Next Article in Journal
Adjoint Algorithm Design of Selective Mode Reflecting Metastructure for BAL Applications
Previous Article in Journal
Pyroclastic Dust from Arequipa-Peru Decorated with Iron Oxide Nanoparticles and Their Ecotoxicological Properties in Water Flea D. magna
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancement of Peroxydisulfate Activation for Complete Degradation of Refractory Tetracycline by 3D Self-Supported MoS2/MXene Nanocomplex

College of Life and Environmental Sciences, Central South University of Forestry and Technology, Changsha 410004, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2024, 14(9), 786; https://doi.org/10.3390/nano14090786
Submission received: 15 April 2024 / Revised: 26 April 2024 / Accepted: 29 April 2024 / Published: 30 April 2024

Abstract

:
Antibiotic abuse, particularly the excessive use of tetracycline (TC), a drug with significant environmental risk, has gravely harmed natural water bodies and even posed danger to human health. In this study, a three-dimensional self-supported MoS2/MXene nanohybrid with an expanded layer spacing was synthesized via a facile one-step hydrothermal method and used to activate peroxydisulfate (PDS) for the complete degradation of TC. The results showed that a stronger •OH signal was detected in the aqueous solution containing MoS2/MXene, demonstrating a superior PDS activation effect compared to MoS2 or Ti3C2TX MXene alone. Under the conditions of a catalyst dosage of 0.4 g/L, a PDS concentration of 0.4 mM, and pH = 5.0, the MoS2/MXene/PDS system was able to fully eliminate TC within one hour, which was probably due to the presence of several reactive oxygen species (ROS) (•OH, SO4•−, and O2•−) in the system. The high TC degradation efficiency could be maintained under the influence of various interfering ions and after five cycles, indicating that MoS2/MXene has good anti-interference and reusability performance. Furthermore, the possible degradation pathways were proposed by combining liquid chromatography–mass spectrometry (LC-MS) data and other findings, and the mechanism of the MoS2/MXene/PDS system on the degradation process of TC was elucidated by deducing the possible mechanism of ROS generation in the reaction process. All of these findings suggest that the MoS2/MXene composite catalyst has strong antibiotic removal capabilities with a wide range of application prospects.
Keywords:
MoS2; MXene; PDS; tetracycline; ROS

1. Introduction

The content of refractory and toxic antibiotics in water bodies is increasing due to the random dumping of personal care products (PPCPs), active pharmaceutical compounds, and other medical supplies in recent years, posing a serious threat to the ecological environment of natural water bodies and even human health [1,2]. Among them, tetracyclines (TCs) are widely used in medicine and animal husbandry due to their low cost and broad-spectrum activity, resulting in significant environmental risks [3,4]. It is difficult to remove TC from the environment using common methods. For instance, adsorption can merely transfer TC in water to the surface of a solid adsorbent without removing it fundamentally, while biological resistance and chemical stability would limit the biodegradation process, resulting in a low degradation efficiency (below 56%) [5,6]. Therefore, an innovative method for eliminating TC from the environment is critical.
Applying advanced oxidation processes (AOPs) to the oxidative destruction of organic compounds is a major area of current research. When compared to the traditional hydroxyl radical (•OH), the sulfate radicals (SO4•−) of sulfate radical-based advanced oxidation processes (SR-AOPs) have a greater redox potential (2.5–3.1 eV vs. 1.8–2.7 eV), a longer half-life (30–40 s vs. 20 ns), a wider pH range, and higher selectivity, providing them an edge in degrading organic contaminants [7,8]. Nevertheless, only when activated under certain conditions (such as UV light [9], ultrasound [10], carbon-based materials [11], and transition metals [12]) that peroxydisulfate (PDS) can produce reactive oxygen species (ROS), including sulfate radicals (SO4•−) and superoxide radicals (O2•−), to properly breakdown organic contaminants [13]. However, UV radiation and ultrasonic waves are limited by their high cost and low energy efficiency, while carbon-based catalysts can only adsorb organics to their surface rather than decompose them [14,15]. It has been reported that transition metals could be useful catalysts for PDS activation [7]. Generally, Co2+ and Fe2+ are the most widely utilized transition metals, and they have a high ability to activate PDS [16]. Unfortunately, limitations such as high cost and toxicity from Co2+ leaching and the conversion of Fe2+ to Fe3+ to produce precipitation would interfere with PDS activation, restricting the application of these two ions in practice [16,17].
Layered transition metal dichalcogenides (TMDs, generally MoS2) have been reported to be employed as inorganic activators or catalysts to activate PDS (S2O82−) to generate ROS directly for pollutant disintegration [16,18]. Furthermore, MoS2 is an exceptional chemical reagent for effectively activating PDS, owing to its low cost, excellent chemical stability, mechanical robustness, non-toxicity/low toxicity, and flexible physical/chemical properties [19]. However, the native interlayer spacing of MoS2 is rather small (0.3 nm), limiting the capacity of ions/molecules to interact with internal S atoms and impeding the efficient activation of PDS by active surfaces and edge sites [17,20]. Therefore, MoS2 nanocomposites with an enlarged interlayer spacing are required. It has been proven that solvothermal synthesis with Mo and S precursors can produce MoS2 nanosheets with an increased interlayer spacing (0.94 nm), along with a surface-exposed plentiful active group, allowing them to exhibit stronger activity [17]. Additionally, MoS2 nanosheets are straightforward to aggregate since their intrinsic high surface energy results in a drop in electron transfer rate and surface-active sites [21]. As a result, directional growth of MoS2 on a hydrophilic, large, specific surface carrier is another effective strategy for increasing layer spacing, thereby exposing more active sites and improving its application in wastewater treatment [22]. In recent years, novel MXene materials (such as Ti3C2TX MXene) have attracted extensive attention in various fields due to their excellent performance [23,24]. MXene, a two-dimensional transition metal substance consisting of selectively etched ternary nitrides or carbides, with ecologically favorable features, a large surface area, excellent chemical stability, unique metallic thermal/electrical conductivity, and hydrophilic qualities, has been widely used in sensors, lithium-ion batteries, supercapacitors, and environmental applications [25]. It is mainly composed of transition metal carbides, and the chemical formula is commonly expressed as Mn+1XnTx, where M is a transition element (e.g., V, Mo, Ti, etc.), X is C or N, and Tx stands for the surface terminations. Among the various MXene materials, Ti3C2Tx is the most widely used [24]. However, the fabrication of devices and functional coatings based on Ti3C2TX remains challenging as they are susceptible to chemical degradation through oxidation to TiO2 [24,26,27]. In addition, the hydrothermal/solvothermal techniques that are typically used to produce MXene-based nanomaterials can accelerate MXene oxidation. It is found that Ti3C2TX MXene is easily oxidized during the hydrothermal process due to the presence of dissolved oxygen, resulting in performance loss [28]. Consequently, a simple solution approach is urgently needed to completely encapsulate and anchor MoS2 nanosheets on the surface of Ti3C2TX MXene to combine the useful properties of these two components into functional heterojunctions and overcome their respective deficiencies for efficient activation of PDS for catalytic degradation of TC.
In this study, we designed an effective and stable MoS2/MXene catalyst for the successful activation of PDS to generate ROS to catalyze the breakdown of residual refractory TC in water. Furthermore, MoS2 nanoflowers were strongly coupled with Ti3C2TX MXene via a well-designed one-step hydrothermal process. In this method, a new sulfur source, sodium diethyldithiocarbamate trihydrate (DDC), and a capping agent, EDTA, were used to substantially widen the interlayer spacing of the nanocomposite and to gain abundant surface-active sites, while the excellent chelating ability of DDC allowed the MoS2 precursor to be tightly coupled to Ti3C2TX MXene, thus effectively preventing the chemical reaction of Ti3C2TX MXene at the initial stage of the hydrothermal reaction [23,29]. The synthesized composites were characterized and analyzed using SEM, FTIR, and XPS, and the effects of different factors on the degradation of TC in water by MoS2/MXene-activated PDS were meticulously investigated, while also inferring the possible degradation mechanism of TC. The findings of this study will contribute to the construction of MoS2, MXene, and their composites, as well as the production of powerful persulfate-activating catalysts, allowing for the efficient purification of refractory persistent organic pollutants in water.

