Next Article in Journal
Comparative Metabolomics to Unravel the Biochemical Mechanism Associated with Rancidity in Pearl Millet (Pennisetum glaucum L.)
Previous Article in Journal
Effects of Abscisic Acid on the Physiological and Biochemical Responses of Saccharina japonica Under High-Temperature Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Peroxymonosulfate Activation by Rice-Husk-Derived Biochar (RBC) for the Degradation of Sulfamethoxazole: The Key Role of Hydroxyl Groups

1
School of Resources and Environmental Engineering, Hefei University of Technology, Hefei 230009, China
2
Key Laboratory of Nanominerals and Pollution Control of Higher Education Institutes, Hefei University of Technology, Hefei 230009, China
3
CAS Key Laboratory of Urban Pollutant Conversion, Department of Environmental Science and Engineering, University of Science and Technology of China, Hefei 230026, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(21), 11582; https://doi.org/10.3390/ijms252111582
Submission received: 14 July 2024 / Revised: 7 August 2024 / Accepted: 17 August 2024 / Published: 29 October 2024
(This article belongs to the Section Biochemistry)

Abstract

:
In this work, rice-husk-derived biochar (RBC) was synthesized by using simple one-step pyrolysis strategies and served as catalysts to activate peroxymonosulfate (PMS) for degrading sulfamethoxazole (SMX). When the annealing temperature (T) = 800 °C, RBC800 exhibits the typical hardwood structure with several micropores and mesoporous. Furthermore, RBC800 obtains more defect sites than RBC600, RBC700, and RBC900. In the RBC800/PMS system, the removal rate of the SMX reached 92.0% under optimal conditions. The kinetic reaction rate constant (kobs) of the RBC800/PMS system was 0.009 min−1, which was about 1.50, 1.28, and 4.50 times that of the RBC600/PMS (kobs = 0.006 min−1), RBC700/PMS (kobs = 0.007 min−1), and RBC900/PMS (kobs = 0.002 min−1) systems, respectively. In the RBC800/PMS system, sulfate radical (SO4•−) is the main active species. Compared with other active sites, the hydroxyl group (C-OH) on the surface of RBC800 interacts more strongly with PMS, which is more likely to promote the stretching of the O-O bond of the PMS, thus breaking into the activated state and significantly reducing the activation energy required for reaction. The degradation intermediates of SMX were speculated, and the toxicity analysis was conducted. Generally, this work reveals in depth the interaction between reactive sites of biochar-based catalysts and PMS at the molecular level.

1. Introduction

Recently, pharmaceuticals and personal care products (PPCPs) have been widely adopted to treat a variety of human and animal diseases [1]. Among them, sulfamethoxazole (SMX) as a typical PPCPs can effectively prevent coccidiosis, diarrhea, gastroenteritis, and other bacterial diseases [2]. SMX is first consumed by humans or animals and enters the wastewater treatment system with excreta. However, most conventional treatment processes cannot degrade SMX effectively, so residual SMX enters the natural water environment with treated municipal wastewater [3]. Residual SMX in aquatic environments may increase the resistance of pathogenic microorganisms, causing negative effects on aquatic life [4]. Therefore, it is important to develop an effective method for eliminating residual SMX in the aquatic environment.
The hydroxyl radical-based advanced oxidation processes (HR-AOPs) are the effective method for degrading emerging contaminants, and their mechanisms are largely dependent on the hydroxyl radical (OH) [5]. However, HR-AOPs also have many shortcomings, such as the difficulty of transporting and storing hydrogen peroxide (H2O2), a narrow pH application range, and higher energy consumption [6]. Recently, persulfate-based advanced oxidation processes (PS-AOPs) have been favored by researchers due to higher oxidizability and wider applicability under the environments (pH = 2.0–8.0) [5]. There are lots of methods for peroxydisulfate (PDS) and peroxymonosulfate (PMS) activation to generate reactive oxygen species (ROSs), such as ultraviolet irradiation, heat, and transition metals [7]. However, these activation methods often have problems such as extensive energy consumption, high cost, metal leaching, etc., which limits their application prospects in actual wastewater [8].
Biochar (BC) has become a research hotspot recently [9]. Biomass feedstock, such as sludge, corncob, wood chips and shrimp shell, is rich and cheap [10]. Besides, BC has abundant functional groups as well as well-developed porous structure, which obtains significant advantages for the synthesis of carbonaceous catalysts [11]. Qi et al. reported that the Enteromorpha based graphene-like biochar (EGB) was synthesized as a PDS activator for the SMX removal [12]. Liu et al. reported that boron-doped graphitic porous biochar (B-KBC) was synthesized and used as catalysts to activate PDS for removing SMX [13]. Besides, Zhao et al. systematically reported the research progress of the PMS activation by biochar-based catalysts [11]. Although some studies have been carried out on the degradation of micropollutants via activating PS by pristine BC, the interaction mechanism between surface active sites of BC and PS is still unclear. In addition, the pyrolysis temperature (T) largely determines the structure and physicochemical properties of catalysts, while the specific surface area (SSAs), micropore structure and graphitization degree of BC will affect the removal rate of organic micropollutants in PS-AOPs [11]. Herein, the structure-activity relationship between the physicochemical properties of catalysts and the catalytic activity needs to be further revealed.
Herein, raw biochar (RBCT) derived from rice husk was prepared as a PMS activator for the elimination of SMX. The difference of physicochemical properties of RBCT at different pyrolysis temperatures was investigated, and the structure-activity relationship between physicochemical properties and catalytic activity was established. Then, the ROSs in the RBC800/PMS system were identified. The degradation pathway of SMX was predicted, and the interaction mechanism between active sites and PMS was revealed by density functional theory (DFT) calculation. The study proposes a deep understanding of PMS activation mechanism and demonstrates great potential of the biochar-based catalysts toward sewage treatment.

