Next Article in Journal
g-C3N4 as Photocatalyst for the Removal of Metronidazole Antibiotic from Aqueous Matrices under Lab and Pilot Scale Conditions
Next Article in Special Issue
Reaction Pathways of Gamma-Valerolactone Hydroconversion over Co/SiO2 Catalyst
Previous Article in Journal
Photocatalytic Degradation of Psychiatric Pharmaceuticals in Hospital WWTP Secondary Effluents Using g-C3N4 and g-C3N4/MoS2 Catalysts in Laboratory-Scale Pilot
Previous Article in Special Issue
Effect of Modulation and Functionalization of UiO-66 Type MOFs on Their Surface Thermodynamic Properties and Lewis Acid–Base Behavior
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Titanium Carbide Composite Hollow Cobalt Sulfide Heterojunction with Function of Promoting Electron Migration for Efficiency Photo-Assisted Electro-Fenton Cathode

State Key Laboratory of Heavy Oil Processing, China University of Petroleum, Beijing 102249, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(2), 253; https://doi.org/10.3390/catal13020253
Submission received: 30 November 2022 / Revised: 14 January 2023 / Accepted: 20 January 2023 / Published: 22 January 2023

Abstract

:
Constructing heterostructure within electrocatalysts proves to be an attractive approach to adjust the interfacial charge redistribution to promote the adsorption of reactive species and accelerate the charge transfer. Herein, we present the one-pot solvothermal synthesis of Ti3C2 supported hollow CoS2/CoS microsphere heterostructure with uneven charge distribution as the cathodic catalyst, which displays a superior quasi-first-order degradation rate (0.031 min−1) for sulfamethazine (SMT) in photo-assisted electric–Fenton (EF) process. CoS2/CoS/Ti3C2 is proven to favor the 2e oxygen reduction reaction (ORR), with H2O2 selectivity up to 76%. The built-in potential present in the heterojunction helps to accelerate electron transfer, thus promoting the production of H2O2. Subsequently, H2O2 is rapidly activated to produce ∙OH due to the synergistic effect of Co and S. Notably, CoS2/CoS/Ti3C2 exhibits enhanced photo-assisted EF (PEF) performance under light. The excellent photocatalysis properties of CoS2/CoS/Ti3C2 are attributed to that the unique hollow microsphere structure of catalyst improves the light absorption, and the uneven charge distribution of CoS2/CoS heterojunctions promotes the separation of photo-generated holes and electrons. Given the above advantages, CoS2/CoS/Ti3C2 cathode delivers a high degradation rate of 98.5%, 91.8%, and 94.5% for SMT, bisphenol A, and sulfadiazine, respectively, with TOC removal efficiency of 76% for SMT with 120 min. This work provides a novel light of the design and construction of efficient PEF cathodes for the treatment of antibiotic wastewater.

Graphical Abstract

1. Introduction

The environmental problems caused by antibiotics and organic pollutants have become more serious with the progress of human activities, however, conventional water treatment methods are difficult to achieve effective removal of such pollutants [1]. As a kind of electrochemical advanced oxidation processes (EAOPs), Electro-Fenton (EF) is a prospective water pollution treatment method [2] that generate H2O2 through the 2e-ORR by in-situ reduce O2 at the cathode [3], and then the ∙OH can be produced by catalysis of H2O2. However, the rapid formation of ∙OH is usually limited by the yield, selectivity, and reactivation of H2O2 [4]. Therefore, the rapid generation and activation of H2O2 is the key technology for efficient EF reaction [5]. Iron-based catalysis, including FeS, FeS2, etc., have been proven to have EF properties for activating H2O2, however, the tepid H2O2 yield greatly hinder the efficacy of EF technology [6]. One of the common solutions is combination metal-based materials with carbonaceous carriers because of that the oxyfunctional groups and defect sites of the carbon materials provide sufficient approach for the 2e ORR [7,8,9,10]. However, the large overpotential and long-term operation of the metal-carbon carrier cathode may cause the metal particles to fall off and carbonaceous carrier poisoning. Additionally, atomic precision doping or defect modulation of carbonaceous carriers may require higher costs. Therefore, it is necessary to find a suitable bifunctional catalyst with efficient 2e ORR selectivity and Fenton performance [11,12,13,14].
Cobalt sulfide has a variety of chemical formulas (such as CoS2, CoS, Co9S8, Co4S3), which have emerged as low-cost and abundant content materials with excellent catalytic activity based on both computational and experimental studies [15,16,17,18]. Among them, the metal cobalt pyrite (CoS2) exhibits excellent 2e ORR activity to produce H2O2 due to the single porphyrin structure of CoS2 inhibiting the O–O bond breaking of and promoting the production of H2O2, and the high 2e ORR selectivity of CoS2 is expected to serve as a high efficiency catalyst in EF reaction [16]. Additionally, hexagonal cobalt sulfide (CoS) is commonly used in Fenton-like fields because of narrow band gap structure and efficient activation of H2O2 performance, which is attributed to that CoS can provide suitable Lewis acid strength to adsorb and catalyze H2O2 to produce ∙OH [17]. Besides, the synergistic effect of Co and S effectively promotes electron transfer and accelerates the recycling of cobalt active species [18,19]. It is well known that cobalt sulfide with different formulas can interact with each other to form heterojunction with two opposite charge space field and built-in electric field area [20,21,22], which will promote the formation of uneven charge distribution. This uneven charge distribution may be able to facilitate the charge carrier transport in the photo-assisted electric–Fenton (PEF) process, thus exhibiting enhanced photo-assisted electrocatalytic activity [23]. However, the application of cobalt sulfide in electrocatalysis is always limited in its general conductivity. Transition metal carbide (Mxenes), as a novel two-dimensional material, exhibits excellent electrical conductivity (Titanium carbide(Ti3C2)~9880 S cm−1) [24,25] and highly exposed metal sites, which has great potential in the field of photocatalysis and electrochemistry, such as li-ion battery, supercapacitors, and photocatalytic hydrogen production [26,27,28,29]. There are reports that Ti metal at the end of titanium carbide exhibits poor oxygen resistance due to the high amount of exposed metal atoms, which promotes the regeneration of active sites [28]. In addition, the Schottky junction formed between Ti3C2 and metal semiconductors can suppress the photo-generated holes and electrons recombination, which also helps to enhance the photo-assisted EF performance [30,31,32]. Therefore, the combination of hollow heterostructure CoS2/CoS with Ti3C2 is meaningful for maintaining the 2e ORR behavior of CoS2 and Fenton behavior of CoS while optimizing activity, electrical conductivity, and sustainability of CoS2/CoS.
In this work, heterojunction CoS2/CoS nanocomposites with Ti3C2 were prepared by simple hydrothermal synthesis. CoS2/CoS/Ti3C2 cathode displayed a superior degradation performance for sulfamethazine (SMT) with degradation rates of 0.031 min−1. Rotating disk electrode (RDE) tests exhibit that CoS2/CoS/Ti3C2 has excellent 2e ORR characteristics, which contributes to promote the rapid production of H2O2. In addition, the uneven charge distribution accelerates the migration of charge carriers to facilitate photo-assisted EF process. UV-Vis DRS proves that CoS2/CoS/Ti3C2 has strong light absorption in the visible and UV regions, and the PL spectrum demonstrates that CoS2/CoS/Ti3C2 is provided with excellent charge-hole separation efficiency, which is attributed to the excellent electron-withdrawing characteristics of Ti3C2, as well as the heterostructure in CoS2/CoS. The total organic carbon (TOC) removal rate of SMT within 120 min is as high as 76%, and the possible degradation pathways of SMT are given based on UPLC-M-QTOF. The above work indicates that CoS2/CoS/Ti3C2 as a novel electrode material enriches the choice of available cathode materials and offers a new way for the degradation of persistent pollutants in the PEF process.