2. Materials and Methods

2.1. Materials

Titanium aluminum carbide (Ti3AlC2) was obtained from 11 Technology Co., Ltd. in Jilin, China. Sodium molybdate dihydrate (Na2MoO4·2H2O), sodium diethyldithiocarbamate trihydrate (DDC), ethylenediaminetetraacetic acid (EDTA), tetracycline hydrochloride (TC), sodium persulfate (Na2S2O8, peroxydisulfate (PDS)), and p-benzoquinone (BQ) were purchased from MACKLIN Biochemical Co., Ltd., Shanghai, China. Sulfuric acid (H2SO4), sodium hydroxide (NaOH), lithium fluoride (LiF), anhydrous ethanol (C2H5OH), sodium chloride (NaCl), sodium nitrate (NaNO3), sodium dihydrogen phosphate (NaH2PO4), sodium carbonate (Na2CO3), and tert-butyl alcohol (TBA) were purchased from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. All chemicals used are of analytical grade, and deionized water was used in all experiments.

2.2. Synthesis of MoS2/MXene

The facile one-step hydrothermal synthesis method of MoS2/MXene is shown in Figure 1. Firstly, the preparation of Ti3C2Tx MXene dispersion followed the method described in previous studies [28,30]. A standard hydrothermal procedure was used to synthesize MoS2/MXene nanocomposites. Briefly, 0.7566 g of sodium molybdate dihydrate and 0.2 g of EDTA were dispersed in 50 mL of the pre-prepared Ti3C2Tx MXene dispersion, which underwent absolute ultrasonic processing for 1 h. After oscillation mixing for 0.5 h, 0.7650 g of DDC was added to the suspension with continuous mixing for 0.5 h. Subsequently, the combination was heated at 180 °C for 24 h in a Teflon-lined hydrothermal autoclave reactor. Upon cooling the solution to room temperature naturally, the resulting precipitate was washed several times with deionized water/absolute ethanol, and then, the MoS2/MXene complex was collected after freeze-drying under vacuum. MXene and MoS2 monomers were prepared under the same conditions for subsequent analysis and comparison as the blank control. In addition, several MoS2/MXene samples were prepared by setting different mass ratios of MoS2 and MXene (1:1, 2:1, and 3:1).

2.3. Characterization

A variety of characterization approaches were used to observe the microscopic surface arrangement and element valence makeup of the prepared MoS2/MXene samples. The morphology and structure of different samples were observed using a scanning electron microscope (SEM; ZEISS Sigma 300, Roedermark Germany) equipped with an energy-dispersive X-ray spectrometer (EDX). The chemical bonds and surface functional groups of MoS2, MXene, and MoS2/MXene were investigated via Fourier-transform infrared spectroscopy (FTIR; Thermo Scientific Nicolet iS20, MA, USA). The surface elemental composition and chemical state were analyzed using X-ray photoelectron spectroscopy (XPS; Thermo Scientific K-Alpha, MA, USA).

2.4. ESR Analysis

To better illustrate the role of reactive oxygen species in the activation of PDS by MoS2/MXene, electron spin resonance (ESR) spectra were obtained using an ESR spectrometer (JESFA200, JEOL) in air, with a resonance frequency of 9225.960 MHz, a microwave power of 0.998 mW, a center field of 329.4 mT, a modulation frequency of 100 kHz, a sweep width of 5 × 1 mT, a time constant of 0.03 s, and a sweep time of 30 s, and 5,5-dimethyl-1-pyrroline -N-oxide (DMPO) was used as a spin trap to capture hydroxyl radicals and generate DMPO-•OH signals using hyperfine splitting at room temperature.

2.5. TC Degradation Experiments

All TC degradation experiments were carried out in 150 mL Erlenmeyer flasks at 25 °C under varying experimental conditions, which included the material system (MXene only, MoS2 only, and MoS2/Mxene), catalyst dosage, PDS consumption, pH, and initial concentration of TC.
Specifically, 50 mL of TC solution was placed in a 150 mL flask. After a certain amount of the catalyst material was added to the solution, which was stirred for 10 min, PDS of a certain concentration was added to initiate the degradation reaction. During the reaction process, 1.0 mL of the reaction solution was withdrawn at specific time intervals (10, 20, 30, 40, 50, and 60 min) with filtering. The recovered solution was filtered through a 0.22 μm filter and diluted to the standard curve concentration range with deionized water. The residual concentration was then detected at 356 nm using a UV–visible spectrophotometer (UV-2700; Shimadzu, Kyoto, Japan).
Additionally, (1) the impacts of inorganic anions (Cl, NO3, H2PO4, and CO32−), which are frequently present in natural water, were investigated in this work at a particular concentration of TC solution, a certain mass of catalyst, and a given molar mass of PDS solution. (2) Different concentrations of EtOH, TBA, and BQ were used as quenchers to study their effects on the oxidative degradation of TC.

3. Results and Discussion

3.1. Synthesis and Characterization of MoS2/MXene

According to our previous reports, MXene is prone to oxidation at high temperatures and has poor stability features, which limits its use to some extent [28]. In this study, MXene, MoS2, and MoS2/MXene nanomaterials were prepared under the same conditions, and it can be seen from the SEM image in Figure 2a that the MXene surface was oxidized at a high temperature, and a significant amount of TiO2 grew on the surface of the initially smooth MXene sheet [21]. The pure MoS2 nanosheets were connected in clusters and dispersed in a petal-like pattern in all directions, as shown in Figure 2b,c. The SEM images of MoS2/MXene at different magnifications are presented in Figure 2d–f. It was observed that numerous MoS2 nanoflowers were uniformly growing on the MXene surface. With further magnification, it could be clearly observed that an intercalation structure between MoS2 and MXene was formed, and the disappearance of TiO2 particles on the MXene surface suggested that the addition of MoS2 partially shielded the oxidation of MXene. The EDX mapping image of MoS2/MXene is displayed in Figure 2g, and the results showed that the elements C, O, Mo, S, and Ti were uniformly distributed. The results of the elemental composition analysis via EDX spectroscopy are shown in Table S1. It was determined that the atomic ratio of Mo to S was about 1:2, suggesting effective synthesis of MoS2, and the high contents of C and O indicated the possibility of the presence of abundant organic functional groups. The SEM results indicated that MoS2/MXene composite materials with a stable 3D intercalation structure were successfully prepared.
As shown in Figure 3a, after being exposed to high temperatures, MXene exhibited a broad absorption peak at 3300–3500 cm−1 and a significant absorption peak at 450–530 cm−1, which were attributed to the stretching vibration of the O-H single bond and the Ti-O-Ti bond, respectively [31]. It indicated that the MXene surface was readily oxidized to generate TiO2 under high temperatures. Additionally, the disappearance of the characteristic peak of the Ti-O-Ti bond was observed in the FTIR chart of MoS2/MXene, indicating that MoS2 could serve as a reactive oxygen barrier for MXene and shield it from oxidation at high temperatures [32]. The characteristic peak at 2926 cm−1 was due to -CH3 stretching [33]. The characteristic peaks at 1055, 1123, 1541, and 1629 cm−1 were assigned to the stretching vibrations of the C-N, C-O, N-H, and C=O bonds of the EDTA fraction attached to the MoS2 surface, respectively [34]. Furthermore, the peaks at 765 and 596 cm−1 corresponded to out-of-plane C-H bond bending and S-S bond stretching, respectively, due to bond formation between the S atoms of MoS2 and DDC [35]. The characteristic peak at 457 cm−1 was associated with S-Mo stretching, indicating the successful formation of MoS2 nanostructures [29,36]. Apparently, MoS2/MXene possesses more functional groups and surface-active sites, which are potentially more conducive to the effective activation of PDS for the degradation of various refractory organics.
To further investigate the chemical state and composition of the synthesized MoS2/MXene nanocomplexes and to identify the functional group species, an XPS examination was performed. As shown in the measured spectra (Figure 3b), the peaks at 532.2, 456.1, 285.1, 228.4, and 161.8 eV corresponded to O 1s, Ti 2p, C 1s, Mo 3d, and S 2p, respectively [29,37]. The results showed that MoS2 and MXene were successfully combined, which is consistent with the findings of the SEM and FTIR analyses. In addition, in the high-resolution spectrum of C1s (Figure 3c), the peaks at 281.4 eV, 284.8 eV, 286.4 eV, and 288.8 eV corresponded to Ti-C, C-C, C-O, and O-C=O, respectively [38]. In the O1s high-resolution spectrum (Figure 3d), the peaks at 529.7 eV, 532.0 eV, and 533.6 eV were associated with Ti-O, C-OH, and C=O, respectively [39]. All of these characterization findings suggested that there were several functional groups on the composite surface. In the Mo 3d spectrum (Figure 3e), the double peaks had lower binding energies (231.5 and 228.3 eV), which were associated with Mo4+ 3d3/2 and 3d5/2 of MoS2, respectively, and the weak characteristic peaks at 233.1 and 235.6 eV were caused by the presence of trace quantities of unreduced Mo6+ [29,40]. Additionally, the peak detected at 225.3 eV could be attributed to the S 2s signal [20]. In the S 2p spectrum (Figure 3f), the obtained double peaks of S 2p3/2 and S 2p1/2 at ~161.0 and ~162.2 eV were attributed to the S2− of MoS2 [29]. These analytical test findings showed that MoS2 and MoS2/MXene with plenty of edge active sites were successfully prepared.