2. Results and Discussion

2.1. Characterization of RBC800

The morphology of RBCT were observed by scanning electron microscopy (SEM). As displayed in Figure 1a–d, RBC600 shows the clumpy structure and incomplete porous structure. When the pyrolysis temperature is 700 °C, RBC700 shows a plate structure, and the pore structure has initially developed. RBC800 derived from rice husks exhibits a typical hardwood structure, with lots of micropores and mesoporous. When the pyrolysis temperature is 900 °C, the carbon network collapses. It can be seen that the porous structure of RBC is relatively developed when the pyrolysis temperature is 800 °C, which is conducive to the catalytic oxidation process. However, excessive pyrolysis temperature (T ≥ 900 °C) may destroy the well-developed porous structure.
As depicted in Figure 1e,f, energy-dispersive X-ray spectroscopy (EDS) analyzed the elemental content of RBC800, in which C content accounted for 65.3%, O content accounted for 24.0%, Si content accounted for 5.5%, N content accounted for 0.2% and Fe content accounted for 5.0%. Besides, Fe element is uniformly distributed on RBC800 surface. These results indicate that rice husk-derived RBC800 naturally contains a small amount of Fe element without additional iron source.
Furthermore, the ultrastructure of crystal lattice of RBC800 was observed in the high-resolution transmission electron microscope (HRTEM) and selected area electron diffraction (SAED) pattern, as depicted in Figure 2a,b. RBC800 is composed of graphitized (sp2C) and disordered carbon (sp3C) structures, and the degree of graphitization and disordered of the catalyst need to be further determined by Raman spectroscopy. The Miller indices of (111) and (220) is consistent with the X-ray diffraction (XRD) results.
XRD patterns of RBCT are depicted in Figure 3a. Obviously, the pronounced diffraction peak at 21.60° was attributed to the (111) plane of SiO2 [14]. Two other weak peaks at 35.6° and 56.2°, corresponding to the (220) and (331) planes, agree well with the crystalline phase of calcite (JCPDS 27-0605).
FT-IR spectra show the functional groups of RBCT (Figure 3b). The peaks at 1250 cm−1, 1622 cm−1, and 3419 cm−1 are respectively attributed to the stretching vibration of C-O, C=O and -OH groups [15]. Xin et al. reported that the content of functional groups can be qualitatively judged by the absorption peak intensity [6]. These results show that the annealing temperature has a crucial effect on the formation of oxygen-containing functional groups. The content of C=O and -OH groups on RBC800 surface is significantly higher than that on RBC600, RBC700 and RBC900, while the C=O and -OH groups on the surface of catalyst can promote electron transfer processes and activate PMS to generate ROSs [16].
N2 adsorption-desorption isotherms (Figure 3c) indicate the specific surface areas (SSAs) of RBC800 (194.86 m2⋅g−1) are larger than that of RBC600 (65.00 m2⋅g−1), RBC700 (124.11 m2⋅g−1) and RBC900 (102.71 m2⋅g−1) (Table S1). The appearance of both D-band and G-band of RBCT reveals the co-presence of disordered and crystalline graphite structures [17] (Figure 3d). Furthermore, the ratio of ID/IG is the key parameter to indicate the defective degree of RBCT [18,19]. The ID/IG value was obtained by calculating the intensity of ration of D peak to G peak [13]. The ID/IG is 0.67, 0.98, 1.07, and 0.88 for RBC600, RBC700, RBC800, and RBC900, respectively. The result demonstrates that RBC800 has obtained abundant defects (vacancy and edge defects) during pyrolysis, which is conducive to catalytic oxidation [20].

2.2. Catalytic Oxidation of SMX

The adsorption process accords with Langmuir model, suggesting the monolayer adsorption of SMX on RBC800 surface (Figure S1). In the pure RBCT system, 23.0%, 27.7%, 32.0%, and 29.0% of SMX could be adsorbed onto RBC600, RBC700, RBC800, and RBC900 within 200 min, respectively (Figure 4a). RBC800 has higher SMX adsorption capacities, due to its larger SSAs and developed porous structure [21].
In the RBC800/PMS system, the degradation efficiency of SMX was 67.3% within 40 min, indicating that BC can activate PMS to a certain extent [22]. In the presence of PMS, the removal rate of RBC600, RBC700, RBC800, and RBC900 within 200 min was 77.0%, 80.0%, 92.0%, and 79.0%, respectively (Figure 4b). The removal process of SMX followed a first-order kinetics behavior, and kobs is the reaction rate constant. The kobs of RBC800/PMS system was 0.009 min−1, being about 1.50, 1.28, and 4.50 times that of RBC600/PMS, RBC700/PMS, and RBC900/PMS systems, respectively (Table S2).
It indicated that the annealing temperature (T) can affect the catalytic performances of catalysts. Within a certain range (T < 900 °C), the increase of pyrolysis temperature is conducive to the development of micropores and promotes the conversion of the organic phase with poor crystallinity into graphitic carbon structure, thus enhancing the catalytic performance of catalysts [23]. However, when the pyrolysis temperature is too high (T = 900 °C), the collapse of the carbon skeleton causes a partial loss of the defects, leading to a decrease in the catalytic activity of RBCT [21]. The PPCPs degradation rates in different systems were studied in Table S3. Besides, the effect of various factors on SMX degra-dation were investigated (Figure S2). The point of zero charge (pHpzc) of RBC800 was displayed in Figure S3. To explore the catalytic performance of RBC800 in depth, the residual PMS concentration was detected by the ABTs colorimetric method. The decomposition rate of PMS in the RBC800/PMS system was 74.9% within 200 min (Figure S4). It was suggested that the favorable degradation rate of SMX in the RBC800/PMS system could be due to the rapid decomposition of PMS.
Co-existing ions can affect the SMX degradation by interacting with ROSs [24]. As shown in Figure S5, typical anions (HCO3, Cl, H2PO4) and humic acid (HA) may restrain the SMX elimination. When 5.0 mM and 10.0 mM Cl were present in the RBC800/PMS system, SMX elimination rate decreased from 92.0% to 73.0% and 58.0%, respectively, within 200 min. The kobs decreased from 0.009 min−1 to 0.003 min−1 and 0.002 min−1, which was due to the generation of Cl, Cl2, and HOCl (Equations (1)–(5)) [25]. When increasing the H2PO4 concentration to 5.0 mM and 10.0 mM, the SMX removal rate decreased to 70.0% and 57.7%, respectively. Similarly, when HCO3 concentration was increased from 0 to 5.0 mM and 10.0 mM, the SMX elimination rate decreased from 92.0% to 66.3% and 60.0%, respectively. These results reveal that H2PO4 and HCO3 possess a quenching effect on reactive radicals (Equations (6)–(9)) [26]. When 5.0 mg⋅L−1 and 10.0 mg⋅L−1 of HA was added into the RBC800/PMS system, kobs decreased from 0.009 min−1 to 0.003 min−1 and 0.002 min−1, respectively. The π-π stacking effect of HA can lead to competing adsorption between HA and PMS, thus hinder the production of active species [27].
C l + S O 4 · S O 4 2 + C l
H S O 5 + C l S O 4 2 + H O C l
H S O 5 + 2 C l + H + S O 4 2 + C l 2 + H 2 O
C l + · O H H O C l ·
H O C l · + H + C l + H 2 O
O H + H 2 P O 4 H 2 P O 4 · + O H
S O 4 · + H 2 P O 4 H 2 P O 4 · + S O 4 2
H C O 3 + · O H H 2 O + C O 3 2
H C O 3 + S O 4 · S O 4 2 + H C O 3 ·
To study the universal applicability of RBC800 in real waterbody, more tests were conducted in various water matrices. The characteristics of different water matrices were listed in Table S4. The removal rate of SMX in tap water (79.0%) was lower than that obtained for deionized water (92.0%) (Figure S6a). Besides, the elimination rate decreased in river water, which might be due to the presence of certain level of ions in the river water [28].
To explore the reusability of RBC800, five consecutive degradation tests were performed (Figure S6b). After four cycles, the elimination rate of SMX within 200 min decreased from 92.0% to 66.3%. This could be due to the sedimentation of intermediates on the surface of RBC800, which occupied the defective sites for activating PMS [15]. To recover the porous structure of passivated RBC800, a thermal treatment (annealing at 450 °C under N2 flow for 2 h) was applied. The result showed that the catalytic activity was partially recovered with 74.0% within 200 min in the RBC800/PMS system.
To further explore the wide applicability of RBC800, the degradation experiments of different pollutants were carried out (Figure S6c). In the RBC800/PMS system, the degradation rates of CEX, CIP, CMP and TCS within 200 min were 87.3%, 91.1%, 90.2% and 89.6%, respectively, and the mineralization rates were 67.4%, 72.3%, 73.7% and 76.1%, respectively. These results showed that the admirable universality of RBC800 as an efficient PMS activator to remove a broad array of typical PPCPs. As displayed in Figure S6d, the RBC800/PMS system had a satisfactory mineralization performance on SMX with a TOC elimination rate of 79.0% within 200 min.