2. Results

The synthesis of CoS2/CoS/Ti3C2 involved of CoS2/CoS and subsequent CoS2/CoS/Ti3C2 (Figure 1). The prepared Ti3C2 was distributed in tetramethylammonium hydroxide (TMAOH) solution and processed with ultrasonic waves, and the exfoliated Ti3C2 was added to the solvothermal process of CoS2/CoS/Ti3C2. Finally, hollow CoS2/CoS grown in situ on the surface of Ti3C2. For comparison, the CoS2, CoS2/CoS, and CoS were obtained under the conditions of VEG:VDMF = 0.5, 0.9, and 2, respectively.
As shown in Figure 2a, all characteristic peaks could be well attributed to the diffraction peaks of CoS (PDF#65-3418) and CoS2 (PDF#89-1492) [33] while VDMF:VEG = 0.5 and 2. The mixed phase of CoS2 and CoS is obtained when the value of VEG:VDMF is 0.9 [22,34]. The characteristic peaks of Ti3C2 (36.4°) can be found in CoS2/CoS/Ti3C2. As shown in Raman spectra (Figure 2b), the peaks at ~293 cm−1 (Eg), and ~393 cm−1 (Ag) match well with the value of CoS2 in reference [22,35]. The peaks at ~475, and ~670 cm−1 represent the Eg, and A1g pattern of CoS [36], respectively. Both the characteristic peaks of CoS2 (~392.5 cm−1) and CoS (~476.0 and 679.0 cm−1) can be discovered in the pattern of CoS2/CoS/Ti3C2. It is worth noting that the peaks corresponding of CoS2 and CoS are red-shifted and blue-shifted in heterojunction, respectively. The opposite shifts of CoS and CoS2 characteristic peaks in Raman indicates the creation of a built-in potential region due to the production of heterojunctions, which help to promote electrons transport and enhance catalytic performance.
The morphology of Ti3C2 before HF etching shows a multi-layered bulk structure (Figure S1). SEM images of exfoliated Ti3C2 and CoS2/CoS/Ti3C2 are detected in Figure 3. The Ti3C2 peeled off by the intercalation method showed a uniform monolithic structure with slight wrinkles on the surface. With CoS2/CoS compounded with Ti3C2, CoS2/CoS is evenly dispersed on the surface of Ti3C2 (Figure 3b,c), which is due to the excellent dispersing effect of lot of terminating groups on Ti3C2 surface. More interestingly, the particles on the surface of Ti3C2 show typical hollow characteristics at high magnification (Figure 3d). The TEM of the CoS/CoS2 heterostructure showed the characteristics of hollow microspheres (Figure 4a), and the diameter of the microspheres is about 50 nm, and the particle size is relatively uniform. The wrinkled Ti3C2 lamella can be seen in Figure 4b,c, which demonstrated that the successfully combination of CoS/CoS2 and Ti3C2. Plain stripe spacing of 0.248 nm, 0.279 nm can be observed in the high-resolution TEM picture (Figure 4d), which corresponds to the plane (200) and (210) of CoS2, and interplanar spacing of 0.193 nm and 0.252 nm can be assigned to the plane (102) and (101) of CoS [22,33]. The clear lattice stripes and the absence of obvious amorphous regions at the interface between CoS2 and CoS confirm the formation of heterojunctions (Figure 4e). The overlapping patterns of diffraction rings of CoS2/CoS indicated both the formation of heterostructure and its good crystallinity (Figure 4f). Therefore, the heterogeneous structure of CoS2/CoS was successfully prepared. The element mappings of TEM and SEM demonstrates that Co and S mainly exist on the microspheres, and S and Co are uniformly distributed on the Ti3C2 surface, indicating that cobalt sulfide and Ti3C2 form a good composite structure (Figure 4g).
Without the addition of Ti3C2, CoS2/CoS exhibits the characteristics of aggregated particles (Figure S2a). Interestingly, the addition of Ti3C2 can form uniformly dispersed hollow microspheres. This difference [37] may be caused by the following reasons. Firstly, the functional group of Ti3C2 have a clear tendency to coordinate with Co2+ ions to generate the stable chelate complexes. The effect of Ti3C2 is commanding the release speed of Co2+ via the coordinating effect of Ti3C2-Co2+ chelate composite, thus slowing down deposition rate of CoS2 and CoS to facilitate the regulation of CoS2/CoS hollow microparticles. In the initial step, the sulfur powder in solution is converted by the reduction of urea, which contributes to the rapid acquisition of soft-template S-core at ambient temperature. At the same time, Co2+ is uniformly dispersed on the surface of S-core. As the reaction proceeds, CoS2 and CoS with interaction could be generated after crystal growth overcoming the anisotropy, thereby successfully introducing the heterogeneous structure. With the rise of reaction temperature, the S-core fades away, and the internal cavity becomes apparent, resulting in the formation of CoS2/CoS hollow microspheres with heterogeneous structure. Some examinations have shown that the hollow structure can exhibit a kinetically good open surface structure and a shorter mass and charge transport diffusion path [37]. Therefore, while this unique structure provides more active sites, the shorter mass and charge diffusion paths accelerate the ORR process of H2O2 synthesis and the subsequent Fenton process, thereby effectively improving the degradation efficiency. In addition, the multiple reflection of light in the hollow structure can also improve the efficiency of light utilization, which may improve the light absorption performance, thereby enhancing the PEF process.
The corresponding characteristic peaks of S 2p and Co 2p are displayed in CoS, CoS2, CoS2/CoS, and the peaks of Ti 2p could also be noticed in CoS2/CoS/Ti3C2 (Figure 5a). The high-resolution Ti 2p spectrum of CoS2/CoS/Ti3C2 exhibits peaks at ~464.8, ~458.5 eV corresponds to Ti-O (Figure 5b), and the peak centered at 472.0 eV could be assign to Ti-F, indicating the existence of Ti3C2 [28]. For the Co 2p spectrum, two distinct peaks can be found at binding energy of ~779 and 797.5 eV, indicating the presence of Co2+, which can be attributed to Co-S [33,34,37,38]. Two correlative characteristic lines at 782.5 and 802.4 eV can be attribute to satellite peaks of Co2+ and Co3+ (Figure 5c). In order to gain a clearer comprehending of the formation of CoS2/CoS heterojunctions as well as the charge distribution at the interface, S 2p was further investigated. The characterized S22− peaks at 162.3, 163.5 eV and the S2− at 161.5, 162.7 eV can be observed in CoS2 and CoS, respectively (Figure 5d). It is not difficult to find that the binding energy of S 2p3/2 (S2−) in CoS2/CoS structure move mildly to a lower energy about 0.1 eV than that in CoS because of additional negative charges in CoS side, in contrast, the energy of S22− is marginally higher than that in CoS2 [39], therefore, the distribution of opposite charges in CoS2/CoS and CoS2/CoS/Ti3C2 can be determined because of the electrons shift of S 2p3/2 in the opposite direction.
The influence of different cobalt sulfide crystal phases and composites on EF property was discussed in Figure 6a. 10 mg/L antibiotic solution in 0.05 mol/L Na2SO4 was considered as degraded pollutant and different samples loaded on carbon cloth (CC) as the cathodes. Note that CoS2/CoS/Ti3C2 displays the highest attenuation performance (~67%, 90 min) under the condition of electricity without light, which is higher than that of CoS (~45%,90 min), CoS2 (~56%, 90 min), and CoS2/CoS (~60%, 90 min). This result may be assigned to the excellent synergy between CoS2, CoS and Ti3C2, which significantly improves the transfer efficiency of charge carriers in 2e ORR, thereby helping to accelerate the formation of ∙OH. pH value generally has a significant impact for PEF systems involving transition metal-based catalyst. The catalytic performance of CoS2/CoS/Ti3C2 at different pH is shown in Figure 6b. As the initial pH decreases from 9 to 3, the removal performance of SMT improves from 48% to 80.5%. CoS2/CoS/Ti3C2 at pH = 3, 5, 7, 9 displays completely different attenuation efficiencies for SMT, which is helpful to show new viewpoint of the influence of catalyst on the efficiency of pollutants, as well as the formation of active components. CoS2/CoS/Ti3C2 shows a well documented prevalence of pH values, with similar attenuation of organic compounds at pH of 5–9. The removal efficiency at pH = 3 is significantly higher than that at other pH conditions due to the lowered oxidation potential of ·OH at high pH values (2.6–2.8 eV at pH = 3 and 1.90 eV at pH = 7).