3.2. Performance Regarding TC Degradation

The effects of different materials and systems, catalyst dosage, PDS consumption, pH, and initial concentration of TC on its degradation were investigated, and the results are shown in Figure 4.
Since the best TC removal effect was obtained when the mass ratio of MoS2 to MXene was 3:1, the composites with this mass ratio were used in subsequent experiments (Figure S1). As shown in Figure S1, when MXene or MoS2 was present alone, the degradation efficiency of TC was limited (41.28% and 55.56%, respectively), while the degradation efficiency of TC became higher (75.30%) when MoS2/MXene was used.
As shown in Figure 4a, the effects of different materials or systems on TC degradation were studied under the same catalyst dosage or PDS concentration conditions. In the presence of PDS alone for 1 h, the degradation efficiency of TC was poor (18.07%), while the addition of MXene or MoS2 accelerated the degradation efficiency of TC (71.82% and 74.06%, respectively). More interestingly, the inclusion of MoS2/MXene significantly accelerated the degradation of TC, which could achieve 100% efficiency at around 50 min. In addition, it can be inferred from Figure S2 that all four systems conformed to the pseudo-first-order kinetic equations when degrading TC, and the parameters are listed in Table S2. These results imply that the vertical development of MoS2 on MXene exposes more active sites, thus enhancing electron mobility across the composites and promoting the rapid degradation of TC [41].
The impacts of different dosages of catalyst and PDS concentrations on TC degradation were also investigated, and the results are displayed in Figure 2b,c. The TC degradation efficiency improved significantly from 57.45% to 100% within 1 h of the process as the catalyst dosage rose from 0.05 g/L to 0.4 g/L. This was attributed to an increase in the rate of production of •OH and other reactive oxygen species (ROS) inside the reaction system as the MoS2/MXene concentration increased. However, despite the increase in catalyst dosage, the degradation efficiency of TC did not improve considerably. This could probably be attributed to the limited number of active sites on the catalyst [42]. Additionally, at a catalyst dosage of 0.4 g/L, the impact of varying PDS concentrations on TC removal was investigated. The rate of TC degradation exhibited a general trend toward improvement as the PDS concentration increased (75.88% 100%). It might be attributed to an increase in the concentration of PDS in the MoS2/MXene/PDS system, which promoted the excitation of S2O82− by MoS2/MXene to form SO4•− while also producing more active species, thereby efficiently catalyzing the breakdown of TC in the solution [43].
Similarly, the effectiveness of TC degradation was also influenced by the initial pH of the reaction system. As shown in Figure 4d, the degradation efficiency of TC under weakly acidic and neutral circumstances was greater than that under acidic and weakly basic conditions. When the initial pH was raised from 3.0 to 5.0, the degradation efficiency of TC rose from 90.56% to 100%, but as the pH was increased further from 7 to 9, the degradation efficiency was inhibited at a rate of 1.09% and 10.36%, respectively. This may be explained by the fact that TC in an aqueous solution occurs as cations under an acidic environment, whereas the MoS2/MXene surface is readily protonated and positively charged, which causes their mutual exclusion and reduces the effectiveness of removing TC [44]. Moreover, the difference in TC degradation efficiency may be attributed to the fact that when the initial pH is too low, a considerable quantity of H+ in the solution will form hydrogen bonds with the O-O group in S2O82−, which negatively impacts PDS activation or is associated with the autolytic degradation of HSO5•− and the generation of SO5•− [45,46]. On the other hand, TC is more easily adsorbed on the surface of the catalyst in weakly acidic or neutral conditions, leading to a higher PDS activation efficiency, and thus, improving the degradation efficiency [47]. In contrast, because TC in an aqueous solution exists as anions under alkaline circumstances and the material’s surface protonation is impaired, its adsorption is diminished, and a considerable amount of OH in the solution competes with TC molecules, thereby limiting their removal [48]. In addition, the impact of the initial concentration of TC was investigated. The results clearly showed that when the initial concentration of TC increased, its degradation efficiency dropped since increasing the initial concentration in the reaction system meant more reactive oxygen species (ROS) were present for catalytic degradation [49].
To further explore the action pattern of •OH production by MoS2/MXene nanoparticles during TC degradation, ESR was employed to detect the DMPO-•OH signal intensity, and the results are shown in Figure 5. It was demonstrated that MoS2/MXene nanoparticles generated significantly more •OH in acidic air than in neutral or basic air (Figure 5a–c). Due to the intermolecular polarization of MoS2, the MoS2/MXene nanocomposite created a piezoelectric potential to stimulate the breakdown of water molecules to form •OH, which was accelerated at higher H+ concentrations [50]. The results clearly demonstrated that MoS2/MXene could produce •OH autonomously in exposed aqueous solutions. Additionally, the intensity peak of DMPO-•OH dropped with an increase in solution pH, showing the strongest signal peak at pH = 4.0 compared to other solution pH values (7.0 and 9.0). The intensity of the DMPO-•OH signal is shown in Figure 5d,e, when MXene and MoS2 were employed as an independent catalyst, respectively. The results revealed that under the same circumstances, the intensity of the DMPO-•OH signal was significantly lower than that of the MoS2/MXene nanocomposite when only MXene or MoS2 was present. Therefore, it could be inferred that the formation of particular heterojunctions between MoS2 and MXene enhanced the nanocomposite’s ability to transmit electrons, thus promoting the generation of •OH [41,51], which potentially stimulated PDS activation and enhanced TC degradation in turn.