2.3. Mechanism Discussion

2.3.1. Identification of ROSs

Quenching tests were conducted to first to determine the dominant ROSs responsible for SMX degradation process [13,29]. Methanol (MeOH) was regarded as a scavenger of OH ((1.6–7.7) × 107 M−1s−1) and SO4•− ((1.2–2.8) × 107 M−1s−1) [30]. In contrast, Tert-butanol (TBA) was used as a scavenger to quench OH (6.0 × 108 M−1s−1) [31]. p-benzoquinone (p-BQ) was used to inhibit O2•− (9.6 × 108 M−1s−1) [24]. The SMX elimination rates decreased from 92.0% to 37.0%, 70.0%, 84.0%, and 52.0% within 200 min after adding MeOH (0.5 M), TBA (0.5 M), p-BQ (20.0 mM), and FFA (0.5 M), respectively (Figure 5). The results showed that MeOH had an obvious inhibitory effect on SMX degradation in the RBC800/PMS system, while TBA had a mild inhibitory effect on SMX degradation. Therefore, SO4•− may be the main ROSs, leading the degradation process of SMX. In addition, O2•− had little effect on the degradation of SMX, while OH and 1O2 were also involved in the degradation of SMX.
ESR was used to monitor the presence of main active species in the RBC800/PMS system using 2,2,6,6-tetramethyl-4-piperidinol (TEMP) and 5,5-dimethyl-1-pyrroline N-oxide (DMPO) as spin-trapping agents [32]. When RBC800, PMS and TEMP were added to the system, a typical triplet signal with the intensity ratio of 1:1:1 verified the existence of 1O2. The self-decomposition of PMS could produce a small amount of 1O2 [33] (Figure 6a). After adding SMX to the RBC800/PMS system, the peak intensity of TEMP-1O2 was significantly weakened, which indicated that 1O2 played a role in the degradation process of SMX. A spectrum with seven main peaks belonging to DMPO-X was observed in the RBC800/PMS system, suggesting that RBC800 activated PMS to produce SO4•− and OH (Figure 6b). After adding SMX to the RBC800/PMS system, DMPO-OH and DMPO-SO4•− signals were significantly weakened, which indicated that OH and SO4•− played a vital role in the degradation process of SMX. According to the quenching experiment and ESR test results, 1O2, SO4•− and OH were produced in the RBC800/PMS system, and SO4•− played a dominant role in the removal of SMX.

