Next Article in Journal
Large-Scale Hydrological Models and Transboundary River Basins
Previous Article in Journal
Energy Poverty Impact on Sustainable Development of Water Resources in China: The Study of an Entropy Recycling Dynamic Two-Stage SBM Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Degradation of Water Pollutants by Biochar Combined with Advanced Oxidation: A Systematic Review

by
Fanrong Kong
1,2,
Jin Liu
2,
Zaixin Xiang
2,
Wei Fan
1,
Jiancong Liu
1,
Jinsheng Wang
2,3,
Yangyang Wang
2,3,*,
Lei Wang
2,3,* and
Beidou Xi
2,3
1
School of Environment, Northeast Normal University, Changchun 130117, China
2
School of Materials and Environmental Engineering, Institute of Urban Ecology and Environment Technology, Shenzhen Polytechnic University, Shenzhen 518055, China
3
Eco-Environmental Science Center (Guangdong, Hong-Kong, Macau), Guangzhou 510555, China
*
Authors to whom correspondence should be addressed.
Water 2024, 16(6), 875; https://doi.org/10.3390/w16060875
Submission received: 17 February 2024 / Revised: 10 March 2024 / Accepted: 12 March 2024 / Published: 18 March 2024
(This article belongs to the Section Wastewater Treatment and Reuse)

Abstract

:
Recently, biochar has emerged as a promising option for environmentally friendly remediation due to its cost-effectiveness, extensive surface area, porosity, and exceptional electrical conductivity. Biochar-based advanced oxidation procedures (BC-AOPs) have gained popularity as an effective approach to breaking down organic pollutants in aqueous environments. It is commonly recognized that the main reactive locations within BC-AOPs consist of functional groups found on biochar, which encompass oxygen-containing groups (OCGs), imperfections, and persistent free radicals (PFRs). Additionally, the existence of metallic components supported on biochar and foreign atoms doped into it profoundly impacts the catalytic mechanism. These components not only modify the fundamental qualities of biochar but also serve as reactive sites. Consequently, this paper offers a comprehensive review of the raw materials, preparation techniques, modification approaches, and composite catalyst preparation within the biochar catalytic system. Special attention is given to explaining the modifications in biochar properties and their impacts on catalytic activity. This paper highlights degradation mechanisms, specifically pathways that include radical and non-radical processes. Additionally, it thoroughly examines the importance of active sites as catalysts and the basic catalytic mechanism of BC-AOPs. Finally, the potential and future directions of environmental remediation using biochar catalysts and advanced oxidation processes (AOPs) are discussed. Moreover, suggestions for future advancements in BC-AOPs are provided to facilitate further development.

Graphical Abstract

1. Introduction

In the past few years, there has been a notable increase in global studies concentrating on the treatment of wastewater containing persistent organic pollutants. These compounds present difficulties because of their resistance to degradation [1]. Recently, AOPs have garnered considerable attention for their exceptional potential and distinct benefits, including thorough oxidation, rapid reaction rates, efficient treatment, and a minimal environmental impact [2,3]. AOPs involve the generation of potent oxidizing agents, such as superoxide radicals (•O2), hydroxyl radicals (•OH), sulfate radicals (SO4•), and singlet oxygen (1O2), through diverse mechanisms like electrical stimulation, acoustic waves, photolysis, and catalytic reactions [4].
This leads to the direct mineralization of recalcitrant substances in sewage or the oxidative degradation of large molecules into more readily biodegradable compounds, thereby enhancing the biodegradability of wastewater [5,6]. In contrast to conventional water treatment methods, AOPs exhibit distinctive features: (1) AOPs relies on the continuous improvement of free radical generation efficiency, and do not produce secondary pollution. This allows sewage treatment to achieve the goal of zero environmental pollution and zero waste discharge. (2) The generated free radicals all exhibit non-selective reactivity, providing a broad spectrum of and a non-discriminatory method for pollutant degradation, ensuring their ability to oxidize most organic pollutants [7]. (3) Free radicals have the advantages of a fast reaction speed, a short time of reaction time, mild reaction conditions, and easy control of operating conditions [8]. (4) AOPs can operate either autonomously or in combination with other processing methods. For example, the combined application of the UV-Fenton process can produce better treatment results, while the Fenton technology alone can effectively address challenging wastewater scenarios, providing cost-effective solutions and improving efficiency [9].
Specifically, AOPs typically employ reactive oxygen species (ROSs) with potent REDOX capabilities to achieve the complete breakdown of recalcitrant contaminants in water systems. ROSs mainly include •OH, SO4•, •O2, and 1O2. The main ROSs generation systems include Fenton oxidation, photocatalysis, persulfate (PS), ozone oxidation, ultrasonic oxidation, and electrocatalysis. Photocatalysis typically entails a semiconductor being exposed to light containing photon energy greater than the band gap energy of the material. As a result, photogenerated holes and photogenerated electrons are formed, leading to the generation of •OH, •O2, and 1O2 when dissolved oxygen, protons (H+), or water (H2O) are present (Equations (1)−(5) in Table 1) [10]. The Fenton system is a traditional advanced oxidation process mode, and the classic Fenton reaction is performed by the •OH produced by the reaction of Fe2+ to H2O2 (Equation (6)). Due to the lower cost of Fe3+, it is also used as a catalyst, which is called a Fenton-like reaction (Equation (7)). In addition, other complex reactions may occur simultaneously in solution, such as the use of other varivalent metals like Mn+, dual-cathode systems, etc., (Equations (8)–(15), Equations (2) and (5)) [2]. The primary mechanism for the electrocatalytic degradation of antibiotics involves both direct and indirect oxidation. Normally, reactions for direct oxidation occur at the anode surface, where electrons transfer directly from the electrode to the antibiotic. The indirect oxidation process primarily includes the generation of active species on either the anode or cathode surface. The most widely used systems are the electrocatalytic system (EC) and photocatalytic system (PEC). For example, Wang et al. [11] stimulate C-O units on the anode with PMS to generate C=O and SO4• (Equation (16)). Following this, the C=O units are diminished on the light-collecting anode to renew C-O (Equation (17)). SO4• has the capability to convert OH to •OH and 1O2 (Equations (18) and (19)) [12,13]. In aqueous solutions, ozone reacts with various components in two ways: through the direct oxidation of molecular ozone, which involves a selective reaction; or an indirect mechanism by which •OH radicals are produced through ozonolysis (Equation (20)), which are non-selective and highly reactive [14]. Ozone and •OH free radicals are powerful chemical oxidants involved in the disinfection and oxidation of pollutants. Ultrasonic (US) ( ))) ) treatment is an important advanced oxidation technology which can be used alone or in combination with other advanced oxidation technologies such as O3/US, H2 O2/US, UV/US, etc. The ultrasonic treatment of water can form free radicals [15,16] (Equations (21)–(23) and Equation (9)). Sound waves react with oxygen in the gas phase to form oxygen atoms, which in turn react with water to produce additional free radicals (Equations (24)–(27)) [17].
•OH and SO4• are both highly reactive free radicals. Compared to the hydroxyl radical •OH, the sulfate radical SO4• has a longer half-life of 30–40 μs and a higher REDOX potential ranging from 2.5 to 3.1 V. Additionally, it demonstrates pH-independent reactivity. The reaction rate of •OH is 106~1010 mol/L.s, and that of SO4• is (3.39 ± 0.22) 109 mol/L.s. The way to distinguish these two types of free radicals is mainly through their production mode and reaction characteristics. •OH is usually produced by processes such as REDOX reactions or photolysis, while SO4• requires the decomposition of persulfate (PS) to produce it. Activated PS-AOPs have gained significant interest for their ability to efficiently degrade pollutants, attributed to their stability, efficacy, and lack of secondary contamination [18]. The central idea focuses on generating active radicals, such as SO4• or •OH. These free radicals drive the oxidation of pollutants, converting them to H2O and CO2 with excellent selectivity and reactivity. Common persulfate activators include persulfate bisulfate (PDS) and peroxybisulfate (PMS), which have different structures and exhibit different catalytic activities. The mechanisms for activating PDS and PMS vary, resulting in differing production of free radicals and REDOX potentials [19].
The bond energy of PS, which measures 140 kJ·mol−1, is too high for direct interaction with pollutants. Usually, activation is required to break the -o-o- bond and produce SO4•. In the presence of light, the persulfate system uses energy from ultraviolet or visible light to break the peroxygen bond in the persulfate ion, resulting in the creation of sulfate radicals (Equations (1)–(3), (28), and (29)) [20]. The same principle applies to the thermal activation of PS (Equations (30)–(32)) [21]. Research has shown that ultrasound (US) treatment triggers the conversion of persulfates into sulfate anion radicals through the assistance of PS in collapsing and fragmenting bubbles (Equation (33)) [22]. Moreover, the role of •OH in the generation of SO4• is brought about by the ultrasound-induced formation of •OH from water (Equations (21) and (34) [23]. Within the electrocatalytic PS system, PS is catalyzed by sulfate ion formation following e-receipt at the cathode (Equation (35)) [24]. The activation of persulfate through alkali also serves as a popular technique, yielding •O2, SO4•, and •OH (Equations (36) and (37)) [25,26]. In addition, the transition metal activation of PS can produce SO4• (Equation (38)). PMS and PDA can self-decompose to generate 1O2 (Equations (39)–(42)) [27].
Table 1. Classification table of different free radical sources.
Table 1. Classification table of different free radical sources.
MethodFree RadicalMechanismReference
Photocatalysis•OH, •O2, 1O2Photocatalyst + UV/vis → h+ + e(1) [10]
h+ + H2O → •OH + H+(2)
e + O2 → •O2(3)
•O2 + H+ + e → H2O2(4)
H2O2 + e → •OH + OH(5)
Fenton process•OHFe2+ + H2O2 → Fe3+ + •OH + OH(6) [2]
Fe3+ + H2O2 → Fe2+ + OOH• + H+(7)
Fe2+ + •OH → Fe3+ + OH(8)
•OH + •OH → H2O2(9)
H2O2 + •OH → H2O + OOH•(10)
•OH + OOH• → H2O + O2(11)
Mn+ + H2O2 → M(n+1)+ + OH + •OH(12)
Fe(OH)2+ + hv → Fe2+ + •OH(13)
e + Fe3+ → Fe2+(14)
h+ + H2O → H+ + •OH(2)
O2 + 2H+ + 2e → H2O2(15)
H2O2 + e → •OH + OH(5)
Electrocatalysis•OH, 1O2, SO4C–O + PMS → C=O + SO4(16) [11,12,13]
C=O + e → C–O(17)
SO4• + H2O → •OH + H+ + SO42−(18)
SO4• + H2O → 1O2 + H+ + SO42−(19)
Ozonation•OH,3O3 + OH + H+ → •OH + 4O2(20) [14]
Ultrasonic•OH))) + H2O → •OH + •H(21) [15,16,17]
•OH + •H → H2O(22)
•OH + •OH → H2O2(9)
•H + •H → H2(23)
))) + O2 → O + O(24)
O + H2O → •OH + •OH(25)
•H + O2 → •OOH(26)
•OOH + •OOH → H2O2 + O2(27)
PS•OH, •O2, 1O2, SO4S2O82− + e + UV/vis → SO4• + SO42−(28) [20,21,22,23,24,25,26,27]
S2O82− + •O2 + UV/vis → SO4• + SO42− + O2(29)
HSO5 + heat → SO4• + •OH(30)
S2O82− + heat → 2SO4(31)
SO4• + H2O → SO42−+ •OH + H+(32)
S2O82− + ))) → 2SO4(33)
S2O82− + •OH → HSO4 + SO4• + 1/2O2(34)
S2O82− + e → SO42− + SO4(35)
S2O82− + H2O → SO42− + HO2 + 2H+(36)
S2O82− + HO2 → SO42− + SO4• + H+ + •O2(37)
S2O82− + Mn+ → M(n+1)+ + SO4• + SO42−(38)
HSO5 + SO52− → HSO4+ SO42− + 1O2(39)
S2O82− → 2SO4(40)
SO4• + H2O → •HSO42−+ •OH(41)
2•OH → H2O + 1/21O2(42)
Biochar-derived catalysts, resembling a representative catalytic material, have been suggested as catalysts for cutting-edge oxidation technologies owing to their remarkable performance, expansive surface area, and various methods of preparation [28,29,30]. Furthermore, as a cutting-edge method in oxidation technologies, the activation of PS entails the implementation of unique catalysts derived from biochar to activate PS. The main mechanism of pollutant degradation is free radical attack. This mechanism, when compared to non-free radical pathways, exhibits lower effectiveness [31]. Therefore, modification of the original biochar-based catalyst is necessary to enhance its performance. Two crucial factors for improving catalyst performance are the introduction of a porous structure and modification with metal doping and abundant heteroatoms [32]. Enhancing the surface area of biochar offers the benefit of boosting the degradation efficiency of ROPs by increasing the absorption of pollutant molecules and exposing surface active sites more effectively. Additionally, modifying the initial carbon network through the introduction of metal doping or heteroatoms can activate the catalyst and redistribute the charge density of nearby carbon atoms, leading to the creation of numerous active sites [33]. As a result, the analysis of catalysts in this research primarily concentrates on elucidating the adjustments made to biochar properties and their impact on catalytic performance.

2. Preparation of Biochar

2.1. Feedstock

The specific conditions during the production of biochar can impact its properties, which are also influenced by the composition of the biomass source [34,35]. The pivotal role is played by the heterogeneity of the feedstock, which includes various organic components like proteins, lignocellulose, lipids, etc., as well as inorganic substances like mineral composition, moisture, and volatile fraction. Carbon fixation and the construction of aromatic structures are greatly influenced by lignocellulose among these constituents. Based on research conducted by Das et al. [36], the investigation revealed that an increase in the level of lignin within the feedstock led to an escalation in the production of biochar. Commonly used feedstocks include terrestrial plant sources, aquatic plant sources, animal manure, and sludge biogenic sources. In most cases, a higher specific surface area (SSA) and lower ash content are superior biochar, and biochar extracted from plants is more eligible than biochar obtained from manure [37]. The variances in the feedstock are evident in the biochar’s morphology and structure. Specifically, aquatic plant sources, such as algae or water hyacinth, typically have higher inorganic and nitrogen content due to minerals and proteins in the feedstock [38].
Compared to biochar derived from plants on land, aquatic biomass biochar generally exhibits cation exchange capacity (CEC) and lower SSA, as well as higher ash content [39,40]. The effectiveness of different sources of biochar depends on the ash content. For example, biochar obtained from animal manure demonstrates reduced carbon content, CEC, SSA, and porosity, as well as dissolved organic carbon (DOC), while displaying elevated ash content [41]. However, it is relatively rich in nitrogen compared to other categories [42]. The nature of sludge biochar is variable and is associated with changes in sludge source over time and space. When utilizing biochar made from sludge, it is crucial to acknowledge the potential hazards linked to the presence of high levels of metallic elements. Recent studies have provided evidence of the importance of these metallic elements in the creation of PFRs [43]. Wang et al. effectively degraded 2,4-dichlorophenol (2,4-DCP) using a biomimetic catalyst composed of iron-coated biochar (referred to as biomimetic Fe⨀BC). Their approach involved utilizing Siberian iris, an aquatic plant source with a considerable ash content and specific surface area [44]. The research is characterized by the possibility of reusing waste materials, in particular “biochar” obtained from iron-rich aquatic plants through direct pyrolysis.

2.2. Biochar Preparation Method

2.2.1. Pyrolysis

Pyrolysis is the process of thermal and chemical decomposition of biomass or living organisms at temperatures above 400 °C in environments with low oxygen levels or inert gases. The process leads to the creation of three primary products: biochar, a solid byproduct; bio-oil, also known as pyrolytic oil, a liquid substance; and finally, syngas, a blend of non-condensable gases. Syngas mostly contains CO, CO2, H2, and CH4. The non-condensable gases generated during the pyrolysis process are commonly known as synthesis gases. There are two main stages. In the initial phase, heat is conveyed to the surface of the particle through radiation or convection, and subsequently to the particle. During pre- and first pyrolysis, a continuous temperature increase results in water removal from biomass pellets. During the duration of the process, the generated volatiles and syngas permeate the particle’s pores, facilitating heat transfer over time [38]. The solid pore expansion is initiated during the second phase, commencing with the conversion of biomass into gas [45]. During the pyrolysis of hot biomass [46], the interaction between volatile gases and syngas experiences significant enhancement due to the extensive enlargement of pores.
Different categories of pyrolysis processes can be classified based on their heating rates. These categories are divided by pyrolysis rate, including slow, fast, and flash. Among these, the most commonly utilized method is slow pyrolysis. In intermittent or continuous systems, the crucial role is played by the size of the particles and the moisture content, especially in continuous systems. During the pre-pyrolysis process, the temperature range typically falls between 200 °C and 246 °C. This marks the initial stage of biomass decomposition. This stage is characterized by multiple transformations, including internal reorganizations, the disruption of chemical bonds, the removal of water molecules, the appearance of unpaired electrons, and the creation of carboxyl groups, carbonyl groups, and hydroperoxides. The primary pyrolysis processes occur rapidly in the second stage, while the char undergoes slow decomposition in the third stage. As a consequence, a solid residue abundant in carbon content, known as biochar, is generated. Typically, the slow pyrolysis process produces a ratio of biochar, bio-oil, and syngas of 0.35:0.3:0.35. Higher heating rates facilitate biomass decomposition, producing more volatiles and less biochar.
Biomass pyrolysis degradation includes cellulose, hemicellulose, lignin, and other trace organic matter. The degradation process is affected by factors such as pyrolysis temperature, heating rate, pressure, biomass particle size, and reactor structure. Hemicellulose decomposition initiates between 240 and 270 °C, cellulose decomposition occurs between 260 and 320 °C, and lignin decomposition starts between 290 and 400 °C. The pyrolysis reaction consists of two basic reactions: one is cellulose pyrolysis, which decomposes and cokes at a lower heating rate and temperature; the other is cellulose pyrolysis, which rapidly volatilizes and produces levoglucan at higher heating rates and temperatures.
In addition to the above three substances, the pyrolysis mechanism of biomass is also significantly affected by a variety of inorganic compounds. In the course of biomass pyrolysis, a high mobility of K and Cl is observed, which allows them to volatilize at lower temperatures. Conversely, high temperatures are necessary for the evaporation of Ca and Mg as they are chemically bonded to organic molecules, either ionically or covalently. On the other hand, organic compounds in biomass and plant cells will decompose at lower temperatures during the pyrolysis process. During this procedure, particular inorganic elements like alkali metals and alkaline earth metals are crucial catalysts. For instance, compounds with K, Ca, and Mg serve as catalysts in the subsequent breakdown of unstable compounds produced during pyrolysis. The unstable compounds primarily consist of gaseous elements like CO, H2, CH4, C2H4, and CO2.