3.3. Effect of Inorganic Anions

The impacts of inorganic anions (Cl, NO3, H2PO4, and CO32−), which are frequently present in natural water, were investigated in this work. As shown in Figure 6a, a relatively low concentration (0.5 mM) of Cl could significantly reduce the removal efficiency of TC. Moreover, increasing the concentration of Cl from 0.5 mM to 1.0 mM enhanced the degradation efficiency of TC, although a further increase would restrict the degradation efficiency. This is probably due to the fact that Cl at low concentrations would combine rapidly with SO4•− to form SO42−, which would hinder the degradation of TC in an aqueous solution, whereas higher concentrations might generate other active species such as Cl and Cl2, which might stimulate TC breakdown (Equations (1) and (2)) [52]. However, when the concentration of Cl was gradually increased to 20 mM, various active species were depleted, and some inorganic compounds were generated in the system (Equation (3)) [53]. Therefore, the degrading efficiency of TC was slightly impacted.
Similarly, when the concentration of NO3 was raised from 0.5 mM to 10 mM, the degradation of TC reduced from 26.55% to 14.66% (Figure 6b), which could be attributed to the generation of NO3 (Equation (4)) [54]. However, when the concentration of NO3 was increased further, the degradation rate of TC reduced, which might be related to the conversion of various highly active chemicals into less active or inactive ones in the system.
In contrast, the degradation rate of TC decreased with an increase in the concentrations of H2PO4 and CO32− (Figure 6c,d), which might be due to the scavenging action of these inorganic anions on SO4•− (Equations (5)–(10)) [55]. Overall, as shown in Figure 6e, the order of the inhibition effect of inorganic anions was CO32− > H2PO4 > NO3 > Cl, which was essentially consistent with the conclusions of previous studies [56].
Cl + S O 4 2 S O 4 2 + Cl
Cl + Cl Cl 2
2 Cl + S 2 O 8 2 2 S O 4 2 + Cl 2
NO 3 + S O 4 S O 4 2 + NO 3
CO 3 2 + S O 4 S O 4 2 + CO 3
CO 3 2 + H 2 O OH + HCO 3
HCO 3 + S O 4 S O 4 2 + HCO 3
H 2 PO 4 + S O 4 S O 4 2 + H 2 PO 4
H 2 PO 4 HPO 4 2 + H +
HPO 4 2 + S O 4 S O 4 2 + HPO 4

3.4. Reusability Assessment of MoS2/MXene

In real application areas, the reusability of a catalyst is a crucial factor. The degradation rate of TC in the MoS2/MXene/PDS system served as an indication for repeated experiments using the catalysts. We conducted five tests with repeated cycles using the MoS2/MXene/PDS system. At the end of each cycle, MoS2/MXene was sonicated after being evenly distributed in a certain volume of ultrapure water. After natural sedimentation and centrifugation, the remaining concentration of TC in the supernatant was determined. Then, the preceding stages were repeated until there was no residual TC precipitated from the filtrate. Finally, the completely processed MoS2/MXene was dried overnight in a vacuum oven at 60 °C and ground to powder for the following cycle. As shown in Figure 6f, the removal rate of TC decreased steadily from 100% to 81.92% with an increase in cycle number. It can be inferred that MoS2/MXene is sufficiently reusable for the degradation of organic contaminants. Furthermore, the decreased performance might be due to a minor overflow of catalytic units contained in the catalyst [54]. The catalyst might also adsorb certain degradation intermediates, thus limiting the degradation effectiveness of TC [56]. Therefore, MoS2/MXene is practicable as an effective system for degrading TC by activating PDS, with some potential reusability.

3.5. Activation Mechanism of MoS2/MXene/PDS System

The distinct functions of reactive oxygen species generated in the MoS2/MXene/PDS system were investigated by employing free radical capture experiments. Some common scavengers and their rate constants are listed in Table S3; ethanol (EtOH), tert-butyl alcohol (TBA), and p-benzoquinone (BQ) were chosen as the scavengers of SO4•−, •OH, and O2•−, respectively, for comparing their reaction constants [57]. As shown in Figure 7, different scavengers impacted the degradation efficiency of TC to varying degrees under the 1 mM addition condition. The inhibitory impact of EtOH in the MoS2/MXene/PDS system became apparent within the first 10 min of the reaction, demonstrating that SO4•− produced by the activation of PDS via MoS2/MXene was primarily accountable for the degradation of TC in the primary phase of the reaction. The effect of TBA became increasingly apparent as the reaction continued, suggesting that when most of the PDS was engaged in the MoS2/MXene/PDS system, further degradation of TC depended on •OH generated by MoS2/MXene in the aqueous solution. The effectiveness of TC removal by EtOH and TBA decreased from 100% to 83.89% and 77.59%, respectively, after one hour of the reaction, demonstrating that SO4•− and •OH were partially responsible for TC degradation. Additionally, the degradation efficiency of TC was somewhat reduced after BQ was introduced. After the reaction had proceeded for one hour, the degradation efficiency of TC reduced from 100% to 85.68%, indicating that O2•− was extremely likely involved in the process. According to the results of the free radical capture experiments, it can be inferred that various reactive oxygen species play a significant role in the degradation of TC in the MoS2/MXene/PDS system, whose effect is in the order of •OH > SO4•− > O2•−. Moreover, the non-significant inhibition further indicated the great stability and high efficiency of the MoS2/MXene/PDS system, as well as its potential for having outstanding adsorption capabilities in addition to acting as a catalyst for PDS activation.
To further explore the surface-active sites involved in the reaction, the elemental changes in both fresh and used MoS2/MXene, including valence and content, were characterized utilizing XPS analysis. As shown in Figure 7b, the peaks at 530.9, 458.7, 284.8, 228.4, and 161.9 eV corresponded to O 1s, Ti 2p, C 1s, Mo 3d, and S 2p, respectively. There was no significant change in the location or content of these peaks, indicating that MoS2/MXene possessed good stability (Table S4). Moreover, the high-resolution XPS spectra of C 1s, O1s, Mo 3d, and S 2p are displayed in Figure 7c–f. As seen in Table S5, the Mo6+ concentration increased from 0.07% to 0.86% after the reaction, whereas the Mo4+ content decreased from 5.93% to 4.54%, demonstrating that the slight oxidation of Mo4+ on the surface is a major factor for PDS activation. It is worth noting that the slight decrease in Mo4+ content might be attributed to the interconversion of MO4+ and Mo6+ [58]. Additionally, the Ti-O content increased from 5.61% to 8.57%, demonstrating that MXene and MoS2 were possibly engaged in the activation process of PDS and that their combined action could efficiently speed up this process. Meanwhile, the contents of C-O and C=O groups reduced from 16.52% and 5.93% to 9.88% and 3.99%, respectively. It was further revealed that both C-O and C=O groups were necessary for the activation of PDS and the generation of reactive oxygen species [59]. More interestingly, in the high-resolution XPS pattern of S, a novel signal peak at 164.0 eV was discovered, which was identified as bridging S22− at the apical of MoS2, suggesting a potential linkage reaction between pure MoS2 and further indicating the involvement of MoS2 in the oxidation reaction [60].
A suggested mechanism of PDS activation by MoS2/MXene for effective TC degradation was proposed after combining the above findings and analyses. Due to the combined piezoelectric and semiconducting properties of MoS2, effective removal of organic dyes could be accomplished in the dark (termed piezoelectric catalysis) [61]. A strong hydrogen peak was still detected in the solution after the removal of MoS2, which was likely due to the extended existence of a high concentration of •OH in the aqueous solution containing MoS2 [50]. This finding is also highly consistent with the results of the ESR detection. Therefore, the following equations (Equations (11)–(21)) were derived to account for the ROS generated during the oxidative degradation of TC, as well as the chemical reactions in the MoS2/MXene/PDS system, based on the experimental findings and previous reports [42,50,56,58,59]. Specifically, in the initial phase of oxidative degradation, electrons and holes are formed due to the piezoelectric catalysis and internal polarization of MoS2, which will stimulate the polarization decomposition of water molecules to produce •OH, and the above processes can be accelerated in acidic solutions exposed to air. Then, the high concentration of •OH promotes SO42− to generate SO4•− in the aqueous solution, which rapidly attaches to organic molecules and triggers their oxidative breakdown [62]. Moreover, PDS can produce weak SO4•− and O2•− in an aqueous solution, while MoS2 can expedite this process and permit cyclic redox, considerably enhancing the utilization of the catalyst. Finally, TC is mineralized by a variety of ROS into different small molecules such as H2O and CO2.
Mo S 2 ( H 2 O ) Piezoelectric e + h +
H 2 O + e H + O H
O H + h + OH
OH + S O 4 2 S O 4 + O H
S O 4 + H 2 O OH + S O 4 2 + H +
S 2 O 8 2 + 2 H 2 O H O 2 + 2 S O 4 2 + 3 H +
S 2 O 8 2 + H O 2 S O 4 + S O 4 2 + O 2 + 3 H +
H O 2 H + + O 2
Mo 4 + + 2 S 2 O 8 2 Mo 6 + + 2 S O 4 + 2 S O 4 2
Mo 6 + + 2 S 2 O 8 2 + 2 H 2 O Mo 4 + + 2 S O 5 + 2 S O 4 2 + 4 H +
OH + S O 4 + O 2 + TC C O 2 + H 2 O