2.3.2. Reaction Mechanism

PMS activation mechanism was further investigated in this work. First, OH and SO4•− generated from the destruction of the O-O bond of the PMS by free-flowing π electrons on the sp2 hybrid carbon of the RBC800 (Equations (10) and (11)). Secondly, 1O2 could be produced by the self-decomposition of PMS (Equations (12) and (13)) [33]. According to FT-IR and XPS results, the surface of RBC800 contained C-OH, COOH, C=O and other oxygen-containing functional groups, which could be used as the active center of RBC800. Two characteristic peaks attributed to C 1s and O 1s were observed on the full XPS spectra with binding energies at 285.1 eV and 531.1 eV, respectively (Figure 7a). The O1s XPS spectra can be decomposed into three components (Figure 7b). The peak at 531.8 eV and 533.3 eV could be assigned to C=O and COOH, respectively. A peak centered at 534.1 eV corresponds to C-OH. After the reaction, the content of COOH decreased from 43.7% to 34.2%, suggesting that COOH, as the active site in the catalytic reaction, activated PMS to produce SO4•−. In addition, C-OH on the surface of RBC800 can also activate PMS to produce SO4•−, and the conversion between SO4•− and OH can be flexible (Equations (14)–(17)) [16]. Due to electrostatic attraction and the interaction between electron donors and acceptors, the adsorption of RBC800 facilitated the uniform distribution of SMX molecules on the surface of the carbon matrix, promoting contact with ROSs [34].
π e l e c t r o n s + H S O 5 S O 4 · + O H
π e l e c t r o n s + H S O 5 · O H + S O 4 2
H S O 5 S O 5 2 + H +
H S O 5 + S O 5 2 S O 4 2 + H S O 4 + 1 O 2
B C O O H + H S O 5 S O 4 · + B C O O + H 2 O
B C O H + H S O 5 S O 4 · + B C O + H 2 O
S O 4 · + H 2 O H + + S O 4 2 + · O H
· O H / S O 4 · / 1 O 2 + S M X i n t e r m e d i a t e s d e g r a d e d   p r o d u c t s + C O 2 + H 2 O
Based on the above analysis results, sp2 hybrid carbon on the RBC800 surface and oxygen-containing functional groups such as COOH, C-OH and C=O could be used as the active sites, but the interaction mechanism between PMS and these active sites has not been deeply revealed. Therefore, density functional theory (DFT) calculations were adopted to further explore the activation mechanism of PMS on the surface of RBC800.
The adsorption processes of PMS on sp2 hybrid carbon network (C/PMS), carbon network edge containing COOH functional group (COOH/PMS), carbon network edge containing C-OH functional group (C-OH/PMS) and carbon network edge containing C=O functional group (C=O/PMS) were studied. The top and side views of all the optimized configurations are shown in Figure 8. The shortest distance (D) between PMS and sp2 hybrid carbon network, COOH, C-OH and C=O were 2.18 Å, 1.46 Å, 1.43 Å and 2.01 Å, respectively (Table S5). Compared with other configurations, the C-OH/PMS configuration has the smallest D value, suggesting that there may be a strong interaction between C-OH and PMS.
The essence of PMS activation is the breaking of O-O bond. The O-O bond length (lo-o) of PMS at sp2 hybrid carbon network, COOH, C-OH and C=O is 1.454 Å, 1.464 Å, 1.468 Å and 1.458 Å, respectively. These results indicate that oxygen-containing functional groups can act as reactive sites and promote the stretching of O-O bond of PMS. The lo-o of PMS is the longest at C-OH, suggesting that the C-OH functional group could largely promote the stretching of the O-O bond of PMS, thus breaking into the activated state.
The energy barrier required for the reaction is calculated by optimizing the reactants and transition states. The energy barriers corresponding to the C/PMS, COOH/PMS, C-OH/PMS and C=O/PMS configurations are 35.16 kcal/mol, 31.93 kcal/mol, 28.12 kcal/mol and 32.28 kcal/mol, respectively. Obviously, the transition state energy barrier corresponding to the C-OH/PMS configuration is low, indicating that the C-OH functional group can significantly reduce the activation energy required for the reaction, and the catalytic activity of this active site is high. Based on the above discussion, Figure 9 displays the degradation mechanism of SMX in the RBC800/PMS system.

2.4. Degradation Pathways of SMX

The intermediates of SMX degradation were investigated by UPLC-TOF/MS (Figures S7 and S8 and Table S6). In pathway I, under the attack of OH, electrophilic addition reaction occurred on the C atom of the isoxazole ring of SMX to produce product I (m/z = 288) [2]. In addition, SO4•− may also attack the olefin double bond on the isoxazole ring of the SMX molecule and form product I [35]. Subsequently, OH further attacks the S-N bond, transforming product I into product II (m/z = 133) and product III (m/z = 190). Surface charge distribution of SMX manifests that negatively charge regions are concentrated around the O and N atoms (Figure S9). Via the attack of ROSs, the N-O bond on the isoxazole ring breaks to produce product IV (m/z = 117), which is eventually mineralized into CO2 and H2O. In pathway II, under the attack of OH, SMX molecules first undergo nitration reaction to produce product VI (m/z = 284), then break S-C bond on SMX molecules and undergo hydroxylation reaction to produce product VII (m/z = 143), and then undergo benzene ring opening reaction and transform into small molecular organic matter, and eventually mineralized into CO2 and H2O.

2.5. Toxicity Assessment

The bioaccumulation factor and developmental toxicity of SMX as well as its intermediate products were estimated by Toxicity Estimation Software Tool (T.E.S.T) (5.1.1.0). T.E.S.T has a simple interface and easy operation, and the software is equipped with a guide, which is easy to use. Users can draw and load the structure of chemical substances by CAS number, SMILES code, substance name, InChi code, DTXSID or manually, then select the prediction end point and method, and change the output path of the result. The software automatically generates a result report after the prediction. The bioaccumulation factor of SMX was 17.11 mg/L. The bioaccumulation factor of P(I), P(II), P(III), P(IV), P(V), P(VI), P(VII) and P(VIII) were 0.67, 0.59, 1.89, 0.44, 0.49, 17.06, 7.04 and 1.34 mg/L, respectively (Figure 10a). The developmental toxicity of SMX was 0.85 mg/L. The developmental toxicity of P(I), P(II), P(III), P(IV), P(V), P(VI), P(VII) and P(VIII) were 0.79, 0.54, 0.60, 0.72, 0.66, 0.78, 0.57 and 0.50 mg/L, respectively (Figure 10b). The bioaccumulation factor and developmental toxicity of all intermediates was lower than that of SMX. According to the results of toxicity analysis, the comprehensive environmental risk of SMX was decreased in the RBC800/PMS system.

3. Materials and Methods

3.1. Preparation of Catalysts

Firstly, Rice husk was washed, oven-dried at 100 °C for 12 h, ground, and then passed through a 100-mesh sieve to acquire thin powders for further use. BC was prepared by pyrolysis at a constant calcination temperature under N2 atmosphere (5 °C/min of heating rate) for 2 h. The resulting powers are further ground, sieved and then stored in a ziplock bag. To further remove ash, soluble salts and other impurities from biochar, and increase the content of acidic functional groups on the BC surface, 100.0 g of the collected sample was added to 300.0 mL of 0.1 M hydrochloric acid (HCl, 36.0–38.0%) solution. The mixture was stirred for 24 h, then filtered, and the biochar is washed with ultra-pure water for several times to stabilize its pH to 6.0. The biochar was dried at 100–105 °C for 12 h after cleaning, and the resulting composities were denoted as RBCT (T = 600, 700, 800, 900 °C).