2.2.2. Co-Precipitation

The process of co-precipitation is a vital technique used in the creation of biochar, which includes the concurrent precipitation of multiple substances from a solution. At the same time, it is also possible to combine the co-precipitation method with the carrier, and it becomes feasible to fix metals, metal oxides, or composite materials on the biochar. And, in biochar preparation, coprecipitation is often combined with other techniques such as pyrolysis and calcination. Thanks to the extensive use of this basic yet essential technique, scientists have been able to investigate the production of magnetic biochar and active biochar from different types of biomass. Do et al. [47] synthesized magnetic biochar through co-precipitation with the Fe2+ and Fe3+ solution, and examined the adsorption capabilities of magnetic biochar on crocin in a water-based solution. Another crucial feature of coprecipitation is that it involves the precipitation of a solute from a solution with the help of a carrier. This method causes the solute to adhere to the carrier rather than dissolve in the water. By employing this technique, it becomes feasible to immobilize metals, metal oxides, or composites onto biochar. Researchers have been able to explore the generation of magnetic biochar and activated biochar from various biomasses due to extensive utilization of this uncomplicated yet essential method. In their research, Saravanan and colleagues [48] investigated the use of the biological macromolecule gulan agglutinative (GK). The preparation of magnetic biochar GK involved co-precipitation of Fe2+ and Fe3+ ions on GK with the presence of biomacromolecules in GK. This was achieved by using an ammonia solution in a 2:1 ratio. This well-crystallized iron-containing biochar demonstrated thermal stability compared to raw GK containing iron oxide particles in a spherical aggregate structure. Additionally, Asgharzadeh et al. [49] followed a similar approach, where they produced biochar (MgBC) from rice bran in alkaline conditions through the co-precipitation method. Following this, a composite of TiO2/MgBC was created utilizing the sol-gel approach. These composites displayed a 97% degradation rate of tetracycline, showcasing impressive durability. A recent study used co-precipitation to prepare MnFe2O4/biochar composites, which were then used to activate H2O2 to degrade tetracycline (TC). Throughout the entire experimental process, it was discovered that the catalyst could be reused and there was minimal leaching of metal ions [50]. Pi et al. [51] utilized a comparable approach to fabricate biochar with magnetite loading for triggering PS to break down TC. Furthermore, the produced catalyst exhibited consistency over three rounds, with leaching reported to be below 3 mg·L−1.

2.2.3. Hydrothermal

The hydrothermal treatment process involves adding the necessary ingredients to purified water and then transferring the resulting mixture to a tightly sealed autoclave lined with PTFE. This ensures that the catalyst can be uniformly prepared under specific reaction conditions. The temperature of the autoclave is then heated and maintained to the set temperature. It is essential to conduct experiments at various time and temperature parameters to determine the optimal conditions for transforming the original substance into the final product, as these conditions could vary based on the characteristics of the original material. Zhang and colleagues [52] extensively investigated the development of magnetic biochar using iron sludge and bio-sludge for H2O2 activation. To maximize the efficiency of degradation, experimental trials were implemented to ascertain the ideal proportions, temperature, and duration of the reaction. The hydrothermal method offers the advantage of a one-step process for preparing biochar-loaded metal oxide catalysts, enabling the formation of metal oxides with diverse structures [53]. Ma et al. [54] utilized corn cobs, an abundant agricultural waste, to produce biochar within 6 h under the existence of FeCl3·6H2O. The procedure ensures that the iron oxide particles are securely attached to the corn cob surface. By hydrothermal method, Fe3+ was introduced to form a mixed structure of large and medium pores in the overall morphology. In a similar vein, Gao et al. [55] conducted a study where they synthesized biochar from pine cones utilizing FeCl3·6H2O. The prepared Fe3O4/P biochars exhibited good crystallization, a porous structure, and close adherence to iron tetraoxide particles.
Taken together, these technologies can make use of different categories of agricultural residues, each with its own advantages and disadvantages. It is important to consider these factors when choosing a biochar preparation method. Pyrolysis is convenient. When produced at high temperatures, it exhibits special porosity and extensive surface area and can be diversified via preparation of different metal forms of catalysts, but an atmosphere of inertia is required. Co-precipitation is the most common method for preparing supported catalysts and mixed-metal catalysts, but the process lacks control and precision. The temperature and pressure conditions required by the hydrothermal method are mild, but the biochar catalyst composites have a relatively low specific surface area. The biomass, characteristics, and degradation effects of biochar-based catalysts prepared by the above three methods are shown in Table 2.

2.3. Modification of Biochar

An important step in preparing biochar is modification. Reasonable modification can change its physical and chemical properties to improve its adsorption and activation properties. This process is crucial in order to optimize its effectiveness. Various methods can be utilized to categorize biochar modification, including physical, chemical, and biological modifications. Changes induced by external energy and matter manifest in the modification of basic structure, alterations in surface functional groups, and variations in loading characteristics. Lee and Shin conducted a comparative study on five different modification techniques, namely acid impregnation, alkali impregnation, oxidation impregnation, MnOx impregnation, and FeOx impregnation [71]. The findings revealed significant alterations in biochar structure, with a 7.3–61.3-fold rise in specific surface area, resulting in an improved capacity for heavy metal adsorption. Specifically, the acid modification technique eliminates impurities and introduces acidic functional groups [72]. Modifying the alkalinity method [73], conversely, elevates the surface area of biochar while introducing OCGs, predominantly -OH and -COOH. Alternatively, physical alterations like ball milling or mechanochemical adjustments aim to break down non-structural components and provoking structural modifications. Ball milling technology is a technology that can reduce the size of a material while increasing the specific surface area. Ball milling technology can also introduce new elements into the material, promoting the exposure and formation of functional groups in the material. If treated with ultraviolet light, oxygen-containing functional groups can also be introduced. Zhang et al. [74] and Peng et al. [75] found that the surface C=O and C-H intensifies after the biochar is treated with ultraviolet light, which improves the polarity and aromatization of the biochar. Steam treatment has the potential to improve the pore arrangement of biochar. It is crucial to acknowledge that every adjustment technique demonstrates its individual restrictions, encompassing energy usage and the likelihood of subsequent contamination. Despite receiving less focus, biological alteration presents a more harmonized ecological strategy. According to Yuvaraj’s findings [76], earthworm-manipulated biochar, coupled with enzymes like metallothionein, β-glucosidase, carboxylesterase, and humic acid, holds promise as a material for combatting pollution.

3. Metals Loading

The addition of metal into biochar has been acknowledged as a successful approach to boost its catalytic potential. There are many studies on the activation of PS by transition metals. Transition metals, with a lone electron in their electronic setup, play a crucial role in the activation of PS. This activation leads to the creation of SO4•, which assists in both gaining and getting rid of electrons. Adding metal to biochar can enhance the catalytic reaction efficiency, increase its utilization effectiveness, and decrease metal contamination [77,78]. PS-AOP systems have received thorough examination in the realm of single-metal, binary, and multi-metallic catalysts. The synthesis of biochar-based materials utilizes metal salts or metal oxides, wherein the metal-loaded biochar is derived in situ from biomass-containing metals [79,80]. Predominantly, metals like Co and Fe are utilized, and various transition metal ions (Mn2+, Co2+, Cu2+, and Fe2+) have exhibited catalytic abilities in activated persulfate systems. After the loading of transition metals, biochar usually generates free radicals for degradation. Mn+ metal exhibits the capacity to facilitate REDOX reactions by supplying the essential electrons for free radical generation. Consequently, PS disrupts the O-O linkage, generating SO4•, which then interacts with H2O to produce •OH. The cyclic process is completed by the reaction of Mn+1 with PMS, resulting in the regeneration of Mn+ (Equations (43)–(45) and (18)). In a thorough investigation conducted by Tan et al. [81], emphasis was placed on producing nanocomposites from biomass with the aid of biochar. Their main approach involved treating biomass before and after processing biochar (Figure 1). Comparable methods have been employed to develop biochar-based substances for activating PS.
Mn+ + HSO5 → Mn+1 + SO4• + OH
Mn+ + HSO5 → Mn+1 + •OH + SO42−
Mn+1 + HSO5 − → Mn+ + SO5• + H+

3.1. Co-Based Catalysts

The activation of PS has garnered significant attention, particularly in relation to the use of catalysts like cobalt oxide and biochar. One particularly effective method of PS activation is the utilization of cobalt-loaded biochar, which shows remarkable performance under acidic conditions. The enhanced activity observed in cobalt ions is due to the creation of multiple SO4• species. These species are generated when divalent and trivalent cobalt ions undergo transformation. However, concerns arise regarding undesired cobalt leaching in the catalytic process, particularly under acidic conditions [82,83,84]. The research analyzed cobalt levels in water subjected to post-PMS catalytic activation reactions under acidic and neutral conditions. The results show that the leaching concentration of cobalt under neutral conditions is reduced by 0.70 mg·L−1 compared with that under acidic conditions. These findings demonstrate that the pH level impacts the cobalt leaching concentration in the catalytic activation process [82]. To tackle the cobalt leaching problem, various carriers for Co3O4 have been investigated by researchers. These include inert materials in one, two, or three dimensions [85,86], metal oxides [87], waste from industrial processes [88], magnetic particles [89], molecular sieves [90], and adsorbents [91]. Adding Co3O4 to the biochar catalyst can notably boost its catalytic performance compared to using Co3O4 alone. This enhancement is likely due to the biochar’s ability to enhance the dispersal of Co3O4, decrease Co leaching, and facilitate the catalyst’s separation/recovery from the treated water. Furthermore, the catalyst has also been designed with additional functionalities, specifically photocatalysis. However, challenges persist, including the material’s monolithic chemical composition, limited acid resistance, and pronounced issues of metal ion leaching. The stability of this material in practical applications, particularly concerning the leaching of metal ions, requires further improvement. Liu reported the degradation of atrazine using cobalt-containing biochar [92]. Chen et al. [93] achieved the degradation of antibiotics through PMS activation using biochar-supported Co3O4. According to the research findings, the promotion of electron transfer between HSO5 and Co species leads to the emergence of a synergistic effect between the loaded Co3O4 and biochar. Consequently, this synergy facilitates the redox cycle of Co3+/Co2+ [94].

3.2. Cu-Based Catalysts

Compared to other metal-biochar materials, copper-biochar materials are cost-effective, making the exploration of their catalytic activation properties economically significant. When examining the comparative performance of two highly stable copper oxide catalysts, namely CuO and Cu2O, it is evident that CuO displays a catalytic activity that is approximately 1.5–2.0 times greater than that of Cu2O [95,96]. Nevertheless, an optimal performance can solely be attained with a heightened CuO load. To illustrate, when subjecting 50 mg·L−1 phenol to a combination of copper oxide and mM Oxone®, superior results are achieved [97]. In addition, by integrating CuO into ZSM5, a carrier characterized by its expansive surface area, exceptional stability, and potent affinity for pollutants, the activity of CuO can be further amplified. This approach not only curtails CuO aggregation but also stimulates the augmentation of the catalytic reaction’s specific surface area [98,99]. CuO, apart from its ability to activate PMS, also demonstrates the capability to activate PS, thereby facilitating the breakdown of specific organic contaminants without generating SO4• [100]. In a study conducted by Li [101], an investigation into the activation of PS through the implementation of copper oxide-loaded biochar was performed. The outcome resulted in the effective breakdown of ROPs. In research by Luo et al. [102], the effectiveness of various metal-supported biochar nanocomposites (CuO/BC, Fe3O4/BC, ZnO/BC) was examined in catalyzing the degradation of Bisphenol A (BPA) (Figure 2a). The reaction pathway, primarily dominated by non-degradation, was elucidated based on the intermediates involved in electron transfer within the catalytic system (Figure 2b).

3.3. Fe-Based Catalysts

Due to their cost-efficiency, environmental friendliness, and superior performance, iron-based materials are widely used in various catalytic applications. Common Fe-based activators for PMS include zero-valent iron (ZVI), ferric oxide, and ferrous oxide. Aggregation poses a significant challenge for nanoscale particles such as nano ZVI, owing to their heightened surface energy and inherent magnetic interactions. If the nanoparticles accumulate in the biochar, the BET and the reactivity of the catalytic reaction will be reduced [103]. In addition, the oxidation of ZVI further reduces its reactivity [104]. The preparation of biochar loaded with ZVI can not only prevent the aggregation behavior, but also improve the effect of pollutant degradation [105]. Previous research has documented the capabilities of biochar/ZVI nanocomposites in eliminating substances such as nonylphenol [104], decabromodiphenyl ether [106], and bisphenol A [107]. Luo et al. found that [108] biochar/ZVI nanocomposites are characterized by oxidation resistance, reusability, and stability. The process of activating PMS with ZVI, as illustrated in Equations (46)–(49), has the capability to produce •OH and SO4• [109,110]:
Fe0 + HSO5 + 2H+ → Fe2+ + H2O + HSO4
Fe2+ + HSO5 → Fe3+ + SO4• + OH
Fe3+ + HSO5 → Fe2+ + SO5• + H+
Fe0 + HSO5 → Fe3+ + •OH + SO42−
Catalyst separation and recovery can be facilitated by loading biochar with Fe3O4 [111], γ-Fe2O3 [112], or CoFe2O4 [113], thus achieving magnetization. The ease of separation and high catalytic activity for persulfate activation are enhanced by the pyrolysis-prepared magnetic biochar, which contains encapsulated iron oxide nanoparticles. Rong et al. [112] synthesized Fe2O3 biochar (γ-Fe2O3-BC) using banana peel as a precursor material by employing hydrothermal technology, activated PS, and degraded BPA. In a similar study, Dong et al. [114] developed a composite material by loading Fe3O4 particles onto biochar. Furthermore, Ouyang et al. [115] synthesized a nanomagnetite biochar composite material, also incorporating biochar loaded with Fe3O4 particles.
Surface iron sites on biochar play a crucial role in Fenton-like systems. Biochar derived from sludge, as prepared by Li et al. [116], features abundant Fe2O3 and Fe2+ sites. At pH = 4.0, the synergistic effect of adsorption and the H2O2 activation process allowed the removal rate of ciprofloxacin to reach 90%. Furthermore, diverse iron sites like Fe0, Fe0.95C0.05, FeAl2O4, and Fe3O4 possess the capacity to induce both homogeneous and heterogeneous Fenton-like reactions concurrently, thereby promoting the decomposition of contaminants. Gan et al. [117] prepared an iron-enriched biochar with Fe0 and Fe0.95C0.05 sites on its surface, which induced Fenton reactions by means of leached Fe2+. Meanwhile, the FeAl2O4 and Fe3O4 sites directly engaged in H2O2 activation through the Fe2+ present on the surface. Additionally, iron sites enhance the stability of sludge-derived biochar. Zhang et al. [52] prepared magnetic biochar composites (MBCs) by mixing bio-sludge with iron-based sludge. The reinforced chemical bonding between biochar and Fe3O4 significantly enhanced the stability and reusability of MBCs, achieving a 98% removal rate for methylene blue after five cycles.
The Fenton-like reaction mechanism of iron-rich biochar can be delineated as follows: (1) Surface multivalent iron sites directly activate H2O2 to produce •OH or, under acidic conditions, leached Fe2+ on the surface activates H2O2 to generate •OH, as shown in Equations (50)–(53) and (7); (2) H2O2 can be activated by the single-electron transport mechanism of PFRs to produce •OH; (3) within a photo-Fenton system, the illumination facilitates the degradation of H2O2 and the subsequent production of •OH, concurrently promoting the regeneration of Fe2+, as presented in Equations (54) and (55); (4) Fenton-like systems mainly utilize •OH as the main free radical-degradation pollutant. Furthermore, biochar made from sludge possesses a porous composition and expansive surface area, enabling it to efficiently absorb various pollutants.
Fe0 + 2H+ → Fe2+ + H2
Fe2+ + H2O2 + H+ → Fe3+ + H2O + •OH
Fe0 + 2Fe3+ → 3Fe2+
Fe3+ + H2O2 → Fe2+ + HO2• + H+
Fe3+ + HO2• → Fe2+ + O2 + H+
H2O2 + hν → 2•OH
Fe3+ + H2O2 + hν → Fe2+ + HO2• + H+

3.4. Mixed-Metal Catalysts

Remarkable focus has been placed on the synergistic mechanism of multi-metal biochar due to its exceptional physicochemical characteristics, remarkable stability, and minimal metal leaching, surpassing that of biochars with only one metal. In multi-metal-loaded biochar, the catalytic efficiency is enhanced through electron transfer between metals. Recently, the development of bimetallic oxide biochars, such as AXB3-XO4-type materials (e.g., CuFe2O4, CoFe2O4, and CoMn2O4), has been explored for advanced applications in persulfate oxidation due to their stable spinel structures, synergistic effects, and stability [118]. Biochar impregnated with ferrite (MFe2O4, where M denotes a transition metal) possesses the capability to ameliorate the downsides associated with metal depletion and particle clustering. Moreover, it effectively tackles the concern of catalyst segregation [119]. Studies have demonstrated the effectiveness of composite materials like CoFe2O4/biochar and CuFe2O4/biochar as activators for the persulfate degradation of BPA [113], isoproterenol [120], and 4-nitrochlorobenzene [121].
Hao et al. [122] discovered a synergistic interaction between iron and manganese ions in Fe-Mn oxide biochar. Initially, Mn primarily governs the reaction, while Fe later assumes the forefront in activation processes. In studying the degradation of tetracycline, the literature has also delved into the application of Mn-fortified magnetic biochar [123] for PDS activation. Additionally, Co3O4-SnO2/BC was used as catalyst to activate PS to degrade thiazole sulfonate [124].
Bimetallic biochar catalysts have been utilized in Fenton-like reactions, exhibiting collaborative impacts on catalyzing hydrogen peroxide (H2O2) [125]. Taking iron-copper-modified chili straw biochar (CuFeO2/BC) as an example, the catalytic degradation of tetracycline by CuFeO2/BC using H2O2 was investigated. The catalysis of •OH production from H2O2 can be facilitated by Fe2+ and Cu2+. Cu2+ undergoes reduction when it reacts with H2O2 and •O2, resulting in the production of Cu+. This process of reduction is intriguing because the standard potential reduction of Cu2+/Cu+ is less than that of Fe3+/Fe2+. Consequently, this allows for the regeneration of Fe2+ via the reaction between Fe3+ and Cu+ [126]. The reaction rate can be accelerated by constructing a suitable bimetal-modified biochar catalyst, which facilitates the cyclic interaction between metals of high and low valence and promotes the production of •OH.
Various transition metals, with a specific emphasis on cobalt, are incorporated into the biochar structure to enhance its catalytic efficiency. The incorporation of multiple metals is a highly promising approach for modifying biochar, offering various benefits including enhanced synergistic effects, minimal metal leaching, and excellent stability. Studies have demonstrated that incorporating metal nanoparticles into biochar can greatly enhance its catalytic performance. Nevertheless, attaining accurate regulation of the metal content during the preparation process remains a major obstacle. Additionally, there is a need to strengthen research on the mechanisms of synergistic effects among different metals.