The influence of catalyst load on degradation performance is shown in Figure 6c. With the increase in catalyst dosage from 3 mg/cm2 to 7 mg/cm2, the degradation rate of SMT improved from 73% to 98.4% (under electricity without light, 120 min, SMT = 10 mg/L). An increase in the catalyst loading capacity brings about an increase in the number of active sites, which contributes to improved degradation performance of SMT. The effect of current is investigated in Figure 6d. When the current intensity reaches 20 mA/cm2, the EF performance (89%) of the CoS2/CoS/Ti3C2 is significantly higher than that of 10 mA/cm2 (43%) and 5 mA/cm2 (24%), which is due to that the increase in current can provide more cathode electrons, thereby enhancing the oxygen reduction activity of the material. However, with the further rise of current intensity, the degradation rate of SMT did not show a significant increase, which may be associated with the occurrence of H2O2 decomposition at high currents. The synergistic effect of photo on EF is shown in Figure 6e, compared to the EF process without light, the removal performance of SMT increased by 10% within 120 min, and SMT decay rate can reach 98.5% under PEF conditions, and the reactions rate is as high as 0.031 min−1 (Figure S4), which is better than that of part reported cathodes (Table S1). Moreover, under the condition of light-only, the removal rate of CoS2/CoS/Ti3C2-cathode can reach 34.5% within 120 min, which is higher than that of CoS2/CoS (28.3%) and CoS (22.0%) under the same conditions. This result indicates that CoS2/CoS/Ti3C2 has better photocatalytic properties, which may be related to the formation of the built-in potential in CoS2/CoS and Schottky junctions between cobalt sulfide and titanium carbide that facilitates the transmission of photo-generated electrons. The removal capacity of different substrates was also examined, which may be directly related to practical applications. As shown in the Figure 6f, sulfadiazine (SDZ) and bisphenol A (BPA) showed similar removal performance as SMT, and the reaction rate constants for SDZ and BPA are 0.023 and 0.021 min−1, respectively, indicating that the CoS2/CoS/Ti3C2-cathode is suitable for the removal of antibiotic contaminants and has good potential for industrial applications. All of these results show that CoS2/CoS/Ti3C2 has excellent photo-assisted EF performance.
To further illustrate the mechanism of the PEF process, we drop-casted CoS2, CoS2/CoS, and CoS2/CoS/Ti3C2 nanomaterials on disk ring-electrode without carbon addition and tested the materials selectivity and activity for two-electron oxygen reduction in alkaline solutions (KOH, 0.1 M) without interference from carbon [16]. The electrocatalytic activity was examined by CV in O2 saturated electrolyte for all three materials (Figure 7a). Compared with CoS2 (−1.08 V) and CoS/CoS2 (−1.05 V), the larger oxygen reduction potential (−1.0 V) and double-layer current value of CoS/CoS2/Ti3C2 indicate its better activity, which may be due to the presence of Ti3C2 and built-in potential in CoS2/CoS accelerate the charge transfer, thereby accelerating the formation of OOH* (Equation (2)), and OOH* reacts with H+ to form H2O2 as the Equation (3). The oxidation peaks at −0.55 V can be ascribed to the oxidation of Co2+, and Co3+ was further converted to Co2+ when it became electrons at the cathode [13]. As the oxidation of Co(II), the sulfide with metal defects is formed on the catalyst surface, which can be further oxidized to generate sulfur intermediate products (S2O32−, S2O42−), thus, the peaks at 0.166 V could be due to the formation of intermediate sulfides, and these sulfur oxides could be further activated to produce sulfate radicals [40,41]. These results prove that CoS2/CoS/Ti3C2 has excellent redox activity, which can be used as the EF cathode to facilitate the produce of active species. The electron transfer capability of the material was studied by Nyquist plots. The impedance test results are shown in Figure 7b, compared with pure CoS and CoS2, and CoS2/CoS significantly reduces the electron transfer resistance, which is attributed to the built-in potential in CoS2/CoS heterojunction helps to increase the charge mobility. After compounding with titanium carbide, the conductivity of CoS/CoS2/Ti3C2 material is further improved, which is due to the excellent conductivity of titanium carbide [22].
LSVs of all the catalysts at different speed are depicted in Figure S3a–c. According to the Koutecky-Levich (K-L) formula, the number of electrons transfer (n) for the ORR is calculated (Text S1) [28]. The number of electrons transfer is calculated to be 2.85, 2.12, and 2.22 for CoS2, CoS2/CoS, and CoS2/CoS/Ti3C2, respectively, demonstrating that CoS2/CoS/Ti3C2 has a tendency to 2e ORR process, which is attributed to that built-in potential formed by the heterostructure helps to accelerate the electron migration rate, thus promoting the formation of OOH* on the CoS2 surface. Although the number (2.22) of electron transfer after composite Ti3C2 is reduced according to the K-L equation, CoS2/CoS/Ti3C2 exhibits a positive onset potential of −0.13 V vs. SCE (Figure S3d), which is higher than that of CoS2/CoS (−0.16 V), indicating that Ti3C2 can enhance the electrocatalytic activity of CoS2. In addition, rotating ring disk electrode (RRDE) tests of CoS2/CoS/Ti3C2-casted disk from 0.2 V to −1 V vs. SCE at 50 mV/s and different rotational speed were recorded in Figure 7c, and corresponding H2O2 selectivity (p) and the n value are calculated on the basis of Equations (S3) and (S4). The H2O2 selectivity of CoS2/CoS/Ti3C2 up to about 76% and remained above 70% over a wide potential range (Figure 7d), and the polarization curves of the CoS2/CoS/Ti3C2 show about 0.75 V vs. RHE of half-wave potential, which is close to thermodynamic limit of two-electron ORR. Apart from the high activity, the stability tests of the material are shown in Figure 7e, with a ring voltage setting of 1.48 V and disc voltage of 0.5 V, the i-t test of CoS2/CoS/Ti3C2 material showed excellent stability with a stable H2O2 selectivity of about 70% within 2 h. The above results indicate that CoS2/CoS/Ti3C2 has excellent 2e ORR activity and stability, and 2e ORR is uptaken at low overpotentials.
The photo adsorption properties of different samples were analyzed by UV-Vis DRS (Figure 8a). CoS exhibits strong light absorption in the ultraviolet and visible regions (edge of approximately 450 nm and 280 nm), and CoS2/CoS and CoS2/CoS/Ti3C2 display stronger absorption in the scope of 280–700 nm. The tauc plot method (Figure 8b) is used to fit the light absorption value as the Equation (1). The band gap of CoS2/CoS/Ti3C2 (~1.95 eV) is slightly lower than that of CoS (~2.25 eV) and CoS2/CoS (~2.06 eV), which indicates that the heterostructure of CoS2/CoS and the schottky junction formed under the modification of Ti3C2 broaden the light absorption range of the material, which helps to absorb more visible light to further improve the catalytic activity. Furthermore, the photocurrent curve (Figure 8c) is used to analyze the electron-hole separation efficiency of the materials. Compared with CoS and CoS2/CoS, CoS2/CoS/Ti3C2 show higher photocurrent intensity and stronger photo response performance, which is attributed to the electron withdrawing effect of Ti3C2 facilitating the separation of photo-generated holes and electrons [42], moreover, the built-in potential in the heterojunction also helps to accelerate the transfer of electrons. PL spectrum (Figure 8d) also proves this fact, and the lowest PL peak of CoS2/CoSTi3C2 indicates that photo-generated electron recombination is well suppressed, which helps to extend the life of photo-generated carriers and holes, thereby improving photocatalytic activity. There is a significant gap in the removal capacity of SMT under light (98.5%) and dark (89%) conditions, respectively (Figure 6e). Under the condition of only light, the effective removal of SMT is about 34.5% within 120 min. The above facts illustrate that CoS2/CoS/Ti3C2 have excellent photo-assisted electrocatalytic activity.
( ɑ h υ ) 1 / n = A ( h υ E g )
O 2 + * + ( H + + e ) OOH *
OOH * + ( H + + e ) H 2 O 2 + *  