3.2. Reaction Procedures

Degradation tests were conducted in 50 mL vials to investigate the catalytic activity of RBCT. The range of catalyst dosing was 0.2–1.0 g/L, the pH was 3.0–11.0, the PMS concentration was 0.2–1.0 mM, and the reaction temperature was set to 25 °C. The batch experiments were carried out at 150 rpm, and 1.0 mL reaction solution was taken periodically and filtered through a 0.22 μm filter. The redundant reaction was restrained by adding 1 mL Na2S2O3 (0.2 M). All tests were conducted in triplicate, and the final values were averaged.

4. Conclusions

Raw biochar (RBCT) was prepared by single one-step pyrolysis strategies using green and cheap rice husk biomass as precursor. The structure-activity relationship between physicochemical properties of RBCT and catalytic activity is discussed in detail. The pyrolysis temperature (T) has a crucial influence on the morphology and structural characteristics of RBCT, which determines the catalytic performance of RBCT. When T < 800 °C, the porous structure of RBCT is not fully developed, and when T > 800 °C, the carbon skeleton of RBCT collapses. When T = 800 °C, RBC800 exhibits a typical hardwood structure with micropores and mesoporous. In addition, RBC800 obtains more defect sites than RBC600, RBC700, and RBC900. Therefore, under optimal conditions, the elimination rate of SMX in the RBC800/PMS system within 200 min was 92.0%, kobs is 0.009 min−1, which is 1.8, 1.6 and 1.5 times that of the RBC600/PMS, RBC700/PMS and RBC900/PMS systems. Compared with similar reports, rice husk-derived biochar has obtained higher degradation performance [36]. Radical quenching, ESR and XPS analysis manifested that the main ROSs in the RBC800/PMS system were measured to be SO4•−. According to the DFT calculation results, compared with other active sites, the C-OH functional group on the surface of RBC800 interacts more strongly with PMS, which is more likely to promote the stretching of the O-O bond of PMS, thus breaking into the activated state and significantly reducing the activation energy required for reaction. This work reveals in depth the interaction mechanism between active sites of biochar-based catalysts and PMS at the molecular level. Besides, the structure-activity relationship between the physicochemical properties of catalysts and the catalytic activity were revealed.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms252111582/s1. References [12,15,26,37,38,39,40,41,42] are cited in the supplementary materials.

Author Contributions

T.L.: Investigation, Methodology, Formal analysis, Data curation, Writing—original draft. C.-X.L.: Investigation, Writing—review & editing. X.C.: Writing—review & editing. Y.C.: Writing—review & editing. K.C.: Supervision, Conceptualization, Funding acquisition. Q.W.: Conceptualization, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by National Key R&D Program of China (2019YFC0408500), Major Science and Technology Projects of Anhui Province (201903a07020009, 202003a07020004), Hefei independent innovation policy loan transfer subsidy project (J2020K07), the National Natural Science Foundation of China (52300016), China Postdoctoral Science Foundation (No. 2023M733379).