4. Heteroatomic Doping

The introduction of heteroatoms into biochar is a feasible method to improve its catalytic efficiency. By introducing C, N, P, S, and other heteroatoms into the carbon lattice to replace the carbon atoms on biochar, the properties of biochar are changed [127,128,129]. The presence of heteroatoms causes shifts in the electron density, boosts electron transportability, raises the number of defect edges, and introduces fresh active locations. As a result, catalytic electron transfer reactions can be expedited. There are typically two doping methods: in situ doping and diffusion doping. The process of in situ doping includes the direct carbonization of biomass precursor materials that contain heteroatoms, resulting in the introduction of heteroatoms into the carbon material [130,131]. On the contrary, diffusion doping involves exposing carbon materials to gases or liquids containing heteroatoms in high-temperature or high-pressure environments, allowing heteroatoms to diffuse into the carbon matrix, thus creating carbon materials with heteroatom doping [132,133].

4.1. Nitrogen Doping

The manipulation of charge distribution and spin density via nitrogen (N) incorporation improves the method in enhancing the catalytic performance of biochar that is initially unreactive [134]. Various external methods for nitrogen doping have been reported, involving different nitrogen sources such as urea [135], melamine [136], dicyandiamide [137], thiourea [138], ethylenediamine [139], polyacrylamide [140], and nitrogen-containing compounds like ammonium nitrate [141], ammonium chloride [142], and ammonium phosphate [143]. Xu et al. [136] conducted an experiment to fabricate biochar with nitrogen doping through a one-step calcination process by utilizing various nitrogen sources. The experimental findings indicated that the nitrogen levels in the initial material significantly impacted the catalytic efficiency of the resulting biochar. Liu et al. [120] applied melamine as the nitrogen supplier and implemented a liquid-phase doping strategy to produce magnetically retrievable biochar doped with nitrogen. After being calcined at a temperature of 800 °C, the biochar displayed an outstanding surface area and graphitization performance, leading to its exceptional catalytic activity and stability. Solid-phase doping is a method that does not rely on solvents and instead involves grinding nitrogen sources and carbon materials together. The resulting nitrogen-doped biochar can be obtained through subsequent calcination under inert atmospheres. Ye et al. [144] created nitrogen-doped biochar by grinding biochar and impregnating it with K2FeO4 and urea. Figure 3 provides an illustration of their methodology. A structural analysis through XRD, Raman, and XPS spectra confirmed the graphite carbon structure and successful nitrogen doping into biochar.
An effective strategy for in situ nitrogen doping, which does not require the use of nitrogen sources, entails directly pyrolyzing precursors containing abundant nitrogen. Natural nitrogen precursors, such as proteins found in residue from spirulina, can be utilized for this purpose. A separate investigation carried out by Xie et al. [145] centered on the synthesis of biochar with N-doping, utilizing Fusarium, which is abundant in nitrogen, as the precursor. Moreover, achieving the accurate management of N-doped biochar characteristics by means of in situ N-doping poses a noteworthy obstacle that necessitates more thorough examination and study.
Biochar containing nitrogen, which has been prepared in advance, showcases a porous structure, abundant defects, a substantial number of nitrogen-containing substances, and a large surface area. These characteristics contribute to its exceptional catalytic performance in degrading ROPs. The size resemblance between carbon atoms and graphene enables the effortless integration of nitrogen atoms into the graphene sheet structure. Furthermore, due to nitrogen’s higher electronegativity compared to carbon, carbon atoms situated next to nitrogen atoms could potentially act as Lewis basic sites [146]. The electronegativity difference between nitrogen and the surrounding carbon accelerates electron transfer and contributes to the formation of defect structures [8,134,147].
Biochar contains dominant nitrogen functional groups, namely pyridine-N and pyrrole-N, structurally resembling nitrogen found in graphene. In a research project led by Yuan et al. [148], the advantageous impacts of graphene-N and pyridine-N, enhancing the electrocatalytic performance of graphene, were investigated. Wang et al. [134] used biochar obtained from different combinations of corncob and urea as a means of PDS activation to decompose sulfadiazine. The researchers found that the nitrogen-doped adaptive biochar was four times more catalytic than before doping. This progress can be attributed to the binding of different sites of active nitrogen (such as pyrrolo-N, pyridine-N, and graphite-N) or the nitrogen functionalization of biochar (including N oxides and amine functional groups), which enhance its catalytic capacity. In the subsequent year, this research group utilized sludge as a raw material to produce sulfurized sludge biochar (SSB) at a temperature of 700 °C and conducted a comparative examination of the activation of PDS and PMS. According to reports [144,149], carbon atoms have a larger radius than graphite-N, with lower electronegativity, allowing electrons to transfer from less electronegative carbon to more electronegative nitrogen. At this juncture, carbon turns electron-deficient and acquires electrons from PMS. The extraction of electrons from PMS results in the formation of SO5•, and subsequently, it can undergo a reaction with water, leading to the production of 1O2 (Equations (56) and (67)).
HSO5 →SO5• + H+ + e
2SO5• + H2O → 1/2 1O2 + 2HSO4
Moreover, due to the larger ionic radius of nitrogen (N) compared to oxygen (O), nitrogen atoms are superior to other atoms, and the beneficial effect of nitrogen atoms is most pronounced in improving visible-light catalytic performance [150,151,152,153]. Li et al. [154] used phenolic resin microspheres as a template and employed a straightforward hydrothermal technique to fabricate nitrogen-doped TiO2 hollow microspheres containing oxygen vacancies. The experimental findings exhibited the significance of oxygen voids and demonstrated how the existence of these voids within hollow microspheres can augment the photocatalytic fixation of nitrogen.

4.2. Sulfur Doping

Sulfur (S) is abundant, and sulfides exhibit diverse structures, providing excellent conditions for the structural design of biomass-based carbon materials. The electronegativity value of sulfur (2.58) is comparable to that of carbon (2.55). Consequently, the inclusion of sulfur atoms into carbon materials does not influence the distribution of electric charge within the atoms. Nevertheless, as sulfur atoms possess a larger atomic radius, the process of doping introduces a higher number of imperfections into the carbon structure, ultimately leading to the reorganization of electron spins [155]. The inclusion of sulfur in the carbon structure causes a disruption in the arrangement of carbon atoms. This disruption leads to the creation of patterns known as thienyl sulfur “-C-S-C-”. Additionally, the presence of sulfur atoms modifies the sp2 conjugation of biochar, thereby ameliorating the electron transfer characteristics of the biochar substance.
The primary method for activating PS in carbon materials doped with sulfur involves using an in situ sulfur-doping strategy. At elevated temperatures, the pyrolysis technique is employed to convert polymers containing sulfur into S-doped carbon materials. Wang et al. [80] illustrated the remarkable ability of sulfurized biochar to effectively eliminate bisphenol A by activating PS. The S-doped biochar produces sulfur atoms at the edges of the biochar, which activates the SP2-hybrid graphene structure, and produces Lewis acid and Lewis base sites within the S-doped biochar. Yaglikci et al. [156] utilized tea as the primary ingredient and sodium thiosulfate pentahydrate as the sulfur dopant in the production of S-doped biochar. Their findings indicate a significant enhancement in the specific capacitance of the biochar post-S modification, showing a substantial increase in comparison to the baseline. This amplification can be ascribed to the incorporation of sulfur atoms, serving as electron donors within the framework and subsequently enhancing the conductivity. Consequently, the introduction of sulfur through the doping process exhibits a favorable impact on the electrochemical capabilities of supercapacitors. Ding and colleagues [138] showcased the positive impact of N-doping on the degradation of isoproturon by activating PMS, whereas the inclusion of S had a negative influence. Additionally, the research revealed that N atoms could substitute sp2 carbon atoms within the carbon framework, whereas S atoms were prone to replacing groups containing oxygen in the carbon structure. The differences observed in sulfur-doped biochar may stem from the varying synthesis methods employed. Furthermore, further studies are necessary to examine the production of biochar doped with sulfur and understand its catalytic mechanisms.

4.3. Boron Doping

Boron is a non-metal found in the third main group of the periodic table. It has three valence electrons and can form bonds with carbon and oxygen. Because of its low electronegativity and similar size to C atoms, boron is considered one of the most popular modified elements in carbon materials. When boron enters the carbon lattice, it forms not only B-C bonds, but also B-O bonds. This modification boosts the transfer of e from the surface of carbon substances, potentially enhancing the electrical conductivity and characteristics of carbon-derived materials [157]. The addition of boron introduces oxygen onto the surface of carbon, enhancing the electronic plane and improving the chemical stability of carbon. In contrast to N-doped carbon, B-doped biochar exhibits prolonged endurance owing to the remarkable stability of boron sites. Studies have shown that B-doped biochar can activate PDS to eliminate sulfamethoxazole [158]. Liu et al. used boric acid and wheat straw to synthesize B-doped biochar by pyrolysis at 900 °C, and evaluated the improvement in the SMX degradation rate [158]. The inclusion of B in biochar enhances its surface affinity for PDS, making it more effective as a Lewis acid site. In addition, B also impacts the electron structure of biochar, leading to an increased electron transfer rate and improved catalytic performance. Both experimental and theoretical calculations confirm these findings.

4.4. Other Atoms Doping

Introducing multiple heteroatoms into biochar showcases the potential synergistic impacts of heteroatoms on improving the PMS activation performance. The investigation of co-doped carbon materials predominantly revolves around incorporating an additional heteroatom onto nitrogen-doped materials. Given that both P and N belong to Group V and are located in the third period, it is observed that the P atom demonstrates a greater atomic radius along with enhanced electron-donating capabilities compared to N. Consequently, this imparts benefits for the modification of the carbon network. Phosphorus, being widely abundant in nature, exhibits promising prospects for non-metallic doping. Consequently, potential elements for co-doping carbon materials encompass nitrogen-sulfur, nitrogen-boron, and nitrogen-phosphorus, among other options.
Currently, the field of PS-AOPs primarily depends on a co-doping method to alter oxidized carbon nanotubes and graphene. In contrast, the exploration of biochar modification remains comparatively restricted. Co-doped biochar exhibits superior catalytic activity compared to biochar doped solely with N or S [138]. In their study, Wang and colleagues [159] conducted an investigation into the synthesis of biochar co-doped with cobalt, sulfur, and N. Through their research, they made a significant observation that the reaction rate of biochar co-doped with sulfur and nitrogen exhibited a remarkable 4.5-fold increase compared to biochar solely doped with nitrogen.

4.5. Metal and Non-Metal Based Biochar

In a recent investigation by Li et al. [160], research was carried out on the domain of PMS activation, wherein the authors introduced the production of biochar co-doped with nitrogen and iron. This research utilized wheat straw, ferrous sulfate, and urea as starting materials to supply Fe and N. The findings indicated that the introduction of iron greatly enhanced the activation efficiency of PMS, resulting in the catalyst becoming magnetic and facilitating its easy separation and retrieval. In the realm of photocatalysis, solely relying on non-metal doping fails to produce remarkable outcomes. Therefore, researchers have attempted to modify TiO2 with different non-metal elements and reducing agents to enhance its response to visible light. Wang’s team investigated the impact of hydrogen reduction on TiO2 photocatalytic materials doped with nitrogen and fluorine for a visible light response [161]. Atomic spin resonance spectroscopy revealed that the Ti3+ and oxygen hole positions in the lattice produced by hydrogen reduction were the same as the substitution positions of nitrogen and fluorine elements, indicating that hydrogen reduced the nitrogen and fluorine in the lattice. The improved photocatalytic effectiveness of the TiO2 with reduced visibility can be credited to the defects formed within the substance following reduction, resulting in disrupted energy bands and chaotic formations. Besides Ti, copper (Cu) can also be co-doped with non-metals. Zhong et al. [162] prepared nitrogen-doped Cu-biochar (N-Cu-BC) composites. The findings revealed that N-Cu-BC showcased a superior electron transfer capability when compared to Cu-biochar (Cu-BC), implying that the introduction of nitrogen can augment catalytic activity through the enhancement of the catalyst’s electron transfer capability. Exploring further non-metal modifications in the Cu-BC system is a highly promising direction for future research.
Nitrogen-doped biochar is the most extensively studied among the various types of heteroatom-doped biochars. While research in this area is still developing, the co-doping approach demonstrates promise in increasing the efficiency of AOPs’ activation. The successful manipulation of biochar’s composition through the addition of diverse atoms is an effective method for regulating it, thereby providing benefits for the degradation of ROPs without the need for radical involvement. Moreover, by incorporating heteroatoms into biochar, the risk of introducing harmful heavy metals that may harm water quality is effectively avoided. Nevertheless, the accurate creation of heteroatom-doped biochar presents considerable difficulties, and our precise understanding of the catalytic mechanisms involved is still uncertain and occasionally contradictory.

5. Catalytic Mechanisms

A large number of scholars have studied the use of biochar-based catalysts to degrade various persistent ROPs through AOPs. However, due to variations in experimental conditions, instruments, and operations, direct comparisons of their effects are not rigorous and reliable. Nevertheless, the findings confirm the potential for biochar-based materials to catalyze the degradation of various ROPs in AOPs. At the same time, many studies have made great progress in the in-depth exploration of the catalytic mechanism of BC-AOPs [163,164,165]. The decomposition of BC-AOPs can be categorized into two pathways: one involving free radicals and the other involving non-free radicals. The free radical pathway initiates charge transfer to generate polar oxidizing entities, activating specific sites such as surface functional groups, PFRs, π-π bonds, hydrogen bonds, defective sites, and transition metals [166,167]. In contrast, there are approximately three non-radical pathways, mainly including electron transfer, 1O2 generation, and surface activation [168,169].