Electron paramagnetic resonance (EPR) was conducted by the spin trap agents of 5,5-dimethyl-1-pyrroline N-oxide (DMPO) to confirm the existence of various reactive oxide species in this reaction process. The result is shown in Figure 9a, and weak DMPO-∙OH signal can be observed under only light conditions, which is attributed to the interaction of photogenerated electrons with ∙O2 to produce ·OH as the Equations (4)–(7). Therefore, the removal of SMT under light conditions alone may be attributed to the ∙O2 (−0.33 eV). After turning on the power, obvious signals of four peaks with the intensity value of 1:2:2:1 (DMPO-·OH) adducts can be found in the EPR spectrum [43], indicating that the existence of ·OH in CoS2/CoS/Ti3C2 system under PEF condition. Furthermore, the weak characteristic peaks of DMPO-SO4 also could be detected, which is due to the oxidation of S2− and SO42− in the solution to sulfur oxides (such as S2O32−, S2O42−), and then sulfur oxides could be further activated to produce SO4, and the high oxidation potential (2.7~3.1 eV) of SO4 can also effectively promote the degradation of the substrates (Figure 9b). Six characteristic peaks of DOMP-·O2 compounding are observed, suggesting the production of ·O2 under the action of light and electricity (Figure 9c). Quenching experiments were performed to clarify the contribution of different ROS in this PEF system. TBA and MeOH act as the trapping agents of ·OH and SO4, respectively. Figure 9d showed that SMT decay rate is significantly inhibited when TBA (100 mM) is added, and the degradation performance is as low as 50.5% in 120 min, indicating the major roles of ·OH in reaction system. In the presence of MeOH (100 mM), the decay rate of SMT still remains around 71.2%, indicating that SO4 has a relatively low contribution for degradation process, therefore, the ∙OHs are critical to the degradation performance in the CoS2/CoS/Ti3C2 system. Anti-ion interference experiments were also carried out, as shown in Figure S6. It can be seen that NO3 and Cl have little effect on the degradation process of SMT, however, the presence of HCO3 makes the removal efficiency of SMT slightly lower (about 20%), but it is undeniable that even in the presence of anions, the removal rate of SMT is still more than 70%, which indicates that the CoS2/CoS/Ti3C2 cathode system has a better ability for resistance to ion interference. Based on the above results, the CoS2/CoS/Ti3C2 cathode system can effectively produce active radicals under EF conditions, and the degradation performance was further enhanced by the addition of light. Compared with FeS2NWs/Ti3C2 cathode 28, CoS2/CoS/Ti3C2 showed higher 2e ORR catalytic activity, higher TOC removal rate (76%), and greater PEF catalytic performance owing to the uneven charge distribution of CoS2/CoS heterojunction. Moreover, due to the excellent 2e ORR properties of the CoS2/CoS/Ti3C2 material, the addition of reduced graphene oxide (RGO) is effectively avoided.
The stability and long-term runnability of the cathode are essential factors in assessing the ability of the material to perform in real-world applications. The removal efficiency of SMT by the CoS2/CoS/Ti3C2 cathode after five cycle tests is well preserved (Figure 9e), decreasing by about 14.5% compared with that for the first cycle. The main reasons for the performance decay of the CoS2/CoS/Ti3C2 cathode may be that deactivation of functional group of Ti3C2 due to the occupation of the active site by contaminants, and quality loss of CoS2/CoS/Ti3C2 due to sulfide being further oxidized to sulfur oxides (such as S2O32−, S2O42−) [40,41]. Despite this, the performance remained around 85% after five cycles, indicating its excellent stability. Figure 9f shows the variation of total organic carbon content with time in SMT degradation process. As the reaction time increased from 30 to 120 min, the mineralization rate of SMT by the CoS2/CoS/Ti3C2 cathode PEF system increased from 31% to 76%. The high mineralization rate is ascribed to the oxidation of reactive oxygen species (·O2, ·OH, SO4). Stability of the catalysts in water treatment system is crucial for their engineering applications. To further reveal the element features of the used catalyst, XRD and XPS of used CoS2/CoS/Ti3C2 were presented in Figure S5. The XRD feature of the used catalyst still has the characteristic peaks of cobalt sulfide, but the corresponding peak intensity is reduced, which may be related to the adsorption of binder polymers and contaminants on the catalyst sites. Moreover, the reusability of the electrode was confirmed through the XPS results of the electrode before and after cycle. The XPS spectrum of Figure S5b shows that the used catalyst has similar characteristics to the fresh CoS2/CoS/Ti3C2 catalyst, and the characteristic peaks corresponding to Co 2p and S 2p are weakened (Figure S6c,d), which may be due to the inevitable ion leaching and sulfide being further oxidized to intermediate sulfur oxides in acidic conditions, and then sulfur oxides could be further activated to produce SO4 to promote the degradation of the SMT. Thus, CoS2/CoS/Ti3C2 can be considered as an effective PEF cathode material for more thorough removal of pollutants.
CoS 2 / CoS / Ti 3 C 2 + h υ     h + + e +  
e + O 2     · O 2
· O 2 + e + 2 H +     H 2 O 2
H 2 O 2 + h υ     2 · OH  
To clarify the degradation pathway of SMT in the CoS2/CoS/Ti3C2 system, the transformation products of SMT were analyzed by QTOF LC/MS (UPLC-M-QTOF). The proposed intermediate compounds are shown in Figure S7. SMT was attacked by by ∙O2, SO4 and ∙OH to form intermediate products, and the reaction pathway (Figure 10) included (1) oxidation process of aniline group in SMT to produce N-(4-methylpyrimidin-2-yl)-4-(hydroxyamino)-benzenesulfonamide (P1, m/z, 281.0) and N-(4-methylpyrimidin-2-yl)-4-nitrobenzene-sulfonamide (P1-1, m/z, 296.1). (2) The attack of ROS could lead to the cleavage of S−N bond to form 4-methylpyrimidin-2-amine (P2, m/z, 110.2) and 4-aminobenzenesulfonic acid, which was further oxidized to form 4-nitrobenzenesulfonic acid (P2-1, m/z, 203.1) [44], and this result was frequently reported in AOPs-based water treatment processes for the degradation of sulfonamide antibiotics [45]. (3) Smiles-type rearrangement of SMT to generate 4-(2-imino-4-methylpyrimidin-1(2H)-yl) aniline, which was further oxidized to form 4-methyl-1-(4-nitrosophenyl) pyrimidin-2(1H)-imine (P3, m/z, 214.9) and 4- methyl-1-(4-nitrophenyl) pyrimidin-2(1H)-imine, (P3-1, m/z, 232.9). (4) The SO2 extrusion of SMT or the oxidation of P3-1 resulted in the formation of nitrobenzene (P4, m/z, 124.9), which could be further oxidized to phenol (P4-1, m/z, 94.9). The intermediate products generated above will become smaller molecules through dealkylation and ring-opening reactions under the action of ROS, and ultimately, the above intermediates could be partially mineralized to carbon dioxide and water.
Given the above results, mechanism of CoS2/CoS/Ti3C2 cathode for removal of micro-pollutants was proposed. The energy of EVB and ECB of CoS2/CoS/Ti3C2 are calculated according to Equations (8) and (9) [46]. According to the calculation results, the values of EVB and ECB are 1.79 eV and −0.46 eV, respectively. Under visible light irradiation, electrons are generated in CB and holes are generated in VB of CoS. Due to the electron-absorbing properties of Ti3C2, the conducting electrons of CoS are further transferred to the surface of it, which facilitates the separation of holes and electrons. The electrons in the Ti3C2 could reduce O2 to generate ·O2 (−0.33 eV) [42]. At the same time, the continuously generated e and h+ could make the CB and VB bend upward, establishing Schottky barriers through the contact interface and facilitating the separation of photo-generated electrons and holes [28,46]. In addition, O2 is rapidly converted to produce H2O2 on the CoS2/CoS/Ti3C2 surface as Equations (5) and (6). The generated H2O2 is then activated on the surface of CoS2/CoS/Ti3C2 and participated in the activated process to generate ·OH as the Equation (10). Due to the synergistic effect of S2− and Co, Co(II) can be regenerated on the surface of the cathode (Equation (11)) [18,19]. In addition, decomposition of H2O2 (Equation (7)) can be improved under irradiation. Sulfate in the solution can be oxidized to form persulfate, which is further activated to produce sulfate radicals [13, 41,42]. Therefore, the ROS produced in the PEF process could attack functional groups of contaminants quickly and effectively.
E VB = χ E e + 0 . 5 E g  
E CB = E VB E g  
where χ is the absolute electronegativity of the semiconductor. Ee is the energy of free electrons on the hydrogen scale (4.50 eV). Eg is the band gap of the semiconductor [20].
Co ( II ) + H 2 O 2     Co ( III ) + · OH + OH  
2 S 2 + 16 Co ( III ) + 8 H 2 O 2 SO 4 2 + 16 H + + 16 Co ( II )  