Data Availability Statement

The authors of this paper also conducted similar experiments and extracted different data sets. The raw data supporting the conclusions of this article will be madeavailable by the authors upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Adeleye, A.S.; Xue, J.; Zhao, Y.; Taylor, A.A.; Zenobio, J.E.; Sun, Y.; Han, Z.; Salawu, O.A.; Zhu, Y. Abundance, fate, and effects of pharmaceuticals and personal care products in aquatic environments. J. Hazard. Mater. 2022, 424, 127284. [Google Scholar] [CrossRef] [PubMed]
  2. Du, J.; Guo, W.; Wang, H.; Yin, R.; Zheng, H.; Feng, X.; Che, D.; Ren, N. Hydroxyl radical dominated degradation of aquatic sulfamethoxazole by Fe(0)/bisulfite/O2: Kinetics, mechanisms, and pathways. Water Res. 2018, 138, 323–332. [Google Scholar] [CrossRef] [PubMed]
  3. Yu, X.; Sui, Q.; Lyu, S.; Zhao, W.; Liu, J.; Cai, Z.; Yu, G.; Barcelo, D. Municipal Solid Waste Landfills: An Underestimated Source of Pharmaceutical and Personal Care Products in the Water Environment. Environ. Sci. Technol. 2020, 54, 9757–9768. [Google Scholar] [CrossRef] [PubMed]
  4. Xu, M.; Li, J.; Yan, Y.; Zhao, X.; Yan, J.; Zhang, Y.; Lai, B.; Chen, X.; Song, L. Catalytic degradation of sulfamethoxazole through peroxymonosulfate activated with expanded graphite loaded CoFe2O4 particles. Chem. Eng. J. 2019, 369, 403–413. [Google Scholar] [CrossRef]
  5. 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]
  6. Xin, S.; Ma, B.; Zhang, C.; Ma, X.; Xu, P.; Zhang, G.; Gao, M.; Xin, Y. Catalytic activation of peroxydisulfate by alfalfa-derived nitrogen self-doped porous carbon supported CuFeO2 for nimesulide degradation: Performance, mechanism and DFT calculation. Appl. Catal. B Environ. 2021, 294, 120247. [Google Scholar] [CrossRef]
  7. Zhen, J.; Zhang, S.; Zhuang, X.; Ahmad, S.; Lee, T.; Si, H.; Cao, C.; Ni, S.-Q. Sulfate radicals based heterogeneous peroxymonosulfate system catalyzed by CuO-Fe3O4-Biochar nanocomposite for bisphenol A degradation. J. Water Process Eng. 2021, 41, 102078. [Google Scholar] [CrossRef]
  8. Oh, W.D.; Zaeni, J.R.J.; Lisak, G.; Lin, K.A.; Leong, K.H.; Choong, Z.Y. Accelerated organics degradation by peroxymonosulfate activated with biochar co-doped with nitrogen and sulfur. Chemosphere 2021, 277, 130313. [Google Scholar] [CrossRef]
  9. Wang, C.; Huang, R.; Sun, R.; Yang, J.; Sillanpää, M. A review on persulfates activation by functional biochar for organic contaminants removal: Synthesis, characterizations, radical determination, and mechanism. J. Environ. Chem. Eng. 2021, 9, 106267. [Google Scholar] [CrossRef]
  10. Liu, T.; Cui, K.; Li, C.X.; Chen, Y.; Wang, Q.; Yuan, X.; Chen, Y.; Liu, J.; Zhang, Q. Efficient peroxymonosulfate activation by biochar-based nanohybrids for the degradation of pharmaceutical and personal care products in aquatic environments. Chemosphere 2023, 311, 137084. [Google Scholar] [CrossRef]
  11. Zhao, C.; Shao, B.; Yan, M.; Liu, Z.; Liang, Q.; He, Q.; Wu, T.; Liu, Y.; Pan, Y.; Huang, J.; et al. Activation of peroxymonosulfate by biochar-based catalysts and applications in the degradation of organic contaminants: A review. Chem. Eng. J. 2021, 416, 128829. [Google Scholar] [CrossRef]
  12. Qi, Y.; Ge, B.; Zhang, Y.; Jiang, B.; Wang, C.; Akram, M.; Xu, X. Three-dimensional porous graphene-like biochar derived from Enteromorpha as a persulfate activator for sulfamethoxazole degradation: Role of graphitic N and radicals transformation. J. Hazard. Mater. 2020, 399, 123039. [Google Scholar] [CrossRef]
  13. Liu, B.; Guo, W.; Wang, H.; Si, Q.; Zhao, Q.; Luo, H.; Ren, N. B-doped graphitic porous biochar with enhanced surface affinity and electron transfer for efficient peroxydisulfate activation. Chem. Eng. J. 2020, 396, 125119. [Google Scholar] [CrossRef]
  14. Li, L.; Lai, C.; Huang, F.; Cheng, M.; Zeng, G.; Huang, D.; Li, B.; Liu, S.; Zhang, M.; Qin, L.; et al. Degradation of naphthalene with magnetic bio-char activate hydrogen peroxide: Synergism of bio-char and Fe-Mn binary oxides. Water Res. 2019, 160, 238–248. [Google Scholar] [CrossRef]
  15. Wang, S.; Wang, J. Activation of peroxymonosulfate by sludge-derived biochar for the degradation of triclosan in water and wastewater. Chem. Eng. J. 2019, 356, 350–358. [Google Scholar] [CrossRef]
  16. Sun, H.; Peng, X.; Zhang, S.; Liu, S.; Xiong, Y.; Tian, S.; Fang, J. Activation of peroxymonosulfate by nitrogen-functionalized sludge carbon for efficient degradation of organic pollutants in water. Bioresour. Technol. 2017, 241, 244–251. [Google Scholar] [CrossRef] [PubMed]
  17. Kim, D.-G.; Ko, S.-O. Effects of thermal modification of a biochar on persulfate activation and mechanisms of catalytic degradation of a pharmaceutical. Chem. Eng. J. 2020, 399, 125377. [Google Scholar] [CrossRef]
  18. He, J.; Xiao, Y.; Tang, J.; Chen, H.; Sun, H. Persulfate activation with sawdust biochar in aqueous solution by enhanced electron donor-transfer effect. Sci. Total Environ. 2019, 690, 768–777. [Google Scholar] [CrossRef]
  19. Shang, Y.; Chen, C.; Zhang, P.; Yue, Q.; Li, Y.; Gao, B.; Xu, X. Removal of sulfamethoxazole from water via activation of persulfate by Fe3C@NCNTs including mechanism of radical and nonradical process. Chem. Eng. J. 2019, 375, 122004. [Google Scholar] [CrossRef]
  20. Yu, J.; Feng, H.; Tang, L.; Pang, Y.; Zeng, G.; Lu, Y.; Dong, H.; Wang, J.; Liu, Y.; Feng, C.; et al. Metal-free carbon materials for persulfate-based advanced oxidation process: Microstructure, property and tailoring. Prog. Mater. Sci. 2020, 111, 100654. [Google Scholar] [CrossRef]
  21. Liu, T.; Wang, Q.; Li, C.; Cui, M.; Chen, Y.; Liu, R.; Cui, K.; Wu, K.; Nie, X.; Wang, S. Synthesizing and characterizing Fe3O4 embedded in N-doped carbon nanotubes-bridged biochar as a persulfate activator for sulfamethoxazole degradation. J. Clean. Prod. 2022, 353, 131669. [Google Scholar] [CrossRef]
  22. Liu, T.; Li, C.; Chen, X.; Chen, Y.; Cui, K.; Wang, D.; Wei, Q. Peroxymonosulfate Activation by Fe@N Co-Doped Biochar for the Degradation of Sulfamethoxazole: The Key Role of Pyrrolic N. Int. J. Mol. Sci. 2024, 25, 10528. [Google Scholar]
  23. You, S.; Ok, Y.S.; Chen, S.S.; Tsang, D.C.W.; Kwon, E.E.; Lee, J.; Wang, C.H. A critical review on sustainable biochar system through gasification: Energy and environmental applications. Bioresour. Technol. 2017, 246, 242–253. [Google Scholar] [CrossRef] [PubMed]
  24. Xiao, K.; Liang, F.; Liang, J.; Xu, W.; Liu, Z.; Chen, B.; Jiang, X.; Wu, X.; Xu, J.; Beiyuan, J.; et al. Ce-embedded N-enriched porous biochar for peroxymonosulfate activation in metronidazole degradation: Applications, mechanism insight and toxicity evaluation. Chem. Eng. J. 2022, 433, 134387. [Google Scholar] [CrossRef]
  25. Hu, Y.; Chen, D.; Zhang, R.; Ding, Y.; Ren, Z.; Fu, M.; Cao, X.; Zeng, G. Singlet oxygen-dominated activation of peroxymonosulfate by passion fruit shell derived biochar for catalytic degradation of tetracycline through a non-radical oxidation pathway. J. Hazard. Mater. 2021, 419, 126495. [Google Scholar] [CrossRef]
  26. Wu, Z.; Wang, Y.; Xiong, Z.; Ao, Z.; Pu, S.; Yao, G.; Lai, B. Core-shell magnetic Fe3O4@Zn/Co-ZIFs to activate peroxymonosulfate for highly efficient degradation of carbamazepine. Appl. Catal. B Environ. 2020, 277, 119136. [Google Scholar] [CrossRef]
  27. Fu, H.; Zhao, P.; Xu, S.; Cheng, G.; Li, Z.; Li, Y.; Li, K.; Ma, S. Fabrication of Fe3O4 and graphitized porous biochar composites for activating peroxymonosulfate to degrade p-hydroxybenzoic acid: Insights on the mechanism. Chem. Eng. J. 2019, 375, 121980. [Google Scholar] [CrossRef]