5.1. Free Radical Pathway

5.1.1. Surface Functional Groups

The adsorption capacity of BC towards aromatic hydrocarbons is significantly influenced by the distribution and properties of surface functional groups. These functional groups include not only OCGs, but also N-containing and S-containing functional groups [170], and can vary in terms of their polarity, hydrophobicity, and distribution. The presence of positive/negative charges, as indicated by Tan et al. [171], can serve as a valuable indicator of the electron-donating/accepting properties of BC. OCGs are particularly abundant and have been highlighted as crucial structural features of biochar, influencing its adsorption and catalytic abilities [172,173]. During the adsorption process, various interactions occur between adsorbates and surface functional groups, especially when it comes to aromatic compounds and ionized aromatic pollutants such as herbicides and antibiotics [174]. The presence of surface charge and hydrogen bonding can improve the adsorption capabilities of BC, leading to a decrease in pollutant levels.
According to several studies [175,176], OCGs are the main active site of activated PS on biochar’s surface. Yan et al. [177] pointed out that biochar with OCGs like -OH and -COOH had the potential to serve as a platform for electron transfer and the stimulation of PS. Huong et al. [178] also found that OCGs containing C=O, -COOH, and phenol groups promoted the formation of free radicals. It is suggested that OCGs may accelerate the direct decomposition of PS and enhance the activation of PS through metal mediation. Additionally, Meng et al. [179] prepared OCG-containing biochar using orange peel as a precursor system. They found a direct relationship between the pyrolysis temperature and the degree of graphitization, the BET, the mesopore ratio, and the content of C=O. The enhancement discussed in this passage is related to improving the adsorption capacity and activation, which consequently results in a higher efficiency in catalyzing PMS. Quinone and hydroquinone groups, present on the biochar surface, operate as redox pairs. They facilitate the exchange of electrons between external oxidants and the reductant [180]. Moreover, OCGs can serve as electron mediators to transmit electrons to H2O2, producing extremely reactive •OH, which assists in the breakdown of organic contaminants.
In addition, the process through which ROSs are generated by OCGs resembles that of dye-sensitized semiconductor photocatalytic systems. The surface of biochar contains OCGs, which can function as a photosensitizer, while biochar particles act as electron shuttles. When OCGs are excited by light, electrons are excited and subsequently transferred from OCGs to an electron acceptor (like O2), resulting in the production of •O2. Similar to the •O2 generated previously, it can further accept electrons or absorb H+ to form H2O2, leading to the production of •OH [181]. The primary photochemical reactions are as follows (Equations (58)–(61) and (54)):
BC-OCG + hv → BC-OCG*
BC-OCG* + O2 → e + BC-OCG+
2•O2 +2H+ → 2HO2• → H2O2
HO2• → e + HO2 +H+ → H2O2
H2O2 + hv → •OH
Functional groups in biochar significantly boost its polarity, improving the interaction between pollutants and polar adsorbent materials [182]. Moreover, these functional groups have the potential to catalyze redox reactions and activate PMS. Table 3 shows the N-containing functional groups on the surface of biochar post-pyrolysis, predominantly consisting of pyridine N, N-amine, and potential pyridine n-oxides [183]. The catalytic activity of the oxygen reduction reaction is related to specific nitrogen configurations, namely pyridine N, graphite N, and pyrrole N, and the presence of amine N can adsorb heavy metals and CO2 [184]. Additionally, incorporating and aligning nitrogen functional groups within biochar can elevate its alkalinity, ultimately augmenting CO2 capturing through Lewis acid–base interactions [185]. These interactions play a vital role in adsorbing acidic gases [186]. The presence of pyridinic N has the ability to transform nearby carbon atoms into Lewis base sites, which promotes reactions with oxygen. This, in turn, improves the electrocatalytic redox capabilities of biochar. On the other hand, by combining graphite N and pyridine N, the electrocatalytic performance is further improved since the carbon atom adjacent to graphite N also acts as the active site for the reaction with oxygen. In terms of catalytic PMS activation, SO4• has been identified as an efficient oxidant for the effective removal of ROPs. They exhibit a longer lifespan and better selectivity compared to •OH [7,8]. By utilizing nitrogen-containing biochar catalysis, the efficiency of the removal of pollutants such as Orange G [141], sulfamethoxazole [134], hyaluronic acid [187], and tetracycline [188] can be significantly enhanced. Pyridine N, pyrrole N, graphite N, and N oxide are recognized as N-containing active groups and have shown remarkable efficacy in electrocatalysis. For instance, the improved elimination ability of sulfamethoxazole is a result of the heightened levels of pyridine N and pyridine N within the catalyst, whereas graphite N exhibits an enhanced catalytic performance when eliminating orange G [141]. Similarly, the removal efficiency of tetracycline is associated with pyridinic N and graphitic N [8]. Extensive research must be conducted on the effective nitrogen functional groups in biochar to enhance its catalytic performance. The mechanisms of action of these functional groups need to be understood in order to achieve this improvement. Biochar contains various sulfur-containing functional groups, such as organic sulfides, sulfoxides, sulfides, thiophenes, sulfones [189,190,191], C-SOx-C [192], S=O [193], C=S, -S-S-, C-S, and -SH [194]. By altering the BET and porosity of biochar, these functional groups can improve its adsorption capacity [195]. Additionally, biochar that contains sulfur-containing functional groups has the capability to chemically adsorb mercury and transform it into mercuric sulfide [190]. On the other hand, low-sulfur or sulfur-free biochar primarily relies on physical adsorption for mercury removal as it lacks chemically active sites that can effectively accommodate mercury [196]. Biochar that contains sulfur effectively absorbs Hg2+ in the soil, thanks to the presence of bacteria that reduce sulfate. For example, the adsorption of Hg2+ by biochar doped with elemental sulfur in soil reached an astonishing 67.11 mg·g−1 [197]. Under anaerobic conditions, the adsorbed Hg2+ can be converted into HgS. In the soil, sulfate-reducing bacteria efficiently absorb the HgS that is generated. Subsequently, this HgS undergoes methylation, effectively separating Hg2+ from the soil [198]. The adsorption performance of biochar for Cd is enhanced by functional groups like C-S and C-S, as well as -SO3H [199], sulfates, sulfides [200,201], organic sulfides, and sulfite [202]. When Cd removal from soil is targeted using sulfur-containing biochar, both biochar and sulfur-containing functional groups work synergistically in adsorbing Cd [203]. Sulfur-containing biochar not only adsorbs Hg2+ and Cd but also facilitates the adsorption of nickel ions. Biochar adsorbents were prepared by Higashikawa et al. using four different biomasses, namely rice husk, chicken manure, sawdust, and sugarcane bagasse. The adsorption capacities of these biochar adsorbents for Ni2+ were subsequently tested at different temperatures. The results showed that the maximum Ni2+ removal range was from 0.20 to 10.9 mg·g−1. Of the biochar varieties examined, chicken manure biochar exhibited the most effective pyrolysis capacity for Ni2+ removal at 650 °C, achieving a rate of 10.9 mg·g−1 [204]. The positive role of S-O and -SOH in Ni2+ adsorption is notable. This phenomenon may be attributed to the interaction between weak acids (such as heavy metals) and weak bases (S), leading to an enhanced affinity of S towards heavy metals [205]. Consequently, the formation of metal sulfides such as PbS, Ag2S, Cu2S, CuS, and ZnS becomes more facile.
Sulfur-doped biochar has the ability to generate sulfur-containing functional groups. These groups of functions can not only function as sites for adsorption, but can also act as active catalytic sites for activating PS. Specifically, the active site on the surface of the catalyst binds to PMS, leading to the transfer of e from the catalyst to PMS. This results in the breakdown of PMS to generate hydroxyl radicals or sulfate radicals. Huang et al. [206] prepared the electron-donating thiophene group (SACx) by ball milling by activating PMS during the decomposition of diethyl phthalate. The findings indicate that S has been effectively infused into the sp2 hybrid biochar framework through the creation of electron-rich thiophene groups (C-S-C), which is recognized as the primary active location for PMS stimulation. Additional research verifies that •OH radicals are generated rather than SO4•, and the quantity of •OH radicals is directly linked to the quantity of C-S-C linkages. The proposed theory suggests that, as a result of ball milling, the amalgamation of biochar and sulfur forms C-S-C, which in turn reacts with reactive O atoms in the O-O connection of PMS, resulting in the formation of free radicals (Equations (62) and (63)). Additionally, C-S-C [207] can cycle to generate SO4• (Equation (64)) through the interaction between SACx+ and another PMS, similar to the Fenton reaction. This causes the contaminant to be degraded by the generated •OH and SO4•, and the C-S-C groups can compete with the contaminant for these active substances. In addition to the C-S-C bond, the S-thiophene group located on the catalyst surface can also activate PS. Jin et al. [208] synthesized S-doped biochar rich in iron FeS2-BC by a combination of ball milling and pyrolysis. When PS is activated, the activation is facilitated by the collaboration of surface ketone groups, Fe(II), and highly reactive S-thiophene groups. The electron-donating S-thiophene groups, as well as S2−/Sn2− groups and ketone groups, assist in the regeneration of Fe(II) in both the lattice and solution by serving as electron donors in direct or indirect electron transfer processes.
SACx + HSO5 → SACx+ + •OH + SO42−
SACx + HSO5 → SACx+ + OH + SO4
SACx+ + HSO5 → SACx + H+ + SO5
Table 3. Biochar with nitrogen functional groups.
Table 3. Biochar with nitrogen functional groups.
BiomassPyrolysis Temperature (°C)N-Functional GroupsRefs.
Chlorella vulgaris600–900Pyrrolic-N, quaternary-N, pyridinic-N [209]
Phragmites australis450Pyridinic-N, pyrrolic-N [142]
Nannochloropsis sp. Spirulina platensis Enteromorpha prolifera400–800Pyridinic-N, pyrrolic-N, quaternary-N [210]
Bamboo600Pyridinic-N, pyrrolic-N, quaternary-N, [211]
Wheat straw300–800Pyridinic-N, amine-N, pyrrolic-N, quaternary-N, NH4-N [212]
Corn straw600Pyridinic-N, pyrrolic/pyridonic-N, graphitic-N, [213]
Corn straw600–800Pyridinic-N pyrrolic-N pyridonic-N, graphitic-N [214]
Straw300–400Nitrile-N, pyridinic-N pyrrolic-N, amine-N NH4-N, NO3-N NO2 -N [215]
A Mixture of sewage, cattle manure and eucalyptus wood chips250–550NH4-N, amine-N [216]

5.1.2. Role of PFRs

PFRs are proposed as a contrast to short-lived free radicals (transient radicals). Transition metals facilitate the formation of these PFRs by accepting electrons transferred from organic compounds such as phenols and quinones [217,218]. Previous research indicates that the organic composition and metal elements present in biochar have an impact on the types of PFRs that are formed. Metal elements can affect the formation of PFRs [219,220]. There are three main types of PFRs in biochar: carbon center PFRs, carbon-centered radicals with adjacent oxygen atoms, and oxygen-centered PFRs [221]. The presence of PFRs in biochar is affected by many aspects, including the biomass type, sample size during pyrolysis, heating rate, residence time, etc. [222,223].
According to research reports, it has been observed that the inclusion of PFRs in biochar can effectively generate electron sources, thereby stimulating the activation of different compounds including H2O2, O2 [224], and PS [225]. Then, •OH, O2•, 1O2, and SO4• are generated. The efficiency of ROS production is influenced by both the concentration and type of free radicals, as well as the presence of PFRs. In Fenton-like systems, PFRs with carbon-centered structures are considered active centers. Yu et al. made a significant discovery regarding carbon-centered PFRs found in biochar. They observed that these PFRs have the capability to transfer electrons to oxygen molecules that are adsorbed, consequently leading to the creation of O2• (Figure 4b) [226]. It is speculated that PFRs with oxygen-centered structures may serve as active centers for •OH generation, while carbon-centered PFRs could be responsible for the production of O2•. In addition, apart from their involvement in the e-transfer procedures for producing PFRs, PFRs also have the ability to directly interact with contaminants. In wheat straw-derived biochar, PFRs with a carbon center have the ability to serve as donors of electrons, facilitating the transfer of e between the surface of the biochar and Fe3+, thus activating electrons in persulfate and producing Fe2+ [227].
Furthermore, the preparation parameters greatly impact the amount and composition of PFRs, encompassing variables including the temperature and duration of pyrolysis [224,225,228]. Due to the presence of biochar-PFRs, hydrochar can also generate ROSs. Chen et al. [181] systematically evaluated the photochemical characteristics of hydrochar and pyrolysis char. Their findings revealed that both forms of biochar predominantly consist of PFRs centered around oxygen, yet pyrolysis char exhibited a greater abundance of PFRs. The pig manure biochar SBC600 at moderate temperature exhibited a higher PFR content, thus showcasing a more pronounced ability to activate H2O2, where oxygen-centered PFRs played a dominant role in electron transfer [220]. Regarding residence time, Wang et al. [229] emphasized that the peak proportion of PFRs was detected following a residence duration of 4 h.
Various scientific investigations have extensively recorded that the production of PFRs is regulated by the existence of phenolic compounds and transition metals in the feedstock [225]. The amount of PFRs is dictated by the phenolic compounds’ composition, whereas the development of PFRs is considerably impacted by transition metals. Remarkably, PFRs with the central presence of oxygen demonstrated superior catalytic efficacy in the PS activation mechanism when contrasted with PFRs centered on carbon [224]. Adding Fe3+ as a dopant boosts the generation of oxygen-centered PFRs, like semi-quinone radicals, and adjacent carbon-centered oxygen-centered PFRs, like phenoxyl radicals. Additionally, Deng and colleagues discovered that only oxygen-centered PFRs act as sites for generating hydroxyl radicals (Figure 4a). Chen Zhang [220] chose pig manure with a higher content of metal elements and organic matter as a raw material to prepare biochar, and activated hydrogen peroxide to degrade SMT. The resulting biochar contained a higher concentration of PFRs, with oxygen-centered PFRs exhibiting a higher catalytic activity.
The electron transfer capabilities of PFRs are determined by their type. Compared with carbon-centered PFRs with oxygen atoms close together, carbon-centered PFRs have stronger electron donating capacity [230]. On the other hand, PFRs centered on oxygen demonstrate a superior catalytic activity in comparison to carbon-centered PFRs with neighboring oxygen atoms as well as carbon-centered PFRs [231,232]. Hence, it becomes essential to regulate the elemental composition of biochar catalysts to optimize their catalytic performance.

5.1.3. π-π Bonds and Hydrogen Bonds

Non-covalent interactions, known as π-π interactions, pertain to the bonding phenomenon observed between planar functional groups and aromatic molecules. These interactions elucidate the mechanism underlying the binding of biochar to ROPs, specifically those featuring benzene rings or C-C functional groups (Figure 5) [171]. Meanwhile, biochar with aromatic rings exhibits a strong electron-donating capability, forming cation-π interactions with TC [233].
The interaction between electron acceptors and donors (EDA) enhances the adsorption of positively charged compounds on the surface of biochar. Moreover, the spatial structure of biochar can be further improved by π-π interactions, increasing its efficiency in removing ROPs [234]. Tang et al. conducted research involving the use of biochar as an adsorbent to remove tetrabromobisphenol A (TBBPA). Following the adsorption process, changes in the C-C functional group were observed in the absorption peak at 1618 cm−1. This alteration indicates the existence of π-π electron interactions between TBBPA and biochar [235]. Zhou et al. further studied alkali acid-modified magnetic biochar. According to their observations, the strong adsorption of TC depends on the π-π stacking mechanism derived from the graphite structure. In addition, ketone and amino functional groups on biochar significantly enhance the electron acceptance ability of TC [236]. Carbon materials with an sp2 hybrid structure have abundant π electrons and are catalytic active sites in biochar [237]. Zhu and colleagues [232] discovered that the percentage decrease in π-π interaction was 1.22% both pre- and post-catalysis. This led to the conclusion that the conversion of π-π interaction from the aromatic ring to PDS results in the generation of SO4•. In addition, according to the observation of Xu et al. [238], the presence of extended electron pairs in pyridine N of sp2-C significantly enhances the transfer of π-π interaction to PMS. The primary process involves π-π interaction with the aromatic ring of ROPs, and the arrangement of ROPs and biochar structures will influence the adsorption capacity of the π-π bond. It should be noted that different organic pollutants may have different spatial structures under different environmental conditions. As such, external conditions like pH levels, temperature variations, and the presence of additional anions can modify the spatial arrangement of ROPs, thereby impacting the adsorption effectiveness of biochar towards ROPs. Hence, it is essential to conduct further research on the structure of ROPs and the correlation between the spatial configuration of biochar and its adsorption capabilities.
In addition to a π-π bond, intermolecular hydrogen bonds can also carry out an adsorption process. For instance, pre-treated biochar containing iron (SDBC) exhibits abundant OCGs, which establishes hydrogen bond interactions with the hydroxyl and amino groups of TC and doxycycline (DOX) during degradation [239]. Moreover, the phenol group of TC can also develop hydrogen bonds with OCGs, facilitating the adsorption process. Various environmental conditions may influence the hydrogen bond strength, with pH being particularly significant in altering the biochar surface potential and antibiotic structure. When both the catalyst surface and antibiotic bear negative charges, this negativity can boost adsorption via hydrogen bonding [240]. Furthermore, polar functional groups in biochar can engage in hydrogen bonding with pollutants as well [241]. Zhou et al. indicated that the interaction between OCGs (such as C-O, Fe-O, and Co-O) and hydrogen bonds on GO/Co-Fe2O4-SDBC can adsorb Imidacloprid (IMI). Fe and Co are responsible for connecting GO/Co-FFe2O4-SDBC and IMI [242].
Figure 5. Mechanisms governing the adsorption of tetrabromobisphenol A (TBBPA) utilizing SDBCs [243].
Figure 5. Mechanisms governing the adsorption of tetrabromobisphenol A (TBBPA) utilizing SDBCs [243].
Water 16 00875 g005

5.1.4. Defect Sites

The uniformity of the carbon lattice structure in carbon materials can be disrupted by defect sites, resulting in the presence of a high chemical potential and unpaired electrons. This imparting of a catalytic capability to defect sites has been observed [144]. The presence of defects, such as zigzag and armchair edges, results in the dispersion of π electrons that play a crucial role in the adsorption process and breaking of O-O bonds during PS activation [244]. To enhance the transformation of OCGs into graphite and defect sites, an ideal pyrolysis temperature must be maintained [245]. Research has indicated that increasing the pyrolysis temperature can cause the conversion of amorphous carbon (sp3-C) to graphitic carbon (sp2-C), which leads to the breakdown of the carbon structure and the formation of additional imperfections [246]. Nevertheless, an abundance of flaws in the structure can hinder both the mechanical integrity and catalytic function of biochar [123]. Defective structures in biochar, which have a high distribution of electron density, can boost electron transport and play a role in triggering the activation of PMS. A study by Ouyang et al. [247] found that imperfect structures like edge defects, curvations, or vacancies create suspended σ bonds. These bonds enable π electrons to move freely from biological carbon to PMS, ultimately leading to the formation of free radicals. According to Kim and Ko, the activation of persulfate depends primarily on the stratification of graphitized structures formed in biochar, defects, and the presence of delocalized electrons in basic groups [166]. Defect positions on the graphitized nitrogen and carbon network are also considered potential active sites for promoting pollutant degradation [144].

5.2. Non-Radical Pathways

Experts in scientific research have increasingly focused on non-free radical pathways, with particular emphasis on the integration of biochar and PS-AOPs. Non-radical oxidation, a surface reaction different from the predominant free radical oxidation in natural solutions, has garnered significant interest in recent years. The non-free radical pathway has several significant benefits: (i) the oxidation capacity of fully activated PS; (ii) self-quenching behavior involves the elimination of free radicals; (iii) minimizing interference from both inorganic and organic substances; (iv) wider operating pH levels; (v) potential selectivity for the degradation of specific contaminants [248].