3. Experimental Section

3.1. Chemicals

Cobalt sulfate heptahydrate (CoSO4∙7H2O, 99.99%), urea (CH4N2O, 99.5%), sublimed sulfur (S, 99.5%), ethylene glycol ((CH2OH)2, 98%), N, N-Dimethylformamide (C3H7NO, 99.8%), tetramethylammonium hydroxide solution (C4H13NO, 25%), titanium aluminum carbide powder (Ti3AlC2, 98%), tert-butanol (TBA, 99.5%), and methanol (MeOH, 99.5%) are chemicals, which were all purchased from Aladdin Chemical Reagent Co., Ltd. (Shanghai, China), unless otherwise stated.

3.2. Preparation of CoS2/CoS and CoS2/CoS/Ti3C2

In the typical procedure, 10 mmol urea, 2 mmol CoSO4, and 25 mmol sulfur powder were added to 60 mL mixture solution of DMF and EG. The mixture solution was stirred for 30 min and then sonicated for 30 min. The well mixed precursor solution is added to Teflon-lined autoclave. The solution is maintained in an automatic heating oven at 180 °C for 12 h. After the hydrothermal reaction, the sample was naturally cooled to room temperature, centrifuged, and washed three times with ethanol and ultrapure water. The samples were dried under vacuum. The CoS2, CoS2/CoS, and CoS were obtained under the conditions of VEG:VDMF = 0.5, 0.9, and 2.0.
Ti3AlC2 was etched in HF solution to obtain Ti3C2. Subsequently, the 25% TMAOH solution was used to intercalate Ti3C2 to obtain a monolithic layer of Ti3C2. Similarly, to obtain CoS2/CoS/Ti3C2, (50 mg) of stripped Ti3C2 was placed in a mixed solution, and the temperature was maintained at 180 °C for 12 h. After centrifugal washing, the sample is vacuum dried.

3.3. Characterization and Experimental Process

Detailed information of preparation of CoS2/CoS/Ti3C2 cathode and degradation test was shown in Supplementary Materials Text S2.

4. Conclusions

In summary, CoS2/CoS/Ti3C2 displayed a superior catalytic performance in photo-assisted-electro-Fenton process compared with other materials (Table S3). The charged surface derived from the built-in potential benefited to facilitate the electron migration. Both the oxygen reduction ability for 2e process and photogenerated electron-hole separation of CoS2/CoS/Ti3C2 were enhanced owing to the uneven charge distribution of CoS2/CoS heterojunction. CoS2/CoS/Ti3C2 cathode demonstrated a high PEF performance with 98.5% degradation rate for SMT at 20 mA/cm2 within 120 min and excellent performance retention of 85% after five cycles. Moreover, ESR signals revealed that the main ROS in the degradation process of SMT was ∙OH. The design and construction of photoelectric Fenton electrocatalysts with heterojunction offers a promising strategy to enhance the performance and competitiveness. In addition, based on the excellent catalytic performance of this material, further research with various substrates (such as other Mxenes materials) is undergoing to further improve the water splitting process and broaden the application scope of CoS2/CoS-based materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13020253/s1, Figure S1. Morphology of Ti3C2 sample after etching with hydrofluoric acid. Figure S2. TEM of (a) CoS2/CoS (b) CoS2/CoS/Ti3C2. Figure S3. LSVs of (a) CoS2, (b) CoS2/CoS, (c) CoS2/CoS/Ti3C2 performed with RDE test in O2-saturated (697 µg/cm2) and the corresponding K-L curves (inset), (d) LSVs of CoS2/CoS and CoS2/CoS/Ti3C2. Figure S4. seudo first order reaction rate of different substrates in CoS2/CoS/Ti3C2 cathode system. Figure S5. (a) XRD pattern and (b-d) XPS of used CoS2/CoS/Ti3C2 catalyst. Figure S6. Resistance to ion interference of CoS2/CoS/Ti3C2 cathode Figure S7. MS spectrum of intermediate products in SMT degradation process. Table S1. Performance comparison with similar materials. References [47,48,49,50,51] are cited in Supplementary Materials.

Author Contributions

Data curation, F.C. and H.L.; Formal analysis, F.C. and F.Y.; Funding acquisition, Y.L.; Methodology, F.C. and Z.W.; Resources, S.C.; Software, N.C. and Z.W.; Validation, F.Y. and Y.L.; Writing—review and editing, F.C., F.Y. and S.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China, grant number No. 21776308.