  28. Chen, Y.; Bai, X.; Ji, Y.; Shen, T. Reduced graphene oxide-supported hollow Co3O4@N-doped porous carbon as peroxymonosulfate activator for sulfamethoxazole degradation. Chem. Eng. J. 2022, 430, 132951. [Google Scholar] [CrossRef]
  29. Peng, L.; Duan, X.; Shang, Y.; Gao, B.; Xu, X. Engineered carbon supported single iron atom sites and iron clusters from Fe-rich Enteromorpha for Fenton-like reactions via nonradical pathways. Appl. Catal. B Environ. 2021, 287, 119963. [Google Scholar] [CrossRef]
  30. Yao, B.; Luo, Z.; Du, S.; Yang, J.; Zhi, D.; Zhou, Y. Magnetic MgFe2O4/biochar derived from pomelo peel as a persulfate activator for levofloxacin degradation: Effects and mechanistic consideration. Bioresour. Technol. 2022, 346, 126547. [Google Scholar] [CrossRef]
  31. Fu, H.; Ma, S.; Zhao, P.; Xu, S.; Zhan, S. Activation of peroxymonosulfate by graphitized hierarchical porous biochar and MnFe2O4 magnetic nanoarchitecture for organic pollutants degradation: Structure dependence and mechanism. Chem. Eng. J. 2019, 360, 157–170. [Google Scholar] [CrossRef]
  32. Li, X.; Yang, Z.; Wu, G.; Huang, Y.; Zheng, Z.; Garces, H.F.; Yan, K. Fabrication of ultrathin lily-like NiCo2O4 nanosheets via mooring NiCo bimetallic oxide on waste biomass-derived carbon for highly efficient removal of phenolic pollutants. Chem. Eng. J. 2022, 441, 136066. [Google Scholar] [CrossRef]
  33. Liu, T.; Cui, K.; Chen, Y.; Li, C.; Cui, M.; Yao, H.; Chen, Y.; Wang, S. Removal of chlorophenols in the aquatic environment by activation of peroxymonosulfate with nMnOx@Biochar hybrid composites: Performance and mechanism. Chemosphere 2021, 283, 131188. [Google Scholar] [CrossRef] [PubMed]
  34. Peng, B.; Chen, L.; Que, C.; Yang, K.; Deng, F.; Deng, X.; Shi, G.; Xu, G.; Wu, M. Adsorption of Antibiotics on Graphene and Biochar in Aqueous Solutions Induced by pi-pi Interactions. Sci. Rep. 2016, 6, 31920. [Google Scholar]
  35. Yang, Y.; Lu, X.; Jiang, J.; Ma, J.; Liu, G.; Cao, Y.; Liu, W.; Li, J.; Pang, S.; Kong, X.; et al. Degradation of sulfamethoxazole by UV, UV/H2O2 and UV/persulfate (PDS): Formation of oxidation products and effect of bicarbonate. Water Res. 2017, 118, 196–207. [Google Scholar] [CrossRef] [PubMed]
  36. Yin, R.; Guo, W.; Wang, H.; Du, J.; Wu, Q.; Chang, J.-S.; Ren, N. Singlet oxygen-dominated peroxydisulfate activation by sludge-derived biochar for sulfamethoxazole degradation through a nonradical oxidation pathway: Performance and mechanism. Chem. Eng. J. 2019, 357, 589–599. [Google Scholar] [CrossRef]
  37. Hung, C.M.; Chen, C.W.; Huang, C.P.; Shiung Lam, S.; Dong, C.D. Peroxymonosulfate activation by a metal-free biochar for sulfonamide antibiotic removal in water and associated bacterial community composition. Bioresour. Technol. 2022, 343, 126082. [Google Scholar] [CrossRef]
  38. Liang, J.; Duan, X.; Xu, X.; Chen, K.; Wu, F.; Qiu, H.; Liu, C.; Wang, S.; Cao, X. Biomass-derived pyrolytic carbons accelerated Fe(III)/Fe(II) redox cycle for persulfate activation: Pyrolysis temperature-depended performance and mechanisms. Appl. Catal. B Environ. 2021, 297, 120446. [Google Scholar] [CrossRef]
  39. Wang, W.; Chen, M. Catalytic degradation of sulfamethoxazole by peroxymonosulfate activation system composed of nitrogen-doped biochar from pomelo peel: Important roles of defects and nitrogen, and detoxification of intermediates. J. Colloid Interface Sci. 2022, 613, 57–70. [Google Scholar] [CrossRef]
  40. Yu, J.; Tang, L.; Pang, Y.; Zeng, G.; Feng, H.; Zou, J.; Wang, J.; Feng, C.; Zhu, X.; Ouyang, X.; et al. Hierarchical porous biochar from shrimp shell for persulfate activation: A two-electron transfer path and key impact factors. Appl. Catal. B Environ. 2020, 260, 118160. [Google Scholar] [CrossRef]
  41. Zhang, H.; Song, Y.; Nengzi, L.-c.; Gou, J.; Li, B.; Cheng, X. Activation of persulfate by a novel magnetic CuFe2O4/Bi2O3 composite for lomefloxacin degradation. Chem. Eng. J. 2020, 379, 122362. [Google Scholar] [CrossRef]
  42. Zhao, Y.; Song, M.; Cao, Q.; Sun, P.; Chen, Y.; Meng, F. The superoxide radicals’ production via persulfate activated with CuFe2O4@Biochar composites to promote the redox pairs cycling for efficient degradation of o-nitrochlorobenzene in soil. J. Hazard. Mater. 2020, 400, 122887. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (ad) SEM images of RBCT; (e) EDS elemental content, and (f) element mappings of RBC800.
Figure 1. (ad) SEM images of RBCT; (e) EDS elemental content, and (f) element mappings of RBC800.
Ijms 25 11582 g001
Figure 2. (a) HRTEM, and (b) SAED pattern of RBC800.
Figure 2. (a) HRTEM, and (b) SAED pattern of RBC800.
Ijms 25 11582 g002
Figure 3. (a) XRD patterns, (b) FTIR spectra, (c) Nitrogen adsorption–desorption isotherms and pore structure (the inset), and (d) Raman spectra of RBCT.
Figure 3. (a) XRD patterns, (b) FTIR spectra, (c) Nitrogen adsorption–desorption isotherms and pore structure (the inset), and (d) Raman spectra of RBCT.
Ijms 25 11582 g003
Figure 4. The adsorption rate (a) and degradation rate (b) of SMX in various reaction systems. Reaction conditions: [SMX]0 = 10.0 mg/L; [RBCT] = 0.4 g/L; [PMS]0 = 0.6 mM; pH = 7.0; Reaction temperature = 25 °C.
Figure 4. The adsorption rate (a) and degradation rate (b) of SMX in various reaction systems. Reaction conditions: [SMX]0 = 10.0 mg/L; [RBCT] = 0.4 g/L; [PMS]0 = 0.6 mM; pH = 7.0; Reaction temperature = 25 °C.
Ijms 25 11582 g004
Figure 5. Effects of ROSs scavengers on the SMX degradation in the RBC800/PMS system. Reaction conditions: [SMX]0 = 10.0 mg/L; [RBC800] = 0.4 g/L; [PMS]0 = 0.6 mM; pH = 7.0; MeOH = 0.5 M; TBA = 0.5 M; FFA = 0.5 M; p-BQ = 20.0 mM; Reaction temperature = 25 °C.
Figure 5. Effects of ROSs scavengers on the SMX degradation in the RBC800/PMS system. Reaction conditions: [SMX]0 = 10.0 mg/L; [RBC800] = 0.4 g/L; [PMS]0 = 0.6 mM; pH = 7.0; MeOH = 0.5 M; TBA = 0.5 M; FFA = 0.5 M; p-BQ = 20.0 mM; Reaction temperature = 25 °C.
Ijms 25 11582 g005
Figure 6. ESR signals of (a) TEMP-1O2 and (b) DMPO-OH and DMPO-SO4•−. (Conditions: [SMX]0 = 10.0 mg/L; [RBC800]0 = 0.4 g/L; [PMS]0 = 0.6 mM; pH = 7.0; Reaction temperature = 25 °C; [TEMP] = [DMPO] = 10.0 mM).
Figure 6. ESR signals of (a) TEMP-1O2 and (b) DMPO-OH and DMPO-SO4•−. (Conditions: [SMX]0 = 10.0 mg/L; [RBC800]0 = 0.4 g/L; [PMS]0 = 0.6 mM; pH = 7.0; Reaction temperature = 25 °C; [TEMP] = [DMPO] = 10.0 mM).
Ijms 25 11582 g006
Figure 7. XPS spectra of full-range survey (a), and O 1s (b) of RBC800.
Figure 7. XPS spectra of full-range survey (a), and O 1s (b) of RBC800.
Ijms 25 11582 g007
Figure 8. The optimization structures of PMS adsorption on different sites and the corresponding transition state. (a) C/PMS, (b) COOH/PMS, (c) C-OH/PMS and (d) C=O/PMS.
Figure 8. The optimization structures of PMS adsorption on different sites and the corresponding transition state. (a) C/PMS, (b) COOH/PMS, (c) C-OH/PMS and (d) C=O/PMS.
Ijms 25 11582 g008
Figure 9. Proposed mechanism of SMX degradation in the RBC800/PMS system.
Figure 9. Proposed mechanism of SMX degradation in the RBC800/PMS system.
Ijms 25 11582 g009
Figure 10. Bioaccumulation factor (a), and developmental toxicity (b) of SMX and its degradation byproducts.
Figure 10. Bioaccumulation factor (a), and developmental toxicity (b) of SMX and its degradation byproducts.
Ijms 25 11582 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