5.2.1. Electronic Transfer

Biochar’s electrical conductivity (EC) is directly linked to both its graphitization degree and mesopore degree [249]. The presence of delocalized π electrons within the graphite layer leads to the creation of overhanging bond states at edges and defects, which in turn promotes rapid electron transfer [250]. Given that electron transfer is fundamental to the reaction process, biochar’s EC plays a crucial role in determining its catalytic activity. Serving as a unique “transporter for electrons”, biochar possesses a high EC, enabling the rapid transfer of electrons to facilitate the swift generation of ROSs and leading to the degradation of ROPs. Moreover, research has illustrated that the elevated EC of biochar supports the process of photoinduced electron transfer while impeding the recombination of electron-hole pairs (e-h+) [245]. Electrochemical impedance spectroscopy (EIS) and cyclic voltammetry (CV) can directly identify the presence of non-radical electron transfer pathways. Research conducted by EIS [251] demonstrated an enhancement in the electron transmission capability of biochar substrates after exposure to KOH and CaCl2 treatment. Through the mediation of electron transfer within the graphite framework, electrons are conveyed from natural substances to PMS, resulting in the degradation of ROPs via non-radical pathways [119].
Biochar surfaces can form metastable complexes (biochar/PS*) that can transfer electrons from ROPs to persulfate (PS). In turn, this process leads to the oxidation of ROPs without the production of ROSs [252]. Wang et al. [79] discovered that biochar plays a key role in enhancing electron transfer to photosystem (PS) at electron-abundant locations, thereby promoting the oxidation of sulfamethazine in adjacent electron-deficient sites. To ensure effective electron transfer, it is crucial that the activation potential of the biochar/PS* system is lower than the oxidation potential of the recalcitrant organic pollutants (ROPs). Stratified porous biochar prepared from shrimp shells by Yu et al. [246] has been used for persulfate activation to remove 2,4-DCP. The degradation mechanism of 2,4-DCP involves a direct double-electron transfer, with 2,4-DCP transferring electrons to either PS or O2 to generate SO4• and •O2.

5.2.2. Single Line State Oxygen

The first excited electronic state of O2, known as 1O2, is highly reactive. When in its excited form, 1O2 can return to its stable state of O2 without engaging in chemical interactions or electron transfer [253]. Understanding the complex mechanism behind the generation of 1O2 has been the focus of the recent literature. Certain researchers have suggested that the generation of 1O2 occurs through the activation of PMS with the presence of oxygen vacancies located on the biochar surface [254]. Furthermore, carbonyl groups and various metallic elements could potentially act as the sites responsible for the production of 1O2. The transformation of O2 to •O2 takes place at active sites like heterogeneous iron [255], soluble Fe2+ [256], and phenol-OH [257,258], followed by subsequent conversion to 1O2. The exact process of 1O2 formation is unknown, but it may be influenced by OCGs, the defect structure, and variable valence metal ions. Particularly, different structures of oxidants may have different effects on the formation mechanism of 1O2, such as the different structures of PMS and PDS, which produce 1O2 in different ways. Moreover, under certain conditions, free radicals may also be transformed into 1O2 [259]. In the case of persulfate activation, there are four possible pathways for 1O2 generation. First, applying the C-O of biochar can produce 1O2, as shown in Figure 6. The Lewis active center of C-O facilitates electron transfer, enhancing the electron density in adjacent carbon rings. This transfer aids in the cleavage of PS’s O-O bond to generate 1O2 [260]. Meng et al. [179] verified through density functional theory (DFT) that the catalytic active center for activating PMS and generating 1O2 is C-O located at the edge. Second, •O2, as a precursor, can also generate 1O2. •O2 reacts with H+ or H2O to form 1O2. It can also combine with SO4• or self-combine to yield 1O2 (Equations (65)–(68)) [261,262]. Thirdly, PS has the ability to transfer electrons to the positively charged carbon atom found in biochar. This process triggers the creation of the superoxide radical anion SO5•. Subsequently, SO5• reacts with H2O to generate 1O2 (Equations (69) and (70)) [80,263]. Fourth, PMS has the ability of self-decomposition to 1O2 [260]. Additionally, Li et al. [264] indicated that iron-doped biochar can make H2O2 to produce 1O2 to degrade tetracycline. And •O2 is a precursor to 1O2 production (Equations (71)–(73)).
•O2 + H+1O2 + H2O2
2•O2 + 2H2O → 1O2 + H2O2 + 2HO
•O2 + SO4•→ 1O2 + SO42−
•O2 + •O21O2 + H2O2
HSO5 + SO5•→ H+ + e
SO5• + H2O → HSO4 + 1O2
•O2 + H2O21O2 +OH + •OH
2H2O21O2 + 2H2O
2•O2 + 2H+1O2 + H2O2
Some studies have shown that 1O2 is the main pathway of ROPs’ degradation in PS and Fenton-like processes. The predominant oxidation pathway observed in PDS systems to eliminate SMX involves non-radical oxidation, wherein the main species is 1O2. The 8N atoms of SMX and the benzene ring are more attractive to 1O2 oxidation, leading to the decomposition of SMX through the non-radical oxidation pathway [265]. Xu et al. observed that, when using nitrogen-doped biochar, •OH, SO4•, and 1O2 are generated to activate PMS, with 1O2 making a more significant contribution to organic degradation in N-biochar-activated PMS. The main cause of the biochar activation of PDS is the presence of OCGs, especially quinones, and transition metal ions such as iron [136]. Several studies also propose that the composition of 1O2 serves a distinct role in pollutant degradation through different ROS functions in SDBC. The generation of 1O2 is intricately linked to the existence of oxygen that is dissolved within water. Consequently, in degradation procedures led by 1O2, modifying the quantity of dissolved oxygen might have a more substantial influence on pollutant degradation.
Figure 6. Elucidation of the mechanisms involved in both radical and non-radical processes during phenol degradation, along with the progression of singlet oxygen evolution [266].
Figure 6. Elucidation of the mechanisms involved in both radical and non-radical processes during phenol degradation, along with the progression of singlet oxygen evolution [266].
Water 16 00875 g006

5.2.3. Surface Activation

One crucial method by which the non-radical pathway operates is via its surface charge, a crucial factor in the interaction between biochar and ROPs or oxidants. The surface charge of biochar is significantly influenced by the pH of the surrounding solution. If the solution’s pH is below the biochar’s pHpzc (zero charge point), the biochar’s surface becomes positively charged [267]. When the biochar’s surface charge is contrary to that of the ROPs or oxidizer, molecules are drawn to the biochar’s surface by electrostatic forces. This phenomenon increases the likelihood that the reactive oxygen species will come into contact with the molecules, and then the activation process will occur at the catalytic active site. Additionally, the scientific literature has highlighted the significance of surface charge deprotonation in facilitating the electron transfer of PFRs located on the biochar surface [226].
Surface activation is a non-free radical process mechanism and a feasible way to degrade refractory ROPs. The degradation process is achieved by forming a surface complex, which is established by a strong electrostatic binding between positively charged C and PS [249]. Wang et al. [134] highlighted the important role played by surface-bound active complexes in N-doped biochar/PDS systems. The observation of a substantial rise in current upon the introduction of PDS suggests a confirmation of its interaction with biochar to create a metastable reaction complex [141]. The catalytic system predominantly relies on surface-bound and reactive PS complexes, along with the generation of 1O2, as the primary non-radical mechanisms. In addition, the biochar surface can bind free radicals to form another metastable substance present on the catalyst surface, which will not be quenched by conventional free radical quenchers, such as p-benzoquinone. It can only be quenched by KI, a quencher specifically designed to quench surfacing bound free radicals [118,119]. Xing and colleagues [268] developed a N and S co-doped biochar for the activation of PS in the degradation of methyl orange (MO). The activation of PS primarily involves the oxidation of surface-bound free radicals, such as SO4• and •OH. The degradation efficiency of MO reached 99% under gentle conditions. However, upon the addition of KI as a quencher, the degradation efficiency of MO decreased to 64% within the first 30 min.

6. Conclusions

This research presents a broad viewpoint for industry professionals on the implementation of biochar in AOPs. The manufacturing of catalysts based on biochar and the most recent advancements in mechanistic studies within the realm of AOPs are thoroughly explained. The synthesis methods and mechanisms employed by biochar-based catalysts are summarized, with a discussion on their roles in adsorption, persulfate activation, and Fenton-like processes. The key findings are as follows:
(1)
The choice of raw materials for biochar, the process of preparation, and the modification of biochar all have a substantial impact on the structure and characteristics of the material. Pyrolysis stands out as the most frequently employed method in biochar production, and the elevated temperature during this process significantly affects the overall performance of the biochar. Metal-doped catalysts, especially mixed-metal-doped catalysts, have a variety of applications such as a high REDOX activity of persulfate activation, versatility, ferromagnetism, and photocatalytic and antibacterial properties. OCGs present on biochar (e.g., -OH and -COOH) and the metals loaded on them react with PS to form free radicals. Therefore, they are suitable for use in heterogeneous catalysts or mixed systems consisting of different AOPs. In catalyst doped with heteroatoms, the heteroatoms not only promote the transfer of electrons, but also generate new active sites. During the degradation of pollutants by AOPs, it is important to consider that the process primarily involves the use of free radicals (such as SO4•, •OH, and •O2) as well as non-free radical pathways (like 1O2). The selection of a specific pathway is heavily influenced by the active sites found on the surface of biochar;
(2)
The mechanism of biochar combined with advanced oxidation to degrade AOPs involves both radical and non-radical pathways. Active sites that produce 1O2 consist of OCGs and defects, with OCGs helping to enhance the electron transfer process. Electron transfer from PFRs can lead to the generation of •OH, SO4•, and •O2. Furthermore, •OH is associated with -OH and oxygen-centered PFRs, while •O2 is linked to quinone structures and carbon-centered PFRs. The adsorption mechanism of biochar for AOPs is driven by hydrogen bonds and π-π interactions. The graphitization degree, pore structure, and specific surface area of biochar impact both the electron transfer and adsorption capabilities of the catalyst, thus playing a crucial role in AOPs’ activation.
As the domain of catalysis progresses, there arises a need for materials that are not only inexpensive but that also exhibit excellent performance. This has led to the emergence of biochar-derived materials, such as activated carbon, which hold great potential as viable substitutes. Despite this progress, achieving practical applications still requires substantial efforts. For example, there is a requirement for extensive investigations into the development of oxygen functional groups (OFGs) on the biochar surface across diverse conditions. This will facilitate the customized conception of OFGs on biochar surfaces. Additional examination is crucial to unravel the characteristics and influential variables of catalysts based on biochar in various reaction systems. Investigating doping methods and functional groups, along with understanding how specific functional groups contribute to the overall catalytic activity, is essential. Additionally, unraveling the complex interactions between different components in real-world applications and how they cross-influence the reaction mechanisms of catalysts is crucial. Numerous obstacles and prospective paths were underscored, encompassing the necessity for thorough examinations concerning catalyst deactivation or toxicity evaluations amidst and subsequent to AOP treatments predicated on biochar. Comprehensive research on functional groups, mechanisms, and toxicity is required to draw specific conclusions. Currently, there is an insufficient number of comprehensive studies on the deactivation of catalysts or evaluations of toxicity throughout and following AOP treatments utilizing biochar. Future studies should focus on addressing these knowledge gaps to advance the understanding and application of BC-AOPs.