Data Availability Statement

All the relevant data used in this study have been provided in the form of figures and tables in the published article, and all data provided in the present manuscript are available to whom it may concern.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cheng, S.; Shen, C.; Zheng, H.; Liu, F.; Li, A. OCNTs encapsulating Fe-Co PBA as efficient chainmail-like electrocatalyst for enhanced heterogeneous electro-Fenton reaction. Appl. Catal. B. 2020, 269, 118–785. [Google Scholar] [CrossRef]
  2. Zhu, Y.; Zhu, R.; Xi, Y.; Zhu, J.; Zhu, G.; He, H. Strategies for enhancing the heterogeneous Fenton catalytic reactivity: A review. Appl. Catal. B 2019, 255, 117–739. [Google Scholar] [CrossRef]
  3. Ganiyu, S.O.; Huong, L.T.X.; Bechelany, M.; Esposito, G.; van Hullebusch, E.D.; Oturan, M.A.; Cretin, M. A hierarchical CoFe-layered double hydroxide modified carbon-felt cathode for heterogeneous electro-Fenton process. J. Mater. Chem. A 2017, 5, 3655–3666. [Google Scholar] [CrossRef]
  4. Haider, M.R.; Jiang, W.L.; Han, J.L.; Sharif, H.M.A.; Ding, Y.C.; Cheng, H.Y.; Wang, A.J. In-situ electrode fabrication from polyaniline derived N-doped carbon nanofibers for metal-free electro-Fenton degradation of organic contaminants. Appl. Catal. B 2019, 256, 117–774. [Google Scholar] [CrossRef]
  5. Zhang, C.; Li, F.; Wen, R.; Zhang, H.; Elumalai, P.; Zheng, Q.; Chen, H.; Yang, Y.; Huang, M.; Ying, G. Heterogeneous electro-Fenton using three-dimension NZVI-BC electrodes for degradation of neonicotinoid wastewater. Water Res. 2020, 182, 115975. [Google Scholar] [CrossRef]
  6. Chu, L.; Sun, Z.; Fang, G.; Cang, L.; Wang, X.; Zhou, D.; Gao, J. Highly effective removal of BPA with boron-doped graphene shell wrapped FeS2 nanoparticles in electro-Fenton process: Performance and mechanism. Sep. Purif. Technol. 2021, 267, 118–680. [Google Scholar] [CrossRef]
  7. Qin, X.; Zhao, K.; Quan, X.; Cao, P.; Chen, S.H. Highly efficient metal-free electro-Fenton degradation of organic contaminants on a bifunctional catalyst. J. Hazard. Mater. 2021, 416, 125859. [Google Scholar] [CrossRef]
  8. Cheng, S.; Zheng, H.; Shen, C.; Jiang, B.; Liu, F.; Li, A. Hierarchical Iron Phosphides Composite Confined in Ultrathin Carbon Layer as Effective Heterogeneous Electro-Fenton Catalyst with Prominent Stability and Catalytic Activity. Adv. Funct. Mater. 2021, 31, 2106311. [Google Scholar] [CrossRef]
  9. Wang, H.; Wang, Y.; Zhang, J.; Liu, X.; Tao, S. Electronic structure engineering through Fe-doping CoP enables hydrogen evolution coupled with electro-Fenton. Nano Energy 2021, 84, 105–943. [Google Scholar] [CrossRef]
  10. Yu, D.; He, J.; Wang, Z.; Pang, H.; Li, L.; Chen, Y.; Zheng, Y.; Zhang, J. Mineralization of norfloxacin in a CoFe–LDH/CF cathode-based heterogeneous electro-fenton system: Preparation parameter optimization of the cathode and conversion mechanisms of H2O2 to ·OH. Chem. Eng. J. 2021, 417, 129–240. [Google Scholar] [CrossRef]
  11. Zhang, D.; Yin, K.; Tang, Y.; Wei, Y.; Tang, H.; Du, Y.; Liu, H.; Chen, Y.; Liu, C. Hollow sea-urchin-shaped carbon-anchored single-atom iron as dual-functional electro-Fenton catalysts for degrading refractory thiamphenicol with fast reaction kinetics in a wide pH range. Chem. Eng. J. 2022, 427, 130–996. [Google Scholar] [CrossRef]
  12. Dong, P.; Chen, X.; Guo, M.; Wu, Z.; Wang, H.; Lin, F.; Zhang, J.; Wang, S.; Sun, C.; Zhao, H. Heterogeneous electro-Fenton catalysis with self-supporting CFP@MnO2-Fe3O4/C cathode for shale gas fracturing flowback wastewater. J. Hazard. Mater. 2021, 412, 125–208. [Google Scholar] [CrossRef]
  13. Guo, M.; Lu, M.; Zhao, H.; Lin, F.; He, F.; Zhang, J.; Wang, S.; Dong, P.; Zhao, C. Efficient electro-Fenton catalysis by self-supported CFP@CoFe2O4 electrode. J. Hazard. Mater. 2022, 423, 127033. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, M.; Gong, Y.; Ma, N.; Zhao, X. Promoted photoelectrocatalytic degradation of BPA with peroxymonosulfate on a MnFe2O4 modified carbon paper cathode. Chem. Eng. J. 2020, 399, 125088. [Google Scholar] [CrossRef]
  15. Guo, M.; Qayum, A.; Dong, S.; Jiao, X.; Chen, D.; Wang, T. In situ conversion of metal (Ni, Co or Fe) foams into metal sulfide (Ni3S2, Co9S8 or FeS) foams with surface grown N-doped carbon nanotube arrays as efficient superaerophobic electrocatalysts for overall water splitting. J. Mater. Chem. A 2020, 8, 9239–9247. [Google Scholar] [CrossRef]
  16. Sheng, H.; Hermes, E.D.; Yang, X.; Ying, D.; Janes, A.N.; Li, W.; Schmidt, J.R. Electrocatalytic Production of H2O2 by Selective Oxygen Reduction Using Earth-Abundant Cobalt Pyrite (CoS2). ACS Catal. 2019, 9, 8433–8442. [Google Scholar] [CrossRef]
  17. Kim, M.; Kim, S.H.; Lee, J.H.; Kim, J. Unravelling lewis acidic and reductive characters of normal and inverse nickel-cobalt thiospinels in directing catalytic H2O2 cleavage. J. Hazard. Mater. 2020, 392, 122347. [Google Scholar] [CrossRef]
  18. Ding, Y.; Hu, Y.; Peng, X.; Xiao, Y.; Huang, J. Micro-nano structured CoS: An efficient catalyst for peroxymonosulfate activation for removal of bisphenol A. Sep. Purif. Technol. 2020, 233, 116022. [Google Scholar] [CrossRef]
  19. Yein, W.T.; Wang, Q.; Wu, J.; Wu, X. Converting CoS-TEA hybrid compound to CoS defective ultrathin nanosheets and their enhanced photocatalytic property. J. Mol. Liq. 2018, 268, 273–283. [Google Scholar] [CrossRef]
  20. Zhao, W.; Feng, Y.; Huang, H.; Zhou, P.; Li, J.; Zhang, L.; Dai, B.; Xu, J.; Zhu, F.; Sheng, N.; et al. A novel Z-scheme Ag3VO4/BiVO4 heterojunction photocatalyst: Study on the excellent photocatalytic performance and photocatalytic mechanism. Appl. Catal. B 2019, 245, 448–458. [Google Scholar] [CrossRef]
  21. Li, Y.; Zhao, Q.; Zhang, Y.; Li, Y.; Fan, L.; Li, F.T.; Li, X. In-situ construction of sequential heterostructured CoS/CdS/CuS for building “electron-welcome zone” to enhance solar-to-hydrogen conversion. Appl. Catal. B 2022, 300, 120–763. [Google Scholar] [CrossRef]
  22. Ma, D.; Hu, B.; Wu, W.; Liu, X.; Zai, J.; Shu, C.; Tadesse Tsega, T.; Chen, L.; Qian, X.; Liu, T.L. Highly active nanostructured CoS2/CoS heterojunction electrocatalysts for aqueous polysulfide/iodide redox flow batteries. Nat. Commun. 2019, 10, 3367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Shi, Y.; Wang, X.; Liu, X.; Ling, C.; Shen, W.; Zhang, L. Visible light promoted Fe3S4 Fenton oxidation of atrazine. Appl. Catal. B 2020, 277, 119–229. [Google Scholar] [CrossRef]