Liu, T.; Li, C.-X.; Chen, X.; Chen, Y.; Cui, K.; Wei, Q. Peroxymonosulfate Activation by Rice-Husk-Derived Biochar (RBC) for the Degradation of Sulfamethoxazole: The Key Role of Hydroxyl Groups. Int. J. Mol. Sci. 2024, 25, 11582. https://doi.org/10.3390/ijms252111582

AMA Style

Liu T, Li C-X, Chen X, Chen Y, Cui K, Wei Q. Peroxymonosulfate Activation by Rice-Husk-Derived Biochar (RBC) for the Degradation of Sulfamethoxazole: The Key Role of Hydroxyl Groups. International Journal of Molecular Sciences. 2024; 25(21):11582. https://doi.org/10.3390/ijms252111582

Chicago/Turabian Style

Liu, Tong, Chen-Xuan Li, Xing Chen, Yihan Chen, Kangping Cui, and Qiang Wei. 2024. "Peroxymonosulfate Activation by Rice-Husk-Derived Biochar (RBC) for the Degradation of Sulfamethoxazole: The Key Role of Hydroxyl Groups" International Journal of Molecular Sciences 25, no. 21: 11582. https://doi.org/10.3390/ijms252111582

APA Style

Liu, T., Li, C.-X., Chen, X., Chen, Y., Cui, K., & Wei, Q. (2024). Peroxymonosulfate Activation by Rice-Husk-Derived Biochar (RBC) for the Degradation of Sulfamethoxazole: The Key Role of Hydroxyl Groups. International Journal of Molecular Sciences, 25(21), 11582. https://doi.org/10.3390/ijms252111582

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