Author Contributions

F.K.: Conceptualization; data curation; formal analysis; investigation; methodology; software; validation; visualization; writing—original draft; writing—review and editing. J.L. (Jin Liu): data curation; formal analysis; investigation; methodology; software. W.F.: investigation; data curation. Z.X.: investigation; data curation. J.L. (Jiancong Liu): investigation; data curation. J.W.: funding acquisition and project administration. Y.W.: methodology; writing—review and editing. L.W.: writing—review and editing; formal analysis. B.X.: methodology; writing—review and editing; funding acquisition and project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Shenzhen Polytechnic Project (6023310038K, 6022312023K), Shenzhen Science and Technology Program (20231128105823001).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hu, Q.; Zhao, X.; Yang, X.J. China’s decadal pollution census. Nature 2017, 543, 491. [Google Scholar] [CrossRef] [PubMed]
  2. Nidheesh, P.V.; Gandhimathi, R.; Ramesh, S.T. Degradation of dyes from aqueous solution by Fenton processes: A review. Environ. Sci. Pollut. Res. 2013, 20, 2099–2132. [Google Scholar] [CrossRef] [PubMed]
  3. Glaze, W.H.; Kang, J.-W.; Chapin, D.H. The chemistry of water treatment processes involving ozone, hydrogen peroxide and ultraviolet radiation. Ozone Sci. Eng. 1987, 9, 335–352. [Google Scholar] [CrossRef]
  4. Faheem; Du, J.; Kim, S.H.; Hassan, M.A.; Irshad, S.; Bao, J. Application of biochar in advanced oxidation processes: Supportive, adsorptive, and catalytic role. Environ. Sci. Pollut. Res. 2020, 27, 37286–37312. [Google Scholar] [CrossRef]
  5. Romero, V.; Acevedo, S.; Marco, P.; Giménez, J.; Esplugas, S. Enhancement of Fenton and photo-Fenton processes at initial circumneutral pH for the degradation of the β-blocker metoprolol. Water Res. 2016, 88, 449–457. [Google Scholar] [CrossRef] [PubMed]
  6. Khan, J.A.; He, X.; Khan, H.M.; Shah, N.S.; Dionysiou, D.D. Oxidative degradation of atrazine in aqueous solution by UV/H2O2/Fe2+, UV/S2O82−/Fe2+ and UV/HSO5/Fe2+ processes: A comparative study. Chem. Eng. J. 2013, 218, 376–383. [Google Scholar] [CrossRef]
  7. Li, Y.; Pan, L.; Zhu, Y.; Yu, Y.; Wang, D.; Yang, G.; Yuan, X.; Liu, X.; Li, H.; Zhang, J. How does zero valent iron activating peroxydisulfate improve the dewatering of anaerobically digested sludge? Water Res. 2019, 163, 114912. [Google Scholar] [CrossRef]
  8. Yu, J.; Tang, L.; Pang, Y.; Zeng, G.; Wang, J.; Deng, Y.; Liu, Y.; Feng, H.; Chen, S.; Ren, X. Magnetic nitrogen-doped sludge-derived biochar catalysts for persulfate activation: Internal electron transfer mechanism. Chem. Eng. J. 2019, 364, 146–159. [Google Scholar] [CrossRef]
  9. Yu, X.; Lin, X.; Li, W.; Feng, W. Effective Removal of Tetracycline by Using Biochar Supported Fe3O4 as a UV-Fenton Catalyst. Chem. Res. Chin. Univ. 2019, 35, 79–84. [Google Scholar] [CrossRef]
  10. Kelly, J.; McDonnell, C.; Skillen, N.; Manesiotis, P.; Robertson, P.K. Enhanced photocatalytic degradation of 2-methyl-4-chlorophenoxyacetic acid (MCPA) by the addition of H2O2. Chemosphere 2021, 275, 130082. [Google Scholar] [CrossRef]
  11. Wang, K.; Liang, G.; Waqas, M.; Yang, B.; Xiao, K.; Zhu, C.; Zhang, J. Peroxymonosulfate enhanced photoelectrocatalytic degradation of ofloxacin using an easily coated cathode. Sep. Purif. Technol. 2020, 236, 116301. [Google Scholar] [CrossRef]
  12. Wu, S.; Hu, Y.H. A comprehensive review on catalysts for electrocatalytic and photoelectrocatalytic degradation of antibiotics. Chem. Eng. J. 2021, 409, 127739. [Google Scholar] [CrossRef]
  13. Farissi, S.; Abubakar, G.A.; Akhilghosh, K.A.; Muthukumar, A.; Muthuchamy, M. Sustainable application of electrocatalytic and photo-electrocatalytic oxidation systems for water and wastewater treatment: A review. Environ. Monit. Assess. 2023, 195, 1447. [Google Scholar] [CrossRef] [PubMed]
  14. Miklos, D.B.; Remy, C.; Jekel, M.; Linden, K.G.; Drewes, J.E.; Hübner, U. Evaluation of advanced oxidation processes for water and wastewater treatment–A critical review. Water Res. 2018, 139, 118–131. [Google Scholar] [CrossRef] [PubMed]
  15. Yanagida, H.; Masubuchi, Y.; Minagawa, K.; Ogata, T.; Takimoto, J.-i.; Koyama, K. A reaction kinetics model of water sonolysis in the presence of a spin-trap. Ultrason. Sonochemistry 1999, 5, 133–139. [Google Scholar] [CrossRef] [PubMed]
  16. Boffito, D.; Crocellà, V.; Pirola, C.; Neppolian, B.; Cerrato, G.; Ashokkumar, M.; Bianchi, C. Ultrasonic enhancement of the acidity, surface area and free fatty acids esterification catalytic activity of sulphated ZrO2–TiO2 systems. J. Catal. 2013, 297, 17–26. [Google Scholar] [CrossRef]
  17. Joseph, J.M.; Destaillats, H.; Hung, H.-M.; Hoffmann, M.R. The sonochemical degradation of azobenzene and related azo dyes: Rate enhancements via Fenton’s reactions. J. Phys. Chem. A 2000, 104, 301–307. [Google Scholar] [CrossRef]
  18. Shi, Y.; Li, J.; Wan, D.; Huang, J.; Liu, Y. Peroxymonosulfate-enhanced photocatalysis by carbonyl-modified g-C3N4 for effective degradation of the tetracycline hydrochloride. Sci. Total Environ. 2020, 749, 142313. [Google Scholar] [CrossRef]
  19. Lv, Y.; Liu, Y.; Wei, J.; Li, M.; Xu, D.; Lai, B. Bisphenol S degradation by visible light assisted peroxymonosulfate process based on BiOI/B4C photocatalysts with Z-scheme heterojunction. Chem. Eng. J. 2021, 417, 129188. [Google Scholar] [CrossRef]
  20. Zu, M.; Zhang, S.; Liu, C.; Liu, P.; Li, D.-S.; Xing, C.; Zhang, S. Portable wastewater treatment system based on synergistic photocatalytic and persulphate degradation under visible light. Sci. China Mater. 2021, 64, 1952–1963. [Google Scholar] [CrossRef]
  21. Wang, Q.; Cao, Y.; Zeng, H.; Liang, Y.; Ma, J.; Lu, X. Ultrasound-enhanced zero-valent copper activation of persulfate for the degradation of bisphenol AF. Chem. Eng. J. 2019, 378, 122143. [Google Scholar] [CrossRef]
  22. Fagan, W.P.; Zhao, J.; Villamena, F.A.; Zweier, J.L.; Weavers, L.K. Synergistic, aqueous PAH degradation by ultrasonically-activated persulfate depends on bulk temperature and physicochemical parameters. Ultrason. Sonochemistry 2020, 67, 105172. [Google Scholar] [CrossRef] [PubMed]
  23. Sizykh, M.; Batoeva, A. Effect of anions on the oxidation of organic compounds with ultrasonically activated persulfate. Russ. J. Phys. Chem. A 2016, 90, 1298–1300. [Google Scholar] [CrossRef]
  24. Aseman-Bashiz, E.; Sayyaf, H. Synthesis of nano-FeS2 and its application as an effective activator of ozone and peroxydisulfate in the electrochemical process for ofloxacin degradation: A comparative study. Chemosphere 2021, 274, 129772. [Google Scholar] [CrossRef]
  25. Santos, A.; Fernandez, J.; Rodriguez, S.; Dominguez, C.; Lominchar, M.; Lorenzo, D.; Romero, A. Abatement of chlorinated compounds in groundwater contaminated by HCH wastes using ISCO with alkali activated persulfate. Sci. Total Environ. 2018, 615, 1070–1077. [Google Scholar] [CrossRef] [PubMed]
  26. Huang, S.; Guo, X.; Duan, W.; Cheng, X.; Zhang, X.; Li, Z. Degradation of high molecular weight polyacrylamide by alkali-activated persulfate: Reactivity and potential application in filter cake removal before cementing. J. Pet. Sci. Eng. 2019, 174, 70–79. [Google Scholar] [CrossRef]
  27. Liu, H.; Tian, Z.; Huang, C.; Wang, P.; Huang, S.; Yang, X.; Cheng, H.; Zheng, X.; Tian, H.; Liu, Z. A novel 3D Co/Mo co-catalyzed graphene sponge-mediated peroxymonosulfate activation for the highly efficient pollutants degradation. Sep. Purif. Technol. 2022, 301, 122035. [Google Scholar] [CrossRef]
  28. Dai, L.; Wu, B.; Tan, F.; He, M.; Wang, W.; Qin, H.; Tang, X.; Zhu, Q.; Pan, K.; Hu, Q. Engineered hydrochar composites for phosphorus removal/recovery: Lanthanum doped hydrochar prepared by hydrothermal carbonization of lanthanum pretreated rice straw. Bioresour. Technol. 2014, 161, 327–332. [Google Scholar] [CrossRef]
  29. Dai, L.; Tan, F.; Wu, B.; He, M.; Wang, W.; Tang, X.; Hu, Q.; Zhang, M. Immobilization of phosphorus in cow manure during hydrothermal carbonization. J. Environ. Manag. 2015, 157, 49–53. [Google Scholar] [CrossRef]
  30. Dai, L.; Yang, B.; Li, H.; Tan, F.; Zhu, N.; Zhu, Q.; He, M.; Ran, Y.; Hu, G. A synergistic combination of nutrient reclamation from manure and resultant hydrochar upgradation by acid-supported hydrothermal carbonization. Bioresour. Technol. 2017, 243, 860–866. [Google Scholar] [CrossRef]
  31. de Souza, D.I.; Dottein, E.M.; Giacobbo, A.; Rodrigues, M.A.S.; de Pinho, M.N.; Bernardes, A.M. Nanofiltration for the removal of norfloxacin from pharmaceutical effluent. J. Environ. Chem. Eng. 2018, 6, 6147–6153. [Google Scholar] [CrossRef]
  32. Zhang, H.; Wang, T.; Chen, W.-Y.; Zhang, Y.; Sun, B.; Pan, W.-P. Derivation of oxygen-containing functional groups on biochar under non-oxygen plasma for mercury removal. Fuel 2020, 275, 117879. [Google Scholar] [CrossRef]
  33. Sun, H.; Yan, Z.; Liu, F.; Xu, W.; Cheng, F.; Chen, J. Self-supported transition-metal-based electrocatalysts for hydrogen and oxygen evolution. Adv. Mater. 2020, 32, 1806326. [Google Scholar] [CrossRef] [PubMed]
  34. Li, S.; Harris, S.; Anandhi, A.; Chen, G. Predicting biochar properties and functions based on feedstock and pyrolysis temperature: A review and data syntheses. J. Clean. Prod. 2019, 215, 890–902. [Google Scholar] [CrossRef]
  35. Tomczyk, A.; Sokołowska, Z.; Boguta, P. Biochar physicochemical properties: Pyrolysis temperature and feedstock kind effects. Rev. Environ. Sci. Bio/Technol. 2020, 19, 191–215. [Google Scholar] [CrossRef]
  36. Das, S.K.; Ghosh, G.K.; Avasthe, R.; Sinha, K. Compositional heterogeneity of different biochar: Effect of pyrolysis temperature and feedstocks. J. Environ. Manag. 2021, 278, 111501. [Google Scholar] [CrossRef] [PubMed]
  37. Wijitkosum, S. Biochar derived from agricultural wastes and wood residues for sustainable agricultural and environmental applications. Int. Soil Water Conserv. Res. 2022, 10, 335–341. [Google Scholar] [CrossRef]
  38. Sun, J.; Norouzi, O.; Mašek, O. A state-of-the-art review on algae pyrolysis for bioenergy and biochar production. Bioresour. Technol. 2022, 346, 126258. [Google Scholar] [CrossRef]
  39. Aravind, S.; Kumar, P.S.; Kumar, N.S.; Siddarth, N. Conversion of green algal biomass into bioenergy by pyrolysis. A review. Environ. Chem. Lett. 2020, 18, 829–849. [Google Scholar] [CrossRef]
  40. Wang, C.; Wang, Y.; Yan, S.; Li, Y.; Zhang, P.; Ren, P.; Wang, M.; Kuang, S. Biochar-amended composting of lincomycin fermentation dregs promoted microbial metabolism and reduced antibiotic resistance genes. Bioresour. Technol. 2023, 367, 128253. [Google Scholar] [CrossRef]
  41. Akdeniz, N. A systematic review of biochar use in animal waste composting. Waste Manag. 2019, 88, 291–300. [Google Scholar] [CrossRef] [PubMed]
  42. Akca, M.O.; OK, S.S.; Deniz, K.; Mohammedelnour, A.; Kibar, M. Spectroscopic characterisation and elemental composition of biochars obtained from different agricultural wastes. J. Agric. Sci. 2021, 27, 426–435. [Google Scholar] [CrossRef]
  43. Gopinath, A.; Divyapriya, G.; Srivastava, V.; Laiju, A.; Nidheesh, P.; Kumar, M.S. Conversion of sewage sludge into biochar: A potential resource in water and wastewater treatment. Environ. Res. 2021, 194, 110656. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, Y.; Wang, L.; Zhang, Y.; Mao, X.; Tan, W.; Zhang, Y.; Wang, X.; Chang, M.; Guo, R.; Xi, B. Perdisulfate-assisted advanced oxidation of 2, 4-dichlorophenol by bio-inspired iron encapsulated biochar catalyst. J. Colloid Interface Sci. 2021, 592, 358–370. [Google Scholar] [CrossRef] [PubMed]
  45. Curtis, L.J.; Miller, D.J. Transport model with radiative heat transfer for rapid cellulose pyrolysis. Ind. Eng. Chem. Res. 1988, 27, 1775–1783. [Google Scholar] [CrossRef]
  46. Anthony, D.B.; Howard, J.B. Coal devolatilization and hydrogastification. AIChE J. 1976, 22, 625–656. [Google Scholar] [CrossRef]
  47. Phuong, D.T.M.; Loc, N.X. Rice straw biochar and magnetic rice straw biochar for safranin O adsorption from aqueous solution. Water 2022, 14, 186. [Google Scholar] [CrossRef]
  48. Saravanan, P.; Vinod, V.; Sreedhar, B.; Sashidhar, R. Gum kondagogu modified magnetic nano-adsorbent: An efficient protocol for removal of various toxic metal ions. Mater. Sci. Eng. C 2012, 32, 581–586. [Google Scholar] [CrossRef]
  49. Asgharzadeh, F.; Kalantary, R.R.; Gholami, M.; Jafari, A.J.; Kermani, M.; Asgharnia, H. TiO2-decorated magnetic biochar mediated heterogeneous photocatalytic degradation of tetracycline and evaluation of antibacterial activity. Biomass Convers. Biorefinery 2021, 13, 8949–8959. [Google Scholar] [CrossRef]
  50. Steele, W.V.; Appelman, E.H. The standard enthalpy of formation of peroxymonosulfate (HSO5) and the standard electrode potential of the peroxymonosulfate-bisulfate couple. J. Chem. Thermodyn. 1982, 14, 337–344. [Google Scholar] [CrossRef]
  51. Rani, S.K.; Easwaramoorthy, D.; Bilal, I.M.; Palanichamy, M. Studies on Mn (II)-catalyzed oxidation of α-amino acids by peroxomonosulphate in alkaline medium-deamination and decarboxylation: A kinetic approach. Appl. Catal. A Gen. 2009, 369, 1–7. [Google Scholar] [CrossRef]
  52. Zhang, H.; Xue, G.; Chen, H.; Li, X. Magnetic biochar catalyst derived from biological sludge and ferric sludge using hydrothermal carbonization: Preparation, characterization and its circulation in Fenton process for dyeing wastewater treatment. Chemosphere 2018, 191, 64–71. [Google Scholar] [CrossRef] [PubMed]
  53. Jung, K.-W.; Lee, S.Y.; Lee, Y.J.; Choi, J.-W. Ultrasound-assisted heterogeneous Fenton-like process for bisphenol A removal at neutral pH using hierarchically structured manganese dioxide/biochar nanocomposites as catalysts. Ultrason. Sonochemistry 2019, 57, 22–28. [Google Scholar] [CrossRef]
  54. Ma, H.; Li, J.-B.; Liu, W.-W.; Miao, M.; Cheng, B.-J.; Zhu, S.-W. Novel synthesis of a versatile magnetic adsorbent derived from corncob for dye removal. Bioresour. Technol. 2015, 190, 13–20. [Google Scholar] [CrossRef] [PubMed]
  55. Gao, H.; Lv, S.; Dou, J.; Kong, M.; Dai, D.; Si, C.; Liu, G. The efficient adsorption removal of Cr (vi) by using Fe3O4 nanoparticles hybridized with carbonaceous materials. RSC Adv. 2015, 5, 60033–60040. [Google Scholar] [CrossRef]
  56. Khataee, A.; Gholami, P.; Kalderis, D.; Pachatouridou, E.; Konsolakis, M. Preparation of novel CeO2-biochar nanocomposite for sonocatalytic degradation of a textile dye. Ultrason. Sonochemistry 2018, 41, 503–513. [Google Scholar] [CrossRef] [PubMed]
  57. Abdul, G.; Zhu, X.; Chen, B. Structural characteristics of biochar-graphene nanosheet composites and their adsorption performance for phthalic acid esters. Chem. Eng. J. 2017, 319, 9–20. [Google Scholar] [CrossRef]
  58. Lu, L.; Shan, R.; Shi, Y.; Wang, S.; Yuan, H. A novel TiO2/biochar composite catalysts for photocatalytic degradation of methyl orange. Chemosphere 2019, 222, 391–398. [Google Scholar] [CrossRef]
  59. Li, C.; Zhang, L.; Gao, Y.; Li, A. Facile synthesis of nano ZnO/ZnS modified biochar by directly pyrolyzing of zinc contaminated corn stover for Pb (II), Cu (II) and Cr (VI) removals. Waste Manag. 2018, 79, 625–637. [Google Scholar] [CrossRef]
  60. Li, M.; Liu, H.; Chen, T.; Dong, C.; Sun, Y. Synthesis of magnetic biochar composites for enhanced uranium (VI) adsorption. Sci. Total Environ. 2019, 651, 1020–1028. [Google Scholar] [CrossRef]
  61. Fang, C.; Zhang, T.; Li, P.; Jiang, R.; Wu, S.; Nie, H.; Wang, Y. Phosphorus recovery from biogas fermentation liquid by Ca–Mg loaded biochar. J. Environ. Sci. 2015, 29, 106–114. [Google Scholar] [CrossRef] [PubMed]
  62. Baig, S.A.; Zhu, J.; Muhammad, N.; Sheng, T.; Xu, X. Effect of synthesis methods on magnetic Kans grass biochar for enhanced As (III, V) adsorption from aqueous solutions. Biomass Bioenergy 2014, 71, 299–310. [Google Scholar] [CrossRef]
  63. Yu, J.-X.; Wang, L.-Y.; Chi, R.-A.; Zhang, Y.-F.; Xu, Z.-G.; Guo, J. Competitive adsorption of Pb2+ and Cd2+ on magnetic modified sugarcane bagasse prepared by two simple steps. Appl. Surf. Sci. 2013, 268, 163–170. [Google Scholar] [CrossRef]