3.6. Possible Degradation Pathways

To further clarify the degradation mechanism of TC, liquid chromatography–mass spectrometry (LC-MS) was used to identify the intermediates present in the MoS2/MXene/PDS system at different times. Figure S3 and Table S6 show the m/z values and supposed structure of TC degradation intermediates at different times in the presence of the MoS2/MXene catalyst, and signal peaks with different m/z values can be observed for these intermediates. Based on the LC-MS data and previous reports, three possible degradation pathways of TC were proposed (Figure 8). In pathway I, the original TC molecule (m/z = 446) is converted to intermediate P11 (m/z = 460) via dehydration, which is followed by deamidation and dealkylation to yield intermediate P12 (m/z = 374). It is then converted to intermediates P13 m/z = 302) and P14 (m/z = 242) through oxidative ring-opening reaction and gradual decarbonylation. Finally, these intermediates are further oxidized to small molecules P15 (m/z = 240), P16 (m/z = 220), P17 (m/z = 228), and P18 (m/z = 99). In pathway II, the original TC molecule is first converted to intermediate P21 (m/z = 416) via demethylation. It then undergoes a dealkylation reaction to produce P22 (m/z = 376), and further deamidation and ring-opening reactions produce intermediate P23 (m/z = 338). Finally, it gradually forms small organic molecules via decarbonylation and oxidation reactions. Alternatively, P21 is slowly converted to other small molecules, namely, P24 (m/z = 360), P25 (m/z = 211), and P26 (m/z = 121), via progressive oxidation. In pathway III, the original TC molecule is first dehydrated and converted to intermediate P31 (m/z = 427). It can then undergo demethylation, dealkylation, and ring-opening reactions to produce P32 (m/z = 319). With further oxidation, other organics, namely, P33 (m/z = 279), P34 (m/z = 207), and P35 (m/z = 181), can be formed. Moreover, P31 can form intermediate P36 (m/z = 301) via a ring-opening reaction on the left side and hydroxylation and then intermediate P37 (m/z = 274) via deamidation. Finally, with further oxidative degradation, it can be transformed into intermediates P38 (m/z = 131) and P39 (m/z = 112). When MoS2/MXene is present in the system, TC can eventually be mineralized into inorganic substances, such as CO2 and H2O, with the oxidation of various ROS.

4. Conclusions

In this study, MoS2/MXene nanocomplexes were prepared in a simple one-step hydrothermal approach utilizing Ti3C2 MXene as a substrate, which was employed as an effective catalyst for the activation of PDS to effectively degrade TC. The characterization of MoS2/MXene was carried out using SEM, FTIR, and XPS, and it was observed that a stable 3D intercalation structure was formed through the self-assembly of MoS2 with MXene under solvent heat. This not only resulted in an expanded layer spacing and increased the number of surface-active sites for the catalyst, but it also successfully prevented MXene from being oxidized at high temperatures and inhibited MoS2 from clustering. More interestingly, MoS2/MXene performed better catalytically than individual MXene and MoS2, implying that tight contact between MoS2 and MXene occurred. The •OH produced was crucial in the activation of PDS and the oxidative destruction of TC when MoS2/MXene was present in the aqueous solutions, particularly in acidic solutions. In addition, TC could be completely removed in about 50 min under the conditions of a catalyst dosage of 0.4 g/L, a PDS concentration of 0.4 mM, and pH = 5.0 at room temperature. Despite the interference of various inorganic anions, TC was removed from the MoS2/MXene/PDS system relatively successfully. Through repeated testing, the recyclability of MoS2/MXene was assessed, and the catalyst was shown to have excellent reusability performance. Furthermore, the mechanism of action of the MoS2/MXene/PDS system on the TC degradation process was determined by extrapolating the formation principle of ROS during the reaction process and examining the intermediates and potential degradation pathways with LC-MS. Therefore, this study demonstrated that MoS2/MXene is a promising composite catalyst for the removal and degradation of persistent organic pollutants in water, and it also expanded the application of MoS2-based, MXene-based, and MoS2/MXene composites in catalytic oxidation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14090786/s1, Figure S1 Effect of different materials on the degradation efficiency of TC. Figure S2 Kinetic fitting models for different materials/systems. Figure S3 MS full scan spectra of intermediates in TC degradation in the MoS2/MXene/PDS system by LC-MS. Table S1 The elemental weight percentages and atomic percentages of MoS2/MXene were calculated by EDX spectroscopy. Table S2 The fitting parameters of kinetic models of TC degradation by MoS2/MXene/PDS. Table S3 The reaction rate constants of radicals with the corresponding scavenger. Table S4 The Elemental peak positions and contents of fresh and used MoS2/MXene. Table S5 The group percentages of fresh and used MoS2/MXene. Table S6 Liquid-phase mass spectrometry (LC-MS) analysis of intermediates and possible structural formulae.

Author Contributions

Conceptualization, Y.S. and R.C.; Software, S.Y.; Formal analysis, X.N., M.F. and H.X.; Investigation, X.N.; Data curation, S.Y. and M.F.; Writing—original draft, Y.S. and S.L.; Writing—review & editing, Y.S., R.C. and H.X.; Supervision, Y.S. and R.C.; Funding acquisition, Y.S. and R.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study received financial support from the Natural Science Foundation of Hunan Province (No. 2022JJ50259); the Scientific Research Project of Education Department of Hunan Province (No. 22B0279, No. 22A0193); the Major program Natural Science Foundation of Hunan Province of China (No. 2021JC0001); the Science and Technology Talent Support Project of Hunan Province (No. 2022TJ-Q19); the National Natural Science Foundation of China (No. 52374424); and the Environmental Research Foundation of Hunan Province of China (No. HBKT-2022010).