  24. Zhao, R.; Qian, Z.; Liu, Z.; Zhao, D.; Hui, X.; Jiang, G.; Wang, C.; Yin, L. Molecular-level heterostructures assembled from layered black phosphorene and Ti3C2 MXene as superior anodes for high-performance sodium ion batteries. Nano Energy 2019, 65, 104037. [Google Scholar] [CrossRef]
  25. Cheng, L.; Chen, Q.; Li, J.; Liu, H. Boosting the photocatalytic activity of CdLa2S4 for hydrogen production using Ti3C2 MXene as a co-catalyst. Appl. Catal. B 2020, 267, 118–379. [Google Scholar] [CrossRef]
  26. Liao, Y.; Qian, J.; Xie, G.; Han, Q.; Dang, W.; Wang, Y.; Lv, L.; Zhao, S.; Luo, L.; Zhang, W.; et al. 2D-layered Ti3C2 MXenes for promoted synthesis of NH3 on P25 photocatalysts. Appl. Catal. B 2020, 273, 119054. [Google Scholar] [CrossRef]
  27. Ai, Z.; Shao, Y.; Chang, B.; Huang, B.; Wu, Y.; Hao, X. Effective orientation control of photogenerated carrier separation via rational design of a Ti3C2(TiO2)@CdS/MoS2 photocatalytic system. Appl. Catal. B 2019, 242, 202–208. [Google Scholar] [CrossRef]
  28. Xia, Y.; Yang, F.; Zhang, B.; Xu, C.; Yang, W.; Li, Y.F. Fabrication of novel FeS2 NWs/Ti3C2 cathode for Photo-Electro-Fenton degradation of sulfamethazine. Chem. Eng. J. 2021, 426, 130–719. [Google Scholar]
  29. Xu, H.; Cao, J.; Shan, C.; Wang, B.; Xi, P.; Liu, W.; Tang, Y. MOF-Derived Hollow CoS Decorated with CeOx Nanoparticles for Boosting Oxygen Evolution Reaction Electrocatalysis. Angew. Chem. Int. Ed. Engl. 2018, 57, 8654–8658. [Google Scholar] [CrossRef]
  30. Al-Musawi, T.J.; Arghavan, S.M.A.; Allahyari, E.; Arghavan, F.S.; Othmani, A.; Nasseh, N. Adsorption of malachite green dye. onto almond peel waste: A study focusing on application of the ANN approach for optimization of the effect of environmental parameters. Biomass Convers. Biorefinery 2020, 273, 119054. [Google Scholar] [CrossRef]
  31. Nasseh, N.; Samadi, M.T.; Ghadirian, M.; Panahi, A.H.; Rezaie, A. Photo-catalytic degradation of tamoxifen by using a novel. synthesized magnetic nanocomposite of FeCl2@ac@ ZnO: A study on the pathway, modeling, and sensitivity analysis using artificial neural network (AAN). J. Environ. Chem. Eng. 2022, 10, 107450. [Google Scholar] [CrossRef]
  32. Rahimi, S.M.; Panahi, A.H.; Allahyari, E.; Nasseh, N. Breaking down of low-biodegradation Acid Red 206 dye using bentonite/Fe3O4/ZnO magnetic nanocomposite as a novel photo-catalyst in presence of UV light. Chem. Phys. Lett. 2022, 794, 139480. [Google Scholar] [CrossRef]
  33. Fenton, J.L.; Schaak, R.E. Structure-Selective Cation Exchange in the Synthesis of Zincblende MnS and CoS Nanocrystals. Angew. Chem. Int. Ed. Engl. 2017, 56, 6464–6467. [Google Scholar] [CrossRef] [PubMed]
  34. Sheng, H.; Janes, A.N.; Ross, R.D.; Kaiman, D.; Huang, J.; Song, B.; Schmidt, J.R.; Jin, S. Stable and selective electrosynthesis of hydrogen peroxide and the electro-Fenton process on CoSe2 polymorph catalysts. Energy Environ. Sci. 2020, 13, 4189–4203. [Google Scholar] [CrossRef]
  35. Kinner, T.; Bhandari, K.P.; Bastola, E.; Monahan, B.M.; Haugen, N.O.; Roland, P.J.; Bigioni, T.P.; Ellingson, R.J. Majority Carrier Type Control of Cobalt Iron Sulfide (CoxFe1–xS2) Pyrite Nanocrystals. J. Phys. Chem. C 2016, 120, 5706–5713. [Google Scholar] [CrossRef]
  36. Peng, S.; Han, X.; Li, L.; Zhu, Z.; Cheng, F.; Srinivansan, M.; Adams, S.; Ramakrishna, S. Unique Cobalt Sulfide/Reduced Graphene Oxide Composite as an Anode for Sodium-Ion Batteries with Superior Rate Capability and Long Cycling Stability. Small 2016, 12, 1359–1368. [Google Scholar] [CrossRef]
  37. Yun, X.; Lu, T.; Zhou, R.; Lu, Z.; Li, J.; Zhu, Y. Heterostructured NiSe2/CoSe2 hollow microspheres as battery-type cathode for hybrid supercapacitors: Electrochemical kinetics and energy storage mechanism. Chem. Eng. J. 2021, 426, 131–328. [Google Scholar] [CrossRef]
  38. Hao, J.; Yang, W.; Peng, Z.; Zhang, C.; Huang, Z.; Shi, W. A nitrogen doping method for CoS2 electrocatalysts with enhanced water oxidation performance. ACS Catal. 2017, 7, 4214–4220. [Google Scholar] [CrossRef]
  39. Zhu, C.; Wang, A.L.; Xiao, W.; Chao, D.; Zhang, X.; Tiep, N.H.; Chen, S.; Wang, J.; Kang, X.; Ding, J.; et al. In Situ Grown Epitaxial Heterojunction Exhibits High-Performance Electrocatalytic Water Splitting. Adv. Mater. 2018, 30, e1705516. [Google Scholar] [CrossRef]
  40. Chu, L.; Sun, Z.; Cang, L.; Fang, G.; Wang, X.; Zhou, D.; Gao, J. A novel sulfite coupling electro-fenton reactions with ferrous sulfide cathode for anthracene degradation. Chem. Eng. J. 2020, 400, 125945. [Google Scholar] [CrossRef]
  41. Xing, M.; Xu, W.; Dong, C.; Bai, Y.; Zeng, J.; Zhou, Y.; Zhang, J.; Yin, Y. Metal sulfides as excellent co-catalysts for H2O2 decomposition in advanced oxidation processes. Chem 2018, 4, 1359–1372. [Google Scholar] [CrossRef]
  42. Zhang, W.; Zhang, Z.; Kwon, S.; Zhang, F.; Stephen, B.; Kim, K.K.; Jung, R.; Kwon, S.; Chung, K.B. Photocatalytic improvement of Mn-adsorbed g-C3N4. Appl. Catal. B 2017, 206, 271–281. [Google Scholar] [CrossRef]
  43. Sopaj, F.; Oturan, N.; Pinson, J.; Podvorica, F.; Oturan, M.A. Effect of the anode materials on the efficiency of the electro-Fenton process for the mineralization of the antibiotic sulfamethazine. Appl. Catal. B. 2016, 199, 331–341. [Google Scholar] [CrossRef]
  44. Yang, F.; He, M.; Wu, T.F.; Hao, A.P.; Zhang, S.B.; Chen, Y.D.; Zhou, S.B.; Zhen, L.Y.; Wang, R.; Yuan, Z.L.; et al. Sulfadiazine oxidation by permanganate: Kinetics, mechanistic investigation and toxicity evaluation. Chem. Eng. J. 2018, 349, 56–65. [Google Scholar] [CrossRef]
  45. Chen, F.; Yang, F.; Liu, H.; Che, S.; Zhang, G.; Xu, C.; Li, Y. One-pot preparation of surface vulcanization Co-Fe bimetallic aerogel for efficient sulfadiazine degradation. Chem. Eng. J. 2022, 430, 132–904. [Google Scholar] [CrossRef]
  46. Wu, Z.; Shen, J.; Ma, N.; Li, Z.; Wu, M.; Xu, D.; Zhang, S.; Feng, W.; Zhu, Y. Bi4O5Br2 nanosheets with vertical aligned facets for efficient visible-light-driven photodegradation of BPA. Appl. Catal. B. 2021, 286, 119–937. [Google Scholar] [CrossRef]