  64. Wang, M.; Sheng, G.; Qiu, Y. A novel manganese-oxide/biochar composite for efficient removal of lead (II) from aqueous solutions. Int. J. Environ. Sci. Technol. 2015, 12, 1719–1726. [Google Scholar] [CrossRef]
  65. Mohan, D.; Kumar, H.; Sarswat, A.; Alexandre-Franco, M.; Pittman Jr, C.U. Cadmium and lead remediation using magnetic oak wood and oak bark fast pyrolysis bio-chars. Chem. Eng. J. 2014, 236, 513–528. [Google Scholar] [CrossRef]
  66. Tang, Z.; Zhao, S.; Qian, Y.; Jia, H.; Gao, P.; Kang, Y.; Lichtfouse, E. Formation of persistent free radicals in sludge biochar by hydrothermal carbonization. Environ. Chem. Lett. 2021, 19, 2705–2712. [Google Scholar] [CrossRef]
  67. Madduri, S.; Elsayed, I. Novel oxone treated hydrochar for the removal of Pb (II) and methylene blue (MB) dye from aqueous solutions. Chemosphere 2020, 260, 127683. [Google Scholar] [CrossRef] [PubMed]
  68. Kohzadi, S.; Marzban, N.; Libra, J.A.; Bundschuh, M.; Maleki, A. Removal of RhB from water by Fe-modified hydrochar and biochar–An experimental evaluation supported by genetic programming. J. Mol. Liq. 2023, 369, 120971. [Google Scholar] [CrossRef]
  69. Petrović, J.; Ercegović, M.; Simić, M.; Kalderis, D.; Koprivica, M.; Milojković, J.; Radulović, D. Novel Mg-doped pyro-hydrochars as methylene blue adsorbents: Adsorption behavior and mechanism. J. Mol. Liq. 2023, 376, 121424. [Google Scholar] [CrossRef]
  70. Jais, F.M.; Ibrahim, S.; Chee, C.Y.; Ismail, Z. Solvothermal growth of the bimetal organic framework (NiFe-MOF) on sugarcane bagasse hydrochar for the removal of dye and antibiotic. J. Environ. Chem. Eng. 2021, 9, 106367. [Google Scholar] [CrossRef]
  71. Lee, H.-S.; Shin, H.-S. Competitive adsorption of heavy metals onto modified biochars: Comparison of biochar properties and modification methods. J. Environ. Manag. 2021, 299, 113651. [Google Scholar] [CrossRef] [PubMed]
  72. Li, J.-H.; Lv, G.-H.; Bai, W.-B.; Liu, Q.; Zhang, Y.-C.; Song, J.-Q. Modification and use of biochar from wheat straw (Triticum aestivum L.) for nitrate and phosphate removal from water . Desalination Water Treat. 2016, 57, 4681–4693. [Google Scholar]
  73. Jin, H.; Capareda, S.; Chang, Z.; Gao, J.; Xu, Y.; Zhang, J. Biochar pyrolytically produced from municipal solid wastes for aqueous As (V) removal: Adsorption property and its improvement with KOH activation. Bioresour. Technol. 2014, 169, 622–629. [Google Scholar] [CrossRef]
  74. Peng, Z.; Zhao, H.; Lyu, H.; Wang, L.; Huang, H.; Nan, Q.; Tang, J. UV modification of biochar for enhanced hexavalent chromium removal from aqueous solution. Environ. Sci. Pollut. Res. 2018, 25, 10808–10819. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, Y.; Chen, Z.; Chen, C.; Li, F.; Shen, K. Effects of UV-modified biochar derived from phytoremediation residue on Cd bioavailability and uptake in Coriandrum sativum L. in a Cd-contaminated soil. Environ. Sci. Pollut. Res. 2021, 28, 17395–17404. [Google Scholar] [CrossRef]
  76. Yuvaraj, A.; Thangaraj, R.; Karmegam, N.; Ravindran, B.; Chang, S.W.; Awasthi, M.K.; Kannan, S. Activation of biochar through exoenzymes prompted by earthworms for vermibiochar production: A viable resource recovery option for heavy metal contaminated soils and water. Chemosphere 2021, 278, 130458. [Google Scholar] [CrossRef] [PubMed]
  77. Liu, W.; Wang, J.; Cai, N.; Wang, J.; Shen, L.; Shi, H.; Yang, D.; Feng, X.; Yu, F. Porous carbon nanofibers loaded with copper-cobalt bimetallic particles for heterogeneously catalyzing peroxymonosulfate to degrade organic dyes. J. Environ. Chem. Eng. 2021, 9, 106003. [Google Scholar] [CrossRef]
  78. Wang, C.; Sun, R.; Huang, R.; Wang, H. Superior fenton-like degradation of tetracycline by iron loaded graphitic carbon derived from microplastics: Synthesis, catalytic performance, and mechanism. Sep. Purif. Technol. 2021, 270, 118773. [Google Scholar] [CrossRef]
  79. Wang, J.; Wang, S. Preparation, modification and environmental application of biochar: A review. J. Clean. Prod. 2019, 227, 1002–1022. [Google Scholar] [CrossRef]
  80. Wang, H.; Guo, W.; Liu, B.; Si, Q.; Luo, H.; Zhao, Q.; Ren, N. Sludge-derived biochar as efficient persulfate activators: Sulfurization-induced electronic structure modulation and disparate nonradical mechanisms. Appl. Catal. B Environ. 2020, 279, 119361. [Google Scholar] [CrossRef]
  81. Tan, X.-F.; Liu, Y.-G.; Gu, Y.-L.; Xu, Y.; Zeng, G.-M.; Hu, X.-J.; Liu, S.-B.; Wang, X.; Liu, S.-M.; Li, J. Biochar-based nano-composites for the decontamination of wastewater: A review. Bioresour. Technol. 2016, 212, 318–333. [Google Scholar] [CrossRef] [PubMed]
  82. Anipsitakis, G.P.; Stathatos, E.; Dionysiou, D.D. Heterogeneous activation of oxone using Co3O4. J. Phys. Chem. B 2005, 109, 13052–13055. [Google Scholar] [CrossRef] [PubMed]
  83. Chen, X.; Chen, J.; Qiao, X.; Wang, D.; Cai, X. Performance of nano-Co3O4/peroxymonosulfate system: Kinetics and mechanism study using Acid Orange 7 as a model compound. Appl. Catal. B: Environ. 2008, 80, 116–121. [Google Scholar] [CrossRef]
  84. Guo, W.; Su, S.; Yi, C.; Ma, Z. Degradation of antibiotics amoxicillin by Co3O4-catalyzed peroxymonosulfate system. Environ. Prog. Sustain. Energy 2013, 32, 193–197. [Google Scholar] [CrossRef]
  85. Shi, P.; Su, R.; Zhu, S.; Zhu, M.; Li, D.; Xu, S. Supported cobalt oxide on graphene oxide: Highly efficient catalysts for the removal of Orange II from water. J. Hazard. Mater. 2012, 229, 331–339. [Google Scholar] [CrossRef]
  86. Lin, K.-Y.A.; Hsu, F.-K.; Lee, W.-D. Magnetic cobalt–graphene nanocomposite derived from self-assembly of MOFs with graphene oxide as an activator for peroxymonosulfate. J. Mater. Chem. A 2015, 3, 9480–9490. [Google Scholar]
  87. Yang, Q.; Choi, H.; Al-Abed, S.R.; Dionysiou, D.D. Iron–cobalt mixed oxide nanocatalysts: Heterogeneous peroxymonosulfate activation, cobalt leaching, and ferromagnetic properties for environmental applications. Appl. Catal. B Environ. 2009, 88, 462–469. [Google Scholar] [CrossRef]
  88. Muhammad, S.; Saputra, E.; Sun, H.J.d.; Izidoro, C.; Fungaro, D.A.; Ang, H.M.; Tadé, M.O.; Wang, S. Coal fly ash supported Co3O4 catalysts for phenol degradation using peroxymonosulfate. Rsc. Adv. 2012, 2, 5645–5650. [Google Scholar] [CrossRef]
  89. Wang, Y.; Sun, H.; Ang, H.M.; Tadé, M.O.; Wang, S. Magnetic Fe3O4/carbon sphere/cobalt composites for catalytic oxidation of phenol solutions with sulfate radicals. Chem. Eng. J. 2014, 245, 1–9. [Google Scholar] [CrossRef]
  90. Shukla, P.; Sun, H.; Wang, S.; Ang, H.M.; Tadé, M.O. Co-SBA-15 for heterogeneous oxidation of phenol with sulfate radical for wastewater treatment. Catal. Today 2011, 175, 380–385. [Google Scholar] [CrossRef]
  91. Shukla, P.; Wang, S.; Singh, K.; Ang, H.; Tadé, M.O. Cobalt exchanged zeolites for heterogeneous catalytic oxidation of phenol in the presence of peroxymonosulphate. Appl. Catal. B Environ. 2010, 99, 163–169. [Google Scholar] [CrossRef]
  92. Liu, B.; Guo, W.; Wang, H.; Si, Q.; Zhao, Q.; Luo, H.; Ren, N. Activation of peroxymonosulfate by cobalt-impregnated biochar for atrazine degradation: The pivotal roles of persistent free radicals and ecotoxicity assessment. J. Hazard. Mater. 2020, 398, 122768. [Google Scholar] [CrossRef]
  93. Chen, L.; Yang, S.; Zuo, X.; Huang, Y.; Cai, T.; Ding, D. Biochar modification significantly promotes the activity of Co3O4 towards heterogeneous activation of peroxymonosulfate. Chem. Eng. J. 2018, 354, 856–865. [Google Scholar] [CrossRef]
  94. Xu, H.; Zhang, Y.; Li, J.; Hao, Q.; Li, X.; Liu, F. Heterogeneous activation of peroxymonosulfate by a biochar-supported Co3O4 composite for efficient degradation of chloramphenicols. Environ. Pollut. 2020, 257, 113610. [Google Scholar] [CrossRef]
  95. Oh, W.-D.; Lua, S.-K.; Dong, Z.; Lim, T.-T. Rational design of hierarchically-structured CuBi2O4 composites by deliberate manipulation of the nucleation and growth kinetics of CuBi2O4 for environmental applications. Nanoscale 2016, 8, 2046–2054. [Google Scholar] [CrossRef] [PubMed]
  96. Feng, Y.; Liu, J.; Wu, D.; Zhou, Z.; Deng, Y.; Zhang, T.; Shih, K. Efficient degradation of sulfamethazine with CuCo2O4 spinel nanocatalysts for peroxymonosulfate activation. Chem. Eng. J. 2015, 280, 514–524. [Google Scholar] [CrossRef]
  97. Ji, F.; Li, C.; Deng, L. Performance of CuO/Oxone system: Heterogeneous catalytic oxidation of phenol at ambient conditions. Chem. Eng. J. 2011, 178, 239–243. [Google Scholar] [CrossRef]
  98. Kušić, H.; Koprivanac, N.; Selanec, I. Fe-exchanged zeolite as the effective heterogeneous Fenton-type catalyst for the organic pollutant minimization: UV irradiation assistance. Chemosphere 2006, 65, 65–73. [Google Scholar] [CrossRef]
  99. Liu, J.; Jiang, G.; Liu, Y.; Di, J.; Wang, Y.; Zhao, Z.; Sun, Q.; Xu, C.; Gao, J.; Duan, A. Hierarchical macro-meso-microporous ZSM-5 zeolite hollow fibers with highly efficient catalytic cracking capability. Sci. Rep. 2014, 4, 7276. [Google Scholar] [CrossRef]
  100. Zhang, T.; Chen, Y.; Wang, Y.; Le Roux, J.; Yang, Y.; Croué, J.-P. Efficient peroxydisulfate activation process not relying on sulfate radical generation for water pollutant degradation. Environ. Sci. Technol. 2014, 48, 5868–5875. [Google Scholar] [CrossRef]
  101. Li, Z.; Liu, D.; Huang, W.; Wei, X.; Huang, W. Biochar supported CuO composites used as an efficient peroxymonosulfate activator for highly saline organic wastewater treatment. Sci. Total Environ. 2020, 721, 137764. [Google Scholar] [CrossRef] [PubMed]
  102. Luo, H.; Lin, Q.; Zhang, X.; Huang, Z.; Fu, H.; Xiao, R.; Liu, S.-S. Determining the key factors of nonradical pathway in activation of persulfate by metal-biochar nanocomposites for bisphenol A degradation. Chem. Eng. J. 2020, 391, 123555. [Google Scholar] [CrossRef]
  103. Wang, C.; Huang, R.; Sun, R. Green one-spot synthesis of hydrochar supported zero-valent iron for heterogeneous Fenton-like discoloration of dyes at neutral pH. J. Mol. Liq. 2020, 320, 114421. [Google Scholar] [CrossRef]
  104. Hussain, I.; Li, M.; Zhang, Y.; Li, Y.; Huang, S.; Du, X.; Liu, G.; Hayat, W.; Anwar, N. Insights into the mechanism of persulfate activation with nZVI/BC nanocomposite for the degradation of nonylphenol. Chem. Eng. J. 2017, 311, 163–172. [Google Scholar] [CrossRef]
  105. Liu, C.-M.; Diao, Z.-H.; Huo, W.-Y.; Kong, L.-J.; Du, J.-J. Simultaneous removal of Cu2+ and bisphenol A by a novel biochar-supported zero valent iron from aqueous solution: Synthesis, reactivity and mechanism. Environ. Pollut. 2018, 239, 698–705. [Google Scholar] [CrossRef] [PubMed]
  106. Li, H.; Zhu, F.; He, S. The degradation of decabromodiphenyl ether in the e-waste site by biochar supported nanoscale zero-valent iron/persulfate. Ecotoxicol. Environ. Saf. 2019, 183, 109540. [Google Scholar] [CrossRef] [PubMed]
  107. Jiang, S.-F.; Ling, L.-L.; Chen, W.-J.; Liu, W.-J.; Li, D.-C.; Jiang, H. High efficient removal of bisphenol A in a peroxymonosulfate/iron functionalized biochar system: Mechanistic elucidation and quantification of the contributors. Chem. Eng. J. 2019, 359, 572–583. [Google Scholar] [CrossRef]
  108. Luo, H.; Lin, Q.; Zhang, X.; Huang, Z.; Liu, S.; Jiang, J.; Xiao, R.; Liao, X. New insights into the formation and transformation of active species in nZVI/BC activated persulfate in alkaline solutions. Chem. Eng. J. 2019, 359, 1215–1223. [Google Scholar] [CrossRef]
  109. Volpe, A.; Pagano, M.; Mascolo, G.; Lopez, A.; Ciannarella, R.; Locaputo, V. Simultaneous Cr (VI) reduction and non-ionic surfactant oxidation by peroxymonosulphate and iron powder. Chemosphere 2013, 91, 1250–1256. [Google Scholar] [CrossRef]
  110. Ghanbari, F.; Moradi, M.; Manshouri, M. Textile wastewater decolorization by zero valent iron activated peroxymonosulfate: Compared with zero valent copper. J. Environ. Chem. Eng. 2014, 2, 1846–1851. [Google Scholar] [CrossRef]
  111. Pi, Z.; Li, X.; Wang, D.; Xu, Q.; Tao, Z.; Huang, X.; Yao, F.; Wu, Y.; He, L.; Yang, Q. Persulfate activation by oxidation biochar supported magnetite particles for tetracycline removal: Performance and degradation pathway. J. Clean. Prod. 2019, 235, 1103–1115. [Google Scholar] [CrossRef]
  112. Rong, X.; Xie, M.; Kong, L.; Natarajan, V.; Ma, L.; Zhan, J. The magnetic biochar derived from banana peels as a persulfate activator for organic contaminants degradation. Chem. Eng. J. 2019, 372, 294–303. [Google Scholar] [CrossRef]
  113. Li, Y.; Ma, S.; Xu, S.; Fu, H.; Li, Z.; Li, K.; Sheng, K.; Du, J.; Lu, X.; Li, X. Novel magnetic biochar as an activator for peroxymonosulfate to degrade bisphenol A: Emphasizing the synergistic effect between graphitized structure and CoFe2O4. Chem. Eng. J. 2020, 387, 124094. [Google Scholar] [CrossRef]
  114. Dong, C.-D.; Chen, C.-W.; Hung, C.-M. Synthesis of magnetic biochar from bamboo biomass to activate persulfate for the removal of polycyclic aromatic hydrocarbons in marine sediments. Bioresour. Technol. 2017, 245, 188–195. [Google Scholar] [CrossRef] [PubMed]
  115. Ouyang, D.; Yan, J.; Qian, L.; Chen, Y.; Han, L.; Su, A.; Zhang, W.; Ni, H.; Chen, M. Degradation of 1, 4-dioxane by biochar supported nano magnetite particles activating persulfate. Chemosphere 2017, 184, 609–617. [Google Scholar] [CrossRef]
  116. Li, J.; Pan, L.; Yu, G.; Xie, S.; Li, C.; Lai, D.; Li, Z.; You, F.; Wang, Y. The synthesis of heterogeneous Fenton-like catalyst using sewage sludge biochar and its application for ciprofloxacin degradation. Sci. Total Environ. 2019, 654, 1284–1292. [Google Scholar] [CrossRef] [PubMed]
  117. Gan, Q.; Hou, H.; Liang, S.; Qiu, J.; Tao, S.; Yang, L.; Yu, W.; Xiao, K.; Liu, B.; Hu, J. Sludge-derived biochar with multivalent iron as an efficient Fenton catalyst for degradation of 4-Chlorophenol. Sci. Total Environ. 2020, 725, 138299. [Google Scholar] [CrossRef] [PubMed]
  118. Chen, L.; Jiang, X.; Xie, R.; Zhang, Y.; Jin, Y.; Jiang, W. A novel porous biochar-supported Fe-Mn composite as a persulfate activator for the removal of acid red 88. Sep. Purif. Technol. 2020, 250, 117232. [Google Scholar] [CrossRef]
  119. 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]
  120. Liu, C.; Chen, L.; Ding, D.; Cai, T. From rice straw to magnetically recoverable nitrogen doped biochar: Efficient activation of peroxymonosulfate for the degradation of metolachlor. Appl. Catal. B Environ. 2019, 254, 312–320. [Google Scholar] [CrossRef]
  121. 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]
  122. Hao, H.; Zhang, Q.; Qiu, Y.; Meng, L.; Wei, X.; Sang, W.; Tao, J. Insight into the degradation of Orange G by persulfate activated with biochar modified by iron and manganese oxides: Synergism between Fe and Mn. J. Water Process Eng. 2020, 37, 101470. [Google Scholar] [CrossRef]
  123. Huang, D.; Zhang, Q.; Zhang, C.; Wang, R.; Deng, R.; Luo, H.; Li, T.; Li, J.; Chen, S.; Liu, C. Mn doped magnetic biochar as persulfate activator for the degradation of tetracycline. Chem. Eng. J. 2020, 391, 123532. [Google Scholar] [CrossRef]
  124. Liu, L.; Li, Y.; Li, W.; Zhong, R.; Lan, Y.; Guo, J. The efficient degradation of sulfisoxazole by singlet oxygen (1O2) derived from activated peroxymonosulfate (PMS) with Co3O4–SnO2/RSBC. Environ. Res. 2020, 187, 109665. [Google Scholar] [CrossRef] [PubMed]
  125. Zhu, S.; Wang, Z.; Ye, C.; Deng, J.; Ma, X.; Xu, Y.; Wang, L.; Tang, Z.; Luo, H.; Li, X. Magnetic Co/Fe nanocomposites derived from ferric sludge as an efficient peroxymonosulfate catalyst for ciprofloxacin degradation. Chem. Eng. J. 2022, 432, 134180. [Google Scholar] [CrossRef]
  126. 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]
  127. Zeng, L.; Li, W.; Cheng, J.; Wang, J.; Liu, X.; Yu, Y. N-doped porous hollow carbon nanofibers fabricated using electrospun polymer templates and their sodium storage properties. Rsc Adv. 2014, 4, 16920–16927. [Google Scholar] [CrossRef]
  128. Chen, Z.; Li, H. The lithium ions storage behavior of heteroatom-mediated echinus-like porous carbon spheres: From co-doping to multi-atom doping. J. Colloid Interface Sci. 2020, 567, 54–64. [Google Scholar] [CrossRef]