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gopal, G.; Alex, S.A.; Chandrasekaran, N.; Mukherjee, A. A review on tetracycline removal from aqueous systems by advanced treatment techniques. RSC Adv. 2020, 10, 27081–27095. [Google Scholar] [CrossRef] [PubMed]
  2. Li, N.; Zhou, L.; Jin, X.; Owens, G.; Chen, Z. Simultaneous removal of tetracycline and oxytetracycline antibiotics from wastewater using a ZIF-8 metal organic-framework. J. Hazard. Mater. 2019, 366, 563–572. [Google Scholar] [CrossRef] [PubMed]
  3. Xin, S.; Liu, G.; Ma, X.; Gong, J.; Ma, B.; Yan, Q.; Chen, Q.; Ma, D.; Zhang, G.; Gao, M. High efficiency heterogeneous Fenton-like catalyst biochar modified CuFeO2 for the degradation of tetracycline: Economical synthesis, catalytic performance and mechanism. Appl. Catal. B Environ. 2021, 280, 119386. [Google Scholar] [CrossRef]
  4. Park, J.; Yamashita, N.; Park, C.; Shimono, T.; Takeuchi, D.M.; Tanaka, H. Removal characteristics of pharmaceuticals and personal care products: Comparison between membrane bioreactor and various biological treatment processes. Chemosphere 2017, 179, 347–358. [Google Scholar] [CrossRef] [PubMed]
  5. Cao, J.; Lai, L.; Lai, B.; Yao, G.; Chen, X.; Song, L. Degradation of tetracycline by peroxymonosulfate activated with zero-valent iron: Performance, intermediates, toxicity and mechanism. Chem. Eng. J. 2019, 364, 45–56. [Google Scholar] [CrossRef]
  6. Huang, D.; Zhang, G.; Yi, J.; Cheng, M.; Lai, C.; Xu, P.; Zhang, C.; Liu, Y.; Zhou, C.; Xue, W. Progress and challenges of metal-organic frameworks-based materials for SR-AOPs applications in water treatment. Chemosphere 2021, 263, 127672. [Google Scholar] [CrossRef] [PubMed]
  7. Gao, L.; Mao, Q.; Luo, S.; Cao, L.; Xie, X.; Yang, Y.; Deng, Y.; Wei, Z. Experimental and theoretical insights into kinetics and mechanisms of hydroxyl and sulfate radicals-mediated degradation of sulfamethoxazole: Similarities and differences. Environ. Pollut. 2020, 259, 113795. [Google Scholar] [CrossRef] [PubMed]
  8. Yin, R.; Guo, W.; Wang, H.; Du, J.; Zhou, X.; Wu, Q.; Zheng, H.; Chang, J.; Ren, N. Enhanced peroxymonosulfate activation for sulfamethazine degradation by ultrasound irradiation: Performances and mechanisms. Chem. Eng. J. 2018, 335, 145–153. [Google Scholar] [CrossRef]
  9. Duan, X.; Sun, H.; Kang, J.; Wang, Y.; Indrawirawan, S.; Wang, S. Insights into heterogeneous catalysis of persulfate activation on dimensional-structured nanocarbons. ACS Catal. 2015, 5, 4629–4636. [Google Scholar] [CrossRef]
  10. Anipsitakis, G.P.; Dionysiou, D.D. Radical generation by the interaction of transition metals with common oxidants. Environ. Sci. Technol. 2004, 38, 3705–3712. [Google Scholar] [CrossRef] [PubMed]
  11. Zhou, H.; Lai, L.; Wan, Y.; He, Y.; Yao, G.; Lai, B. Molybdenum disulfide (MoS2): A versatile activator of both peroxymonosulfate and persulfate for the degradation of carbamazepine. Chem. Eng. J. 2020, 384, 123264. [Google Scholar] [CrossRef]
  12. Wang, J.; Wang, S. Activation of persulfate (PS) and peroxymonosulfate (PMS) and application for the degradation of emerging contaminants. Chem. Eng. J. 2018, 334, 1502–1517. [Google Scholar] [CrossRef]
  13. Ji, Y.; Dong, C.; Kong, D.; Lu, J. New insights into atrazine degradation by cobalt catalyzed peroxymonosulfate oxidation: Kinetics, reaction products and transformation mechanisms. J. Hazard. Mater. 2015, 285, 491–500. [Google Scholar] [CrossRef]
  14. Giannakis, S.; Lin, K.-Y.A.; Ghanbari, F. A review of the recent advances on the treatment of industrial wastewaters by Sulfate Radical-based Advanced Oxidation Processes (SR-AOPs). Chem. Eng. J. 2021, 406, 127083. [Google Scholar] [CrossRef]
  15. Ma, D.; Yi, H.; Lai, C.; Liu, X.; Huo, X.; An, Z.; Li, L.; Fu, Y.; Li, B.; Zhang, M. Critical review of advanced oxidation processes in organic wastewater treatment. Chemosphere 2021, 275, 130104. [Google Scholar] [CrossRef] [PubMed]
  16. Sheng, B.; Yang, F.; Wang, Y.; Wang, Z.; Li, Q.; Guo, Y.; Lou, X.; Liu, J. Pivotal roles of MoS2 in boosting catalytic degradation of aqueous organic pollutants by Fe (II)/PMS. Chem. Eng. J. 2019, 375, 121989. [Google Scholar] [CrossRef]
  17. Chen, Y.; Zhang, G.; Liu, H.; Qu, J. Confining free radicals in close vicinity to contaminants enables ultrafast fenton-like processes in the interspacing of MoS2 membranes. Angew. Chem. Int. Ed. 2019, 58, 8134–8138. [Google Scholar] [CrossRef]
  18. Liu, C.; Kong, D.; Hsu, P.-C.; Yuan, H.; Lee, H.-W.; Liu, Y.; Wang, H.; Wang, S.; Yan, K.; Lin, D. Rapid water disinfection using vertically aligned MoS2 nanofilms and visible light. Nat. Nanotechnol. 2016, 11, 1098–1104. [Google Scholar] [CrossRef]
  19. Ai, K.; Ruan, C.; Shen, M.; Lu, L. MoS2 nanosheets with widened interlayer spacing for high-efficiency removal of mercury in aquatic systems. Adv. Funct. Mater. 2016, 26, 5542–5549. [Google Scholar] [CrossRef]
  20. Yang, S.; Liu, Q.; Wan, X.; Jiang, J.; Ai, L. Edge-rich MoS2 nanosheets anchored on layered Ti3C2 MXene for highly efficient and rapid catalytic reduction of 4-nitrophenol and methylene blue. J. Alloys Compd. 2022, 891, 161900. [Google Scholar] [CrossRef]
  21. Li, X.; Lv, X.; Sun, X.; Yang, C.; Zheng, Y.-Z.; Yang, L.; Li, S.; Tao, X. Edge-oriented, high-percentage 1T′-phase MoS2 nanosheets stabilize Ti3C2 MXene for efficient electrocatalytic hydrogen evolution. Appl. Catal. B Environ. 2021, 284, 119708. [Google Scholar] [CrossRef]
  22. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional nanocrystals produced by exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed]
  23. Seh, Z.W.; Fredrickson, K.D.; Anasori, B.; Kibsgaard, J.; Strickler, A.L.; Lukatskaya, M.R.; Gogotsi, Y.; Jaramillo, T.F.; Vojvodic, A. Two-dimensional molybdenum carbide (MXene) as an efficient electrocatalyst for hydrogen evolution. ACS Energy Lett. 2016, 1, 589–594. [Google Scholar] [CrossRef]
  24. Habib, T.; Zhao, X.; Shah, S.A.; Chen, Y.; Sun, W.; An, H.; Lutkenhaus, J.L.; Radovic, M.; Green, M.J. Oxidation stability of Ti3C2Tx MXene nanosheets in solvents and composite films. NPJ 2D Mater. Appl. 2019, 3, 8. [Google Scholar] [CrossRef]
  25. Navitski, I.; Ramanaviciute, A.; Ramanavicius, S.; Pogorielov, M.; Ramanavicius, A. MXene-Based Chemo-Sensors and Other Sensing Devices. Nano-Mater. 2024, 14, 447. [Google Scholar] [CrossRef] [PubMed]