  47. Deng, F.; Li, S.; Zhou, M.; Zhu, Y.; Qiu, S.; Li, K.; Ma, F.; Jiang, J. A biochar modified nickel-foam cathode with iron-foam catalyst in electro-Fenton for sulfamerazine degradation. Appl. Catal. B 2019, 256, 117–796. [Google Scholar] [CrossRef]
  48. Zhang, Y.; Chen, Z.; Wu, P.; Duan, Y.; Zhou, L.; Lai, Y.; Li, S. Three-dimensional heterogeneous Electro-Fenton system with a novel catalytic particle electrode for Bisphenol A removal. J. Hazard. Mater. 2020, 393, 120448. [Google Scholar] [CrossRef]
  49. Chen, Y.P.; Yang, L.M.; Chen, J.P.; Zheng, Y.M. Electrospun spongy zero-valent iron as excellent electro-Fenton catalyst for enhanced sulfathiazole removal by a combination of adsorption and electro-catalytic oxidation. J. Hazard. Mater. 2019, 371, 576–585. [Google Scholar] [CrossRef]
  50. Liu, K.; Yu, M.; Wang, H.; Wang, J.; Liu, W.; Hoffmann, M. Multiphase Porous Electrochemical Catalysts Derived from Iron-Based Metal-Organic Framework Compounds. Environ. Sci. Technol. 2019, 53, 6474–6482. [Google Scholar] [CrossRef] [Green Version]
  51. Xie, W.; Yuan, S.; Mao, X.; Hu, W.; Liao, P.; Tong, M.; Alshawabkeh, A.N. Electrocatalytic activity of Pd-loaded Ti/TiO2 nanotubes cathode for TCE reduction in groundwater. Water Res. 2013, 47, 3573–3582. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration for the composition of CoS2/CoS/Ti3C2.
Figure 1. Schematic illustration for the composition of CoS2/CoS/Ti3C2.
Catalysts 13 00253 g001
Figure 2. (a) X-ray powder diffraction (XRD) and (b) Raman characterization.
Figure 2. (a) X-ray powder diffraction (XRD) and (b) Raman characterization.
Catalysts 13 00253 g002
Figure 3. SEM images of (a) Ti3C2, (bd) CoS2/CoS/Ti3C2 under different magnifications.
Figure 3. SEM images of (a) Ti3C2, (bd) CoS2/CoS/Ti3C2 under different magnifications.
Catalysts 13 00253 g003
Figure 4. (ae) CoS2/CoS/Ti3C2 HRTEM images. (f) Different lattice fringes correspond to diffraction rings patterns (g) corresponding EDS elemental mapping of C, Co, Ti, and S.
Figure 4. (ae) CoS2/CoS/Ti3C2 HRTEM images. (f) Different lattice fringes correspond to diffraction rings patterns (g) corresponding EDS elemental mapping of C, Co, Ti, and S.
Catalysts 13 00253 g004
Figure 5. (a) XPS survey spectra of CoS2, CoS, CoS2/CoS, CoS2/CoS/Ti3C2, and high-resolution XPS spectra (b) Ti 2p of CoS2/CoS/Ti3C2. High-resolution (c) Co 2p, (d) S 2p of CoS2, CoS, CoS2/CoS, and CoS2/CoS/Ti3C2.
Figure 5. (a) XPS survey spectra of CoS2, CoS, CoS2/CoS, CoS2/CoS/Ti3C2, and high-resolution XPS spectra (b) Ti 2p of CoS2/CoS/Ti3C2. High-resolution (c) Co 2p, (d) S 2p of CoS2, CoS, CoS2/CoS, and CoS2/CoS/Ti3C2.
Catalysts 13 00253 g005
Figure 6. Removal efficiency of prepared samples in different reaction systems, (a) CoS2, CoS, and CoS2/CoS, CoS2/CoS/Ti3C2 (pH ≈ 5), (b) different initial pH (c) different catalyst loading (pH ≈ 3), (d) different current intensity, (e) EF under different light conditions, (f) Different reaction substrates. Unless otherwise stated, the reaction conditions are SMT = 10 mg/L, 100 mL; catalyst = 5 mg/cm2; I = 20 mA/cm2; under electricity without light.
Figure 6. Removal efficiency of prepared samples in different reaction systems, (a) CoS2, CoS, and CoS2/CoS, CoS2/CoS/Ti3C2 (pH ≈ 5), (b) different initial pH (c) different catalyst loading (pH ≈ 3), (d) different current intensity, (e) EF under different light conditions, (f) Different reaction substrates. Unless otherwise stated, the reaction conditions are SMT = 10 mg/L, 100 mL; catalyst = 5 mg/cm2; I = 20 mA/cm2; under electricity without light.
Catalysts 13 00253 g006
Figure 7. (a) CV of CoS2, CoS2/CoS, and CoS2/CoS/Ti3C2 in O2-saturated electrolyte. (b) EIS plots of different samples. (c) LSVs of CoS2/CoS/Ti3C2 performed with RRDE test in O2-saturated. (d) The H2O2 selectivity and number of electron transfers, (e) i-t curve of disc-ring electrode at 1600 rpm.
Figure 7. (a) CV of CoS2, CoS2/CoS, and CoS2/CoS/Ti3C2 in O2-saturated electrolyte. (b) EIS plots of different samples. (c) LSVs of CoS2/CoS/Ti3C2 performed with RRDE test in O2-saturated. (d) The H2O2 selectivity and number of electron transfers, (e) i-t curve of disc-ring electrode at 1600 rpm.
Catalysts 13 00253 g007
Figure 8. (a) DRS spectra, (b) the corresponding Tauc plots (the colored dotted lines are tangents to the corresponding solid lines), (c) transient photocurrent responses, and (d) PL spectrum of CoS, CoS2/CoS, CoS2/CoS/Ti3C2.
Figure 8. (a) DRS spectra, (b) the corresponding Tauc plots (the colored dotted lines are tangents to the corresponding solid lines), (c) transient photocurrent responses, and (d) PL spectrum of CoS, CoS2/CoS, CoS2/CoS/Ti3C2.
Catalysts 13 00253 g008
Figure 9. (a) DMPO-∙OH spin-trapping EPR spectra in the condition of (a) only light, (b) PEF, (c) DMPO-O2 spin-trapping EPR spectra, (d) quenching experiment, (e) cycle performance, (f) TOC removal rate.
Figure 9. (a) DMPO-∙OH spin-trapping EPR spectra in the condition of (a) only light, (b) PEF, (c) DMPO-O2 spin-trapping EPR spectra, (d) quenching experiment, (e) cycle performance, (f) TOC removal rate.
Catalysts 13 00253 g009
Figure 10. Proposed reaction pathway for the oxidation of SMT at ambient pH by CoS2/CoS/Ti3C2 cothade system.
Figure 10. Proposed reaction pathway for the oxidation of SMT at ambient pH by CoS2/CoS/Ti3C2 cothade system.
Catalysts 13 00253 g010
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

Chen, F.; Yang, F.; Che, S.; Liu, H.; Chen, N.; Wu, Z.; Li, Y. Titanium Carbide Composite Hollow Cobalt Sulfide Heterojunction with Function of Promoting Electron Migration for Efficiency Photo-Assisted Electro-Fenton Cathode. Catalysts 2023, 13, 253. https://doi.org/10.3390/catal13020253

AMA Style

Chen F, Yang F, Che S, Liu H, Chen N, Wu Z, Li Y. Titanium Carbide Composite Hollow Cobalt Sulfide Heterojunction with Function of Promoting Electron Migration for Efficiency Photo-Assisted Electro-Fenton Cathode. Catalysts. 2023; 13(2):253. https://doi.org/10.3390/catal13020253

Chicago/Turabian Style

Chen, Fengjiang, Fan Yang, Sai Che, Hongchen Liu, Neng Chen, Zhijie Wu, and Yongfeng Li. 2023. "Titanium Carbide Composite Hollow Cobalt Sulfide Heterojunction with Function of Promoting Electron Migration for Efficiency Photo-Assisted Electro-Fenton Cathode" Catalysts 13, no. 2: 253. https://doi.org/10.3390/catal13020253

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