  129. Niu, W.; Li, L.; Liu, X.; Wang, N.; Liu, J.; Zhou, W.; Tang, Z.; Chen, S. Mesoporous N-doped carbons prepared with thermally removable nanoparticle templates: An efficient electrocatalyst for oxygen reduction reaction. J. Am. Chem. Soc. 2015, 137, 5555–5562. [Google Scholar] [CrossRef]
  130. Ding, Y.; Li, Y.; Dai, Y.; Han, X.; Xing, B.; Zhu, L.; Qiu, K.; Wang, S. A novel approach for preparing in-situ nitrogen doped carbon via pyrolysis of bean pulp for supercapacitors. Energy 2021, 216, 119227. [Google Scholar] [CrossRef]
  131. Yu, J.; Li, X.; Cui, Z.; Chen, D.; Pang, X.; Zhang, Q.; Shao, F.; Dong, H.; Yu, L.; Dong, L. Tailoring in-situ N, O, P, S-doped soybean-derived porous carbon with ultrahigh capacitance in both acidic and alkaline media. Renew. Energy 2021, 163, 375–385. [Google Scholar] [CrossRef]
  132. Liu, S.; Yang, Z.; Li, M.; Liu, L.; Wang, Y.; Lv, W.; Qin, Z.; Zhao, X.; Zhu, P.; Wang, G. FeS-decorated hierarchical porous N, S-dual-doped carbon derived from silica-ionogel as an efficient catalyst for oxygen reduction reaction in alkaline media. Electrochim. Acta 2018, 265, 221–231. [Google Scholar] [CrossRef]
  133. Liu, Y.; Xiao, Z.; Liu, Y.; Fan, L.-Z. Biowaste-derived 3D honeycomb-like porous carbon with binary-heteroatom doping for high-performance flexible solid-state supercapacitors. J. Mater. Chem. A 2018, 6, 160–166. [Google Scholar] [CrossRef]
  134. Wang, H.; Guo, W.; Liu, B.; Wu, Q.; Luo, H.; Zhao, Q.; Si, Q.; Sseguya, F.; Ren, N. Edge-nitrogenated biochar for efficient peroxydisulfate activation: An electron transfer mechanism. Water Res. 2019, 160, 405–414. [Google Scholar] [CrossRef] [PubMed]
  135. Oh, W.-D.; Lisak, G.; Webster, R.D.; Liang, Y.-N.; Veksha, A.; Giannis, A.; Moo, J.G.S.; Lim, J.-W.; Lim, T.-T. Insights into the thermolytic transformation of lignocellulosic biomass waste to redox-active carbocatalyst: Durability of surface active sites. Appl. Catal. B Environ. 2018, 233, 120–129. [Google Scholar] [CrossRef]
  136. Xu, L.; Wu, C.; Liu, P.; Bai, X.; Du, X.; Jin, P.; Yang, L.; Jin, X.; Shi, X.; Wang, Y. Peroxymonosulfate activation by nitrogen-doped biochar from sawdust for the efficient degradation of organic pollutants. Chem. Eng. J. 2020, 387, 124065. [Google Scholar] [CrossRef]
  137. Terrell, E.; Garcia-Perez, M. Application of nitrogen-based blowing agents as an additive in pyrolysis of cellulose. J. Anal. Appl. Pyrolysis 2019, 137, 203–211. [Google Scholar] [CrossRef]
  138. Ding, D.; Yang, S.; Qian, X.; Chen, L.; Cai, T. Nitrogen-doping positively whilst sulfur-doping negatively affect the catalytic activity of biochar for the degradation of organic contaminant. Appl. Catal. B: Environ. 2020, 263, 118348. [Google Scholar] [CrossRef]
  139. Li, B.; Li, K. Effect of nitric acid pre-oxidation concentration on pore structure and nitrogen/oxygen active decoration sites of ethylenediamine-modified biochar for mercury (II) adsorption and the possible mechanism. Chemosphere 2019, 220, 28–39. [Google Scholar] [CrossRef]
  140. Lee, K.S.; Park, C.W.; Phiri, I.; Ko, J.M. New design for Polyaniline@ Multiwalled carbon nanotubes composites with bacteria doping for supercapacitor electrodes. Polymer 2020, 210, 123014. [Google Scholar] [CrossRef]
  141. Zhu, S.; Huang, X.; Ma, F.; Wang, L.; Duan, X.; Wang, S. Catalytic removal of aqueous contaminants on N-doped graphitic biochars: Inherent roles of adsorption and nonradical mechanisms. Environ. Sci. Technol. 2018, 52, 8649–8658. [Google Scholar] [CrossRef] [PubMed]
  142. Wang, L.; Yan, W.; He, C.; Wen, H.; Cai, Z.; Wang, Z.; Chen, Z.; Liu, W. Microwave-assisted preparation of nitrogen-doped biochars by ammonium acetate activation for adsorption of acid red 18. Appl. Surf. Sci. 2018, 433, 222–231. [Google Scholar] [CrossRef]
  143. Zhou, Q.; Jiang, X.; Li, X.; Jia, C.Q.; Jiang, W. Preparation of high-yield N-doped biochar from nitrogen-containing phosphate and its effective adsorption for toluene. RSC Adv. 2018, 8, 30171–30179. [Google Scholar] [CrossRef] [PubMed]
  144. Ye, S.; Zeng, G.; Tan, X.; Wu, H.; Liang, J.; Song, B.; Tang, N.; Zhang, P.; Yang, Y.; Chen, Q. Nitrogen-doped biochar fiber with graphitization from Boehmeria nivea for promoted peroxymonosulfate activation and non-radical degradation pathways with enhancing electron transfer. Appl. Catal. B Environ. 2020, 269, 118850. [Google Scholar] [CrossRef]
  145. Xie, Y.; Hu, W.; Wang, X.; Tong, W.; Li, P.; Zhou, H.; Wang, Y.; Zhang, Y. Molten salt induced nitrogen-doped biochar nanosheets as highly efficient peroxymonosulfate catalyst for organic pollutant degradation. Environ. Pollut. 2020, 260, 114053. [Google Scholar] [CrossRef] [PubMed]
  146. Guo, D.; Shibuya, R.; Akiba, C.; Saji, S.; Kondo, T.; Nakamura, J. Active sites of nitrogen-doped carbon materials for oxygen reduction reaction clarified using model catalysts. Science 2016, 351, 361–365. [Google Scholar] [CrossRef] [PubMed]
  147. Ho, S.-H.; Li, R.; Zhang, C.; Ge, Y.; Cao, G.; Ma, M.; Duan, X.; Wang, S.; Ren, N.-q. N-doped graphitic biochars from C-phycocyanin extracted Spirulina residue for catalytic persulfate activation toward nonradical disinfection and organic oxidation. Water Res. 2019, 159, 77–86. [Google Scholar] [CrossRef]
  148. Yan, P.; Liu, J.; Yuan, S.; Liu, Y.; Cen, W.; Chen, Y. The promotion effects of graphitic and pyridinic N combinational doping on graphene for ORR. Appl. Surf. Sci. 2018, 445, 398–403. [Google Scholar] [CrossRef]
  149. Gong, K.; Du, F.; Xia, Z.; Durstock, M.; Dai, L. Nitrogen-doped carbon nanotube arrays with high electrocatalytic activity for oxygen reduction. science 2009, 323, 760–764. [Google Scholar] [CrossRef]
  150. Ravidhas, C.; Anitha, B.; Moses Ezhil Raj, A.; Ravichandran, K.; Sabari Girisun, T.; Mahalakshmi, K.; Saravanakumar, K.; Sanjeeviraja, C. Effect of nitrogen doped titanium dioxide (N-TiO2) thin films by jet nebulizer spray technique suitable for photoconductive study. J. Mater. Sci. Mater. Electron. 2015, 26, 3573–3582. [Google Scholar] [CrossRef]
  151. Patel, N.; Jaiswal, R.; Warang, T.; Scarduelli, G.; Dashora, A.; Ahuja, B.; Kothari, D.; Miotello, A. Efficient photocatalytic degradation of organic water pollutants using V–N-codoped TiO2 thin films. Appl. Catal. B Environ. 2014, 150, 74–81. [Google Scholar] [CrossRef]
  152. Asahi, R.; Morikawa, T. Nitrogen complex species and its chemical nature in TiO2 for visible-light sensitized photocatalysis. Chem. Phys. 2007, 339, 57–63. [Google Scholar] [CrossRef]
  153. Tavares, C.; Marques, S.; Viseu, T.; Teixeira, V.; Carneiro, J.; Alves, E.; Barradas, N.; Munnik, F.; Girardeau, T.; Rivière, J.-P. Enhancement in the photocatalytic nature of nitrogen-doped PVD-grown titanium dioxide thin films. J. Appl. Phys. 2009, 106, 113535. [Google Scholar] [CrossRef]
  154. Li, C.; Gu, M.; Gao, M.; Liu, K.; Zhao, X.; Cao, N.; Feng, J.; Ren, Y.; Wei, T.; Zhang, M. N-doping TiO2 hollow microspheres with abundant oxygen vacancies for highly photocatalytic nitrogen fixation. J. Colloid Interface Sci. 2022, 609, 341–352. [Google Scholar] [CrossRef] [PubMed]
  155. Wan, H.; Ju, X.; He, T.; Chen, T.; Zhou, Y.; Zhang, C.; Wang, J.; Xu, Y.; Yao, B.; Zhuang, W. Sulfur-doped porous carbon as high-capacity anodes for lithium and sodium ions batteries. J. Alloys Compd. 2021, 863, 158078. [Google Scholar] [CrossRef]
  156. Yaglikci, S.; Gokce, Y.; Yagmur, E.; Aktas, Z. The performance of sulphur doped activated carbon supercapacitors prepared from waste tea. Environ. Technol. 2019, 41, 36–48. [Google Scholar] [CrossRef]
  157. Paraknowitsch, J.P.; Thomas, A. Doping carbons beyond nitrogen: An overview of advanced heteroatom doped carbons with boron, sulphur and phosphorus for energy applications. Energy Environ. Sci. 2013, 6, 2839–2855. [Google Scholar] [CrossRef]
  158. 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]
  159. Wang, S.; Wang, J. Peroxymonosulfate activation by Co9S8@ S and N co-doped biochar for sulfamethoxazole degradation. Chem. Eng. J. 2020, 385, 123933. [Google Scholar] [CrossRef]
  160. Li, X.; Jia, Y.; Zhou, M.; Su, X.; Sun, J. High-efficiency degradation of organic pollutants with Fe, N co-doped biochar catalysts via persulfate activation. J. Hazard. Mater. 2020, 397, 122764. [Google Scholar] [CrossRef]
  161. Wang, W.; Lu, C.; Ni, Y.; Su, M.; Xu, Z. A new sight on hydrogenation of F and NF doped {0 0 1} facets dominated anatase TiO2 for efficient visible light photocatalyst. Appl. Catal. B Environ. 2012, 127, 28–35. [Google Scholar] [CrossRef]
  162. Zhong, Q.; Lin, Q.; Huang, R.; Fu, H.; Zhang, X.; Luo, H.; Xiao, R. Oxidative degradation of tetracycline using persulfate activated by N and Cu codoped biochar. Chem. Eng. J. 2020, 380, 122608. [Google Scholar] [CrossRef]
  163. Ren, T.-L.; Ma, X.-W.; Wu, X.-Q.; Yuan, L.; Lai, Y.-L.; Tong, Z.-H. Degradation of imidazolium ionic liquids in a thermally activated persulfate system. Chem. Eng. J. 2021, 412, 128624. [Google Scholar] [CrossRef]
  164. Mani, P.; Kim, Y.; Lakhera, S.K.; Neppolian, B.; Choi, H. Complete arsenite removal from groundwater by UV activated potassium persulfate and iron oxide impregnated granular activated carbon. Chemosphere 2021, 277, 130225. [Google Scholar] [CrossRef] [PubMed]
  165. Dominguez, C.M.; Rodriguez, V.; Montero, E.; Romero, A.; Santos, A. Abatement of dichloromethane using persulfate activated by alkali: A kinetic study. Sep. Purif. Technol. 2020, 241, 116679. [Google Scholar] [CrossRef]
  166. 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]
  167. Pan, X.; Gu, Z.; Chen, W.; Li, Q. Preparation of biochar and biochar composites and their application in a Fenton-like process for wastewater decontamination: A review. Sci. Total Environ. 2021, 754, 142104. [Google Scholar] [CrossRef] [PubMed]
  168. Sun, C.; Chen, T.; Huang, Q.; Zhan, M.; Li, X.; Yan, J. Activation of persulfate by CO2-activated biochar for improved phenolic pollutant degradation: Performance and mechanism. Chem. Eng. J. 2020, 380, 122519. [Google Scholar] [CrossRef]
  169. Wan, Z.; Sun, Y.; Tsang, D.C.; Iris, K.; Fan, J.; Clark, J.H.; Zhou, Y.; Cao, X.; Gao, B.; Ok, Y.S. A sustainable biochar catalyst synergized with copper heteroatoms and CO2 for singlet oxygenation and electron transfer routes. Green Chem. 2019, 21, 4800–4814. [Google Scholar] [CrossRef]
  170. Gupta, V.; Moradi, O.; Tyagi, I.; Agarwal, S.; Sadegh, H.; Shahryari-Ghoshekandi, R.; Makhlouf, A.; Goodarzi, M.; Garshasbi, A. Study on the removal of heavy metal ions from industry waste by carbon nanotubes: Effect of the surface modification: A review. Crit. Rev. Environ. Sci. Technol. 2016, 46, 93–118. [Google Scholar] [CrossRef]
  171. Tan, X.-F.; Zhu, S.-S.; Wang, R.-P.; Chen, Y.-D.; Show, P.-L.; Zhang, F.-F.; Ho, S.-H. Role of biochar surface characteristics in the adsorption of aromatic compounds: Pore structure and functional groups. Chin. Chem. Lett. 2021, 32, 2939–2946. [Google Scholar] [CrossRef]
  172. Tan, X.; Liu, Y.; Zeng, G.; Wang, X.; Hu, X.; Gu, Y.; Yang, Z. Application of biochar for the removal of pollutants from aqueous solutions. Chemosphere 2015, 125, 70–85. [Google Scholar] [CrossRef]
  173. Xu, R.-K.; Xiao, S.-C.; Yuan, J.-H.; Zhao, A.-Z. Adsorption of methyl violet from aqueous solutions by the biochars derived from crop residues. Bioresour. Technol. 2011, 102, 10293–10298. [Google Scholar] [CrossRef] [PubMed]
  174. Chen, Z.; Xiao, X.; Chen, B.; Zhu, L. Quantification of chemical states, dissociation constants and contents of oxygen-containing groups on the surface of biochars produced at different temperatures. Environ. Sci. Technol. 2015, 49, 309–317. [Google Scholar] [CrossRef] [PubMed]
  175. Diao, Z.-H.; Zhang, W.-X.; Liang, J.-Y.; Huang, S.-T.; Dong, F.-X.; Yan, L.; Qian, W.; Chu, W. Removal of herbicide atrazine by a novel biochar based iron composite coupling with peroxymonosulfate process from soil: Synergistic effect and mechanism. Chem. Eng. J. 2021, 409, 127684. [Google Scholar] [CrossRef]
  176. Xu, J.; Zhang, X.; Sun, C.; Wan, J.; He, H.; Wang, F.; Dai, Y.; Yang, S.; Lin, Y.; Zhan, X. Insights into removal of tetracycline by persulfate activation with peanut shell biochar coupled with amorphous Cu-doped FeOOH composite in aqueous solution. Environ. Sci. Pollut. Res. 2019, 26, 2820–2834. [Google Scholar] [CrossRef] [PubMed]
  177. Yan, J.; Han, L.; Gao, W.; Xue, S.; Chen, M. Biochar supported nanoscale zerovalent iron composite used as persulfate activator for removing trichloroethylene. Bioresour. Technol. 2015, 175, 269–274. [Google Scholar] [CrossRef] [PubMed]
  178. Huong, P.T.; Jitae, K.; Al Tahtamouni, T.; Tri, N.L.M.; Kim, H.-H.; Cho, K.H.; Lee, C. Novel activation of peroxymonosulfate by biochar derived from rice husk toward oxidation of organic contaminants in wastewater. J. Water Process Eng. 2020, 33, 101037. [Google Scholar] [CrossRef]
  179. Meng, H.; Nie, C.; Li, W.; Duan, X.; Lai, B.; Ao, Z.; Wang, S.; An, T. Insight into the effect of lignocellulosic biomass source on the performance of biochar as persulfate activator for aqueous organic pollutants remediation: Epicarp and mesocarp of citrus peels as examples. J. Hazard. Mater. 2020, 399, 123043. [Google Scholar] [CrossRef] [PubMed]
  180. Klupfel, L.; Keiluweit, M.; Kleber, M.; Sander, M. Redox properties of plant biomass-derived black carbon (biochar). Environ. Sci. Technol. 2014, 48, 5601–5611. [Google Scholar] [CrossRef]
  181. Chen, N.; Huang, Y.; Hou, X.; Ai, Z.; Zhang, L. Photochemistry of hydrochar: Reactive oxygen species generation and sulfadimidine degradation. Environ. Sci. Technol. 2017, 51, 11278–11287. [Google Scholar] [CrossRef] [PubMed]
  182. Rivera-Utrilla, J.; Sánchez-Polo, M.; Gómez-Serrano, V.; Álvarez, P.; Alvim-Ferraz, M.; Dias, J. Activated carbon modifications to enhance its water treatment applications. An overview. J. Hazard. Mater. 2011, 187, 1–23. [Google Scholar] [CrossRef] [PubMed]
  183. Chen, W.; Yang, H.; Chen, Y.; Xia, M.; Chen, X.; Chen, H. Transformation of nitrogen and evolution of N-containing species during algae pyrolysis. Environ. Sci. Technol. 2017, 51, 6570–6579. [Google Scholar] [CrossRef]
  184. Ding, W.; Wei, Z.; Chen, S.; Qi, X.; Yang, T.; Hu, J.; Wang, D.; Wan, L.J.; Alvi, S.F.; Li, L. Space-confinement-induced synthesis of pyridinic-and pyrrolic-nitrogen-doped graphene for the catalysis of oxygen reduction. Angew. Chem. 2013, 125, 11971–11975. [Google Scholar] [CrossRef]
  185. Saha, D.; Kienbaum, M.J. Role of oxygen, nitrogen and sulfur functionalities on the surface of nanoporous carbons in CO2 adsorption: A critical review. Microporous Mesoporous Mater. 2019, 287, 29–55. [Google Scholar] [CrossRef]
  186. Xu, X.; Zheng, Y.; Gao, B.; Cao, X. N-doped biochar synthesized by a facile ball-milling method for enhanced sorption of CO2 and reactive red. Chem. Eng. J. 2019, 368, 564–572. [Google Scholar] [CrossRef]
  187. Xu, J.; Zhang, X.; Sun, C.; He, H.; Dai, Y.; Yang, S.; Lin, Y.; Zhan, X.; Li, Q.; Zhou, Y. Catalytic degradation of diatrizoate by persulfate activation with peanut shell biochar-supported nano zero-valent iron in aqueous solution. Int. J. Environ. Res. Public Health 2018, 15, 1937. [Google Scholar] [CrossRef]
  188. Pan, L.; Cao, Y.; Zang, J.; Huang, Q.; Wang, L.; Zhang, Y.; Fan, S.; Tang, J.; Xie, Z. Preparation of iron-loaded granular activated carbon catalyst and its application in tetracycline antibiotic removal from aqueous solution. Int. J. Environ. Res. Public Health 2019, 16, 2270. [Google Scholar] [CrossRef]
  189. Zhang, H.; Wang, T.; Sui, Z.; Zhang, Y.; Sun, B.; Pan, W.-P. Enhanced mercury removal by transplanting sulfur-containing functional groups to biochar through plasma. Fuel 2019, 253, 703–712. [Google Scholar] [CrossRef]
  190. Huang, S.; Liang, Q.; Geng, J.; Luo, H.; Wei, Q. Sulfurized biochar prepared by simplified technic with superior adsorption property towards aqueous Hg (II) and adsorption mechanisms. Mater. Chem. Phys. 2019, 238, 121919. [Google Scholar] [CrossRef]
  191. Hsi, H.-C.; Tsai