  26. Adomavičiūtė-Grabusovė, S.; Ramanavičius, S.; Popov, A.; Šablinskas, V.; Gogotsi, O.; Ramanavičius, A. Selective enhancement of SERS spectral bands of salicylic acid adsorbate on 2D Ti3C2Tx-based MXene film. Chemosensors 2021, 9, 223. [Google Scholar] [CrossRef]
  27. Sarycheva, A.; Gogotsi, Y. Raman Spectroscopy Analysis of the Structure and Surface Chemistry of Ti3C2Tx MXene. Chem. Mater. 2020, 32, 3480–3488. [Google Scholar] [CrossRef]
  28. Chen, R.; Cheng, Y.; Wang, P.; Wang, Y.; Wang, Q.; Yang, Z.; Tang, C.; Xiang, S.; Luo, S.; Huang, S. Facile synthesis of a sandwiched Ti3C2Tx MXene/nZVI/fungal hypha nanofiber hybrid membrane for enhanced removal of Be (II) from Be (NH2)2 complexing solutions. Chem. Eng. J. 2021, 421, 129682. [Google Scholar] [CrossRef]
  29. Kumar, N.; Fosso-Kankeu, E.; Ray, S.S. Achieving controllable MoS2 nanostructures with increased interlayer spacing for efficient removal of Pb (II) from aquatic systems. ACS Appl. Mater. Interfaces 2019, 11, 19141–19155. [Google Scholar] [CrossRef] [PubMed]
  30. Gogotsi, Y.; Anasori, B. The rise of MXenes. ACS Nano 2019, 13, 8491–8494. [Google Scholar] [CrossRef] [PubMed]
  31. Othman, Z.; Sinopoli, A.; Mackey, H.R.; Mahmoud, K.A. Efficient Photocatalytic Degradation of Organic Dyes by AgNPs/TiO2/Ti3C2Tx MXene Composites under UV and Solar Light. ACS Omega 2021, 6, 33325–33338. [Google Scholar] [CrossRef] [PubMed]
  32. Mao, X.; Zou, Y.; Xu, F.; Sun, L.; Chu, H.; Zhang, H.; Zhang, J.; Xiang, C. Three-dimensional self-supporting Ti3C2 with MoS2 and Cu2O nanocrystals for high-performance flexible supercapacitors. ACS Appl. Mater. Interfaces 2021, 13, 22664–22675. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, Y.; Zheng, K.; Zhan, W.; Huang, L.; Liu, Y.; Li, T.; Yang, Z.; Liao, Q.; Chen, R.; Zhang, C. Highly effective stabilization of Cd and Cu in two different soils and improvement of soil properties by multiple-modified biochar. Ecotoxicol. Environ. Saf. 2021, 207, 111294. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, X.; Yang, S.; Yu, B.; Tan, Q.; Zhang, X.; Cong, H. Advanced modified polyacrylonitrile membrane with enhanced adsorption property for heavy metal ions. Sci. Rep. 2018, 8, 1260. [Google Scholar] [CrossRef]
  35. Li, L.; Qu, Q.; Bai, W.; Yang, F.; Chen, Y.; Zhang, S.; Ding, Z. Sodium diethyldithiocarbamate as a corrosion inhibitor of cold rolled steel in 0.5 M hydrochloric acid solution. Corros. Sci. 2012, 59, 249–257. [Google Scholar] [CrossRef]
  36. Wang, Z.; Mi, B. Environmental applications of 2D molybdenum disulfide (MoS2) nanosheets. Environ. Sci. Technol. 2017, 51, 8229–8244. [Google Scholar] [CrossRef] [PubMed]
  37. Hwang, S.K.; Kang, S.-M.; Rethinasabapathy, M.; Roh, C.; Huh, Y.S. MXene: An emerging two-dimensional layered material for removal of radioactive pollutants. Chem. Eng. J. 2020, 397, 125428. [Google Scholar] [CrossRef]
  38. Liu, C.; Zhang, H.; Li, R.; Li, X.; Tang, P.; Wang, Y.; Yang, B.; Qiao, Z.; Yang, G. Laser triggered exothermic chemical reaction in Au nanoparticle@Ti3C2 MXene membrane: A route toward efficient light to high-temperature pulse conversion. Chem. Eng. J. 2021, 420, 127672. [Google Scholar] [CrossRef]
  39. Tang, T.; Yin, Z.; Chen, J.; Zhang, S.; Sheng, W.; Wei, W.; Xiao, Y.; Shi, Q.; Cao, S. Novel pn heterojunction Bi2O3/Ti3+-TiO2 photocatalyst enables the complete removal of tetracyclines under visible light. Chem. Eng. J. 2021, 417, 128058. [Google Scholar] [CrossRef]
  40. Liu, J.; Jia, Z.; Zhou, W.; Liu, X.; Zhang, C.; Xu, B.; Wu, G. Self-assembled MoS2/magnetic ferrite CuFe2O4 nanocomposite for high-efficiency microwave absorption. Chem. Eng. J. 2022, 429, 132253. [Google Scholar] [CrossRef]
  41. Jiao, S.; Liu, L. Friction-induced enhancements for photocatalytic degradation of MoS2@Ti3C2 nanohybrid. Ind. Eng. Chem. Res. 2019, 58, 18141–18148. [Google Scholar] [CrossRef]
  42. Wang, Y.; Ji, Q.; Xu, J.; Wan, J.; Wang, L. Activation of peroxydisulfate using N-doped carbon-encapsulated Ni species for efficient degradation of tetracycline. Sep. Purif. Technol. 2021, 276, 119369. [Google Scholar] [CrossRef]
  43. He, D.; Zhu, K.; Huang, J.; Shen, Y.; Lei, L.; He, H.; Chen, W. N, S co-doped magnetic mesoporous carbon nanosheets for activating peroxymonosulfate to rapidly degrade tetracycline: Synergistic effect and mechanism. J. Hazard. Mater. 2022, 424, 127569. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, P.; Yap, P.-S.; Lim, T.-T. C-N-S tridoped TiO2 for photocatalytic degradation of tetracycline under visible-light irradiation. Appl. Catal. A Gen. 2011, 399, 252–261. [Google Scholar] [CrossRef]
  45. Du, W.; Zhang, Q.; Shang, Y.; Wang, W.; Li, Q.; Yue, Q.; Gao, B.; Xu, X. Sulfate saturated biosorbent-derived Co-S@NC nanoarchitecture as an efficient catalyst for peroxymonosulfate activation. Appl. Catal. B Environ. 2020, 262, 118302. [Google Scholar] [CrossRef]
  46. He, J.; Wan, Y.; Zhou, W. ZIF-8 derived Fe-N coordination moieties anchored carbon nanocubes for efficient peroxymonosulfate activation via non-radical pathways: Role of FeNx sites. J. Hazard. Mater. 2021, 405, 124199. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, Y.; Zhang, H.; Chen, L. Ultrasound enhanced catalytic ozonation of tetracycline in a rectangular air-lift reactor. Catal. Today 2011, 175, 283–292. [Google Scholar] [CrossRef]
  48. Yang, L.; Shengwang, G.; Lijun, W.; Jianchao, Z.; Hong, G.; Xunfeng, X. Tetracycline degradation in aqueous solution by porous MoS2/g-C3N4. Chin. J. Environ. Eng. 2019, 13, 818–825. [Google Scholar]
  49. Gao, Y.; Chen, Z.; Zhu, Y.; Li, T.; Hu, C. New insights into the generation of singlet oxygen in the metal-free peroxymonosulfate activation process: Important role of electron-deficient carbon atoms. Environ. Sci. Technol. 2019, 54, 1232–1241. [Google Scholar] [CrossRef] [PubMed]
  50. Wu, J.M.; Sun, Y.-G.; Chang, W.-E.; Lee, J.-T. Piezoelectricity induced water splitting and formation of hydroxyl radical from active edge sites of MoS2 nanoflowers. Nano Energy 2018, 46, 372–382. [Google Scholar] [CrossRef]
  51. Hou, W.; Sun, Y.; Zhang, Y.; Wang, T.; Wu, L.; Du, Y.; Zhong, W. Mixed-dimensional heterostructure of few-layer MXene based vertical aligned MoS2 nanosheets for enhanced supercapacitor performance. J. Alloys Compd. 2021, 859, 157797. [Google Scholar] [CrossRef]
