efficiency of MoS
2/Co
3O
4. This facile S-scheme system can efficiently weaken the recombination rate of photoinduced charge carriers in the heterojunction and improve the recombination of limited photo-generated electron and hole pairs, hence increasing the excellent redox activity of the photocatalyst. Finally, the enhanced photodegradation efficiency could be attributed to the increased visible light absorption ability, more reactive sites and powerful transfer of photo response charge carriers. In addition, the interface influence is larger than the efficiency of the photogenerated hole and electron to recombine. The electrons of the conduction band of MoS
2 contribute to the pollutant removal reduction reaction. Then, the holes of Co
3O
4 are oxidized in the valence band by the photocatalytic process. This novel S-scheme nanocomposite possesses better redox ability [
19,
26,
27], which apparently enhances the capability of photocatalytic pollutants removal. The CB and VB of MC with visible light activity, are induced to create photoexcited electrons and holes, respectively. On the other hand, the photo-response charge carriers in MoS
2 and Co
3O
4 will recombine rapidly, implying a weakening in photocatalytic capability. Although, owing to the presence of Co
3O
4 cocatalyst, photoexcited electrons will rapidly migrate to MoS
2, and also further water reduction so hydrogen evolution will appear. Furthermore, the presence of MoS
2 can weaken over-potential and serve more active sites for photocatalytic and can further enlarge the visible light absorption of Co
3O
4.
3. Materials and Methods
Sodium molybdenum oxide dehydrates (Na2MoO4·2H2O) (≥99.5%), l-cysteine (C3H7NO2S), Cobalt chloride hexahydrate (CoCl2·6H2O, ≥98%), Methyl orange (MO, ≥85%), Triton X-100, Potassium hydroxide (KOH, ≥85%), 1-Methyl-2-pyrrolidone (NMP, ≥99%), absolute ethanol (CH3CH2OH, ≥95%), and acetone (CH3COCH3, ≥99.5%) were used from Sigma-Aldrich (St. Louis, MO, USA). Deionized water was utilized to produce solutions to be examined. All the chemicals were adopted directly without purification.
Typically, 1.2 mmol of Na
2MoO
4·2H
2O and 2.5 mmol the
l-cysteine was used to 40 mL of deionized (DI) water to develop a homogeneous solution via stirring and ultrasonic treatment for 30 min. CoCl
2·6H
2O (2.5 mmol) was added in 40 mL distilled water and a certain amount of surfactant (1%,
w/
w, Triton X-100) was added subsequently. Subsequently, the homogeneous solution was transformed into a Teflon-lined stainless-steel autoclave (100 mL) and reacted at 180 °C for 20 h. After cooling down to room temperature naturally, the pristine dark precipitate was obtained by centrifugation, washed three times with deionized water, NMP, absolute ethanol and acetone, and dried at 70 °C in a vacuum oven overnight. A series of MoS
2/Co
3O
4 photocatalysts was provided by increasing various doses of Co
3O
4 (5 wt%~20 wt%); the individual samples are denoted as MC-X (X = 5 wt%, 10 wt%, 15 wt%, and 20 wt%). Finally, the nanocomposite is prepared via the hydrothermal method, as shown in
Scheme 2.
X-ray diffraction patterns (XRD) were determined by a Bruker D8 Advanced diffractometer (Billerica, MA, USA) at 5–75° with a scanning step of 20°/min at room temperature. Hitachi UV-4100 UV–visible spectrophotometer and Edinburgh Analytical Instruments F-7000 (Tokyo, Japan) were adopted to check the UV–vis diffusive reflectance spectra and photoluminescence (PL) spectra with an excitation wavelength of 620 nm. The microstructure and morphology of different samples also were tested by employing scanning electron microscopy (SEM) micrographs (ZEISS AURIGA, Oberkochen, Germany) and transmission electron microscopy (TEM) experiment (JEOL JEM-2100F, Tokyo, Japan). Electron paramagnetic resonance (EPR) spectra were observed on a Bruker A300 (Munich, Germany) at room temperature under atmospheric conditions. A CH Instruments electrochemical workstation with a standard three-electrode system was recorded to carry out electrochemical measurements in 0.2 M Na2SO4 solution. Furthermore, the saturated Pt electrode and Hg/Hg2Cl2 electrode were used as the counter electrode and reference electrode, respectively. The working electrode was used according to the next procedure: the as-prepared photocatalysts (20 mg) were dispersed in 10 mL of DI water to obtain a homogeneous suspension with a consecutive ultrasonic process. The chemical states and surface species were tested by X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific, Waltham, MA, USA).
Photodegradation of MO pollutants over various MC-X photocatalysts was performed at room temperature with visible light activity. A 350 W Xe lamp was employed as the visible-light source. A total of 20 mg of samples were added in 100 mL MO aqueous solution (20 mg L−1). The above reaction solution was further stirred under the dark for 20 min to organize absorption-desorption equilibrium between photocatalysts and MO compounds. Under light activity, a definite quantity of reaction solution was carried out from the combination solution at contributed intervals, followed by centrifugation to eliminate all the materials. The concentration alters of MO pollutants were obtained via UV-vis spectroscopy under photocatalytic activity.