  52. Ao, X.; Liu, W.; Sun, W.; Cai, M.; Ye, Z.; Yang, C.; Lu, Z.; Li, C. Medium pressure UV-activated peroxymonosulfate for ciprofloxacin degradation: Kinetics, mechanism, and genotoxicity. Chem. Eng. J. 2018, 345, 87–97. [Google Scholar] [CrossRef]
  53. Peng, L.; Shang, Y.; Gao, B.; Xu, X. Co3O4 anchored in N, S heteroatom co-doped porous carbons for degradation of organic contaminant: Role of pyridinic N-Co binding and high tolerance of chloride. Appl. Catal. B Environ. 2021, 282, 119484. [Google Scholar] [CrossRef]
  54. Gao, Y.; Zou, D. Efficient degradation of levofloxacin by a microwave-3D ZnCo2O4/activated persulfate process: Effects, degradation intermediates, and acute toxicity. Chem. Eng. J. 2020, 393, 124795. [Google Scholar] [CrossRef]
  55. Hu, L.; Zhang, G.; Liu, M.; Wang, Q.; Wang, P. Enhanced degradation of Bisphenol A (BPA)by peroxymonosulfate with Co3O4-Bi2O3 catalyst activation: Effects of pH, inorganic anions, and water matrix. Chem. Eng. J. 2018, 338, 300–310. [Google Scholar] [CrossRef]
  56. Liu, Z.; Gao, Z.; Wu, Q. Activation of persulfate by magnetic zirconium-doped manganese ferrite for efficient degradation of tetracycline. Chem. Eng. J. 2021, 423, 130283. [Google Scholar] [CrossRef]
  57. Pu, M.; Wan, J.; Zhang, F.; Brusseau, M.L.; Ye, D.; Niu, J. Insight into degradation mechanism of sulfamethoxazole by metal-organic framework derived novel magnetic Fe@C composite activated persulfate. J. Hazard. Mater. 2021, 414, 125598. [Google Scholar] [CrossRef]
  58. Zhu, Y.; Wang, F.; Zhou, B.; Chen, H.; Yuan, R.; Zhang, Y.; Geng, H.; Liu, Y.; Wang, H. Photo-assisted Fe2+ modified molybdenum disulfide activated potassium persulfate to degrade sulfadiazine: Insights into the degradation pathway and mechanism from density functional theory. Chem. Eng. J. 2022, 435, 134904. [Google Scholar] [CrossRef]
  59. Yang, Y.; Ji, W.; Li, X.; Lin, H.; Chen, H.; Bi, F.; Zheng, Z.; Xu, J.; Zhang, X. Insights into the mechanism of enhanced peroxymonosulfate degraded tetracycline using metal organic framework derived carbonyl modified carbon-coated Fe0. J. Hazard. Mater. 2022, 424, 127640. [Google Scholar] [CrossRef] [PubMed]
  60. Cao, P.; Peng, J.; Liu, S.; Cui, Y.; Hu, Y.; Chen, B.; Li, J.; Zhai, M. Tuning the composition and structure of amorphous molybdenum sulfide/carbon black nanocomposites by radiation technique for highly efficient hydrogen evolution. Sci. Rep. 2017, 7, 16048. [Google Scholar] [CrossRef] [PubMed]
  61. Liu, S.; Jing, B.; Nie, C.; Ao, Z.; Duan, X.; Lai, B.; Shao, Y.; Wang, S.; An, T. Piezoelectric activation of peroxymonosulfate by MoS2 nanoflowers for the enhanced degradation of aqueous organic pollutants. Environ. Sci. Nano 2021, 8, 784–794. [Google Scholar] [CrossRef]
  62. Yao, Y.; Cai, Y.; Lu, F.; Wei, F.; Wang, X.; Wang, S. Magnetic recoverable MnFe2O4 and MnFe2O4-graphene hybrid as heterogeneous catalysts of peroxymonosulfate activation for efficient degradation of aqueous organic pollutants. J. Hazard. Mater. 2014, 270, 61–70. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the synthesis of MoS2/MXene.
Figure 1. Schematic diagram of the synthesis of MoS2/MXene.
Nanomaterials 14 00786 g001
Figure 2. SEM images of (a) MXene, (b,c) MoS2, and (df) MoS2/MXene at different magnifications, and (g) EDX electron microscope scan images and elemental distribution.
Figure 2. SEM images of (a) MXene, (b,c) MoS2, and (df) MoS2/MXene at different magnifications, and (g) EDX electron microscope scan images and elemental distribution.
Nanomaterials 14 00786 g002
Figure 3. (a) FTIR spectra of MXene, MoS2, and MoS2/MXene; (b) XPS survey spectra of MoS2/MXene, (c) C 1s, (d) O 1s, and (e) Mo 3d; and (f) S 2p spectra of MoS2/MXene.
Figure 3. (a) FTIR spectra of MXene, MoS2, and MoS2/MXene; (b) XPS survey spectra of MoS2/MXene, (c) C 1s, (d) O 1s, and (e) Mo 3d; and (f) S 2p spectra of MoS2/MXene.
Nanomaterials 14 00786 g003
Figure 4. Effects of different factors on TC degradation: (a) material system; (b) catalyst dosage; (c) PDS concentration; (d) pH; and (e) TC initial concentration.
Figure 4. Effects of different factors on TC degradation: (a) material system; (b) catalyst dosage; (c) PDS concentration; (d) pH; and (e) TC initial concentration.
Nanomaterials 14 00786 g004
Figure 5. Intensity of DMPO-•OH signals in the presence of (ac) MoS2/MXene, pH = 4, 7, or 9; (d) Ti3C2 MXene, pH = 4; and (e) MoS2, pH = 4.
Figure 5. Intensity of DMPO-•OH signals in the presence of (ac) MoS2/MXene, pH = 4, 7, or 9; (d) Ti3C2 MXene, pH = 4; and (e) MoS2, pH = 4.
Nanomaterials 14 00786 g005
Figure 6. (ae) Effects of different inorganic anions on the MoS2/MXene/PDS system, and (f) the degradation of TC in the recycling study of the MoS2/MXene/PDS system.
Figure 6. (ae) Effects of different inorganic anions on the MoS2/MXene/PDS system, and (f) the degradation of TC in the recycling study of the MoS2/MXene/PDS system.
Nanomaterials 14 00786 g006
Figure 7. (a) Effects of different free radical bursting agents; (b) the XPS survey spectra of fresh and used MoS2/MXene, (c) C 1s, (d) O 1s, and (e) Mo 3d; and (f) S 2p spectra of fresh and used MoS2/MXene.
Figure 7. (a) Effects of different free radical bursting agents; (b) the XPS survey spectra of fresh and used MoS2/MXene, (c) C 1s, (d) O 1s, and (e) Mo 3d; and (f) S 2p spectra of fresh and used MoS2/MXene.
Nanomaterials 14 00786 g007
Figure 8. Possible degradation pathways of TC.
Figure 8. Possible degradation pathways of TC.
Nanomaterials 14 00786 g008
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

Song, Y.; Chen, R.; Li, S.; Yu, S.; Ni, X.; Fang, M.; Xie, H. Enhancement of Peroxydisulfate Activation for Complete Degradation of Refractory Tetracycline by 3D Self-Supported MoS2/MXene Nanocomplex. Nanomaterials 2024, 14, 786. https://doi.org/10.3390/nano14090786

AMA Style

Song Y, Chen R, Li S, Yu S, Ni X, Fang M, Xie H. Enhancement of Peroxydisulfate Activation for Complete Degradation of Refractory Tetracycline by 3D Self-Supported MoS2/MXene Nanocomplex. Nanomaterials. 2024; 14(9):786. https://doi.org/10.3390/nano14090786

Chicago/Turabian Style

Song, Yuxia, Runhua Chen, Shihai Li, Shali Yu, Xiaoli Ni, Minglong Fang, and Hanyun Xie. 2024. "Enhancement of Peroxydisulfate Activation for Complete Degradation of Refractory Tetracycline by 3D Self-Supported MoS2/MXene Nanocomplex" Nanomaterials 14, no. 9: 786. https://doi.org/10.3390/nano14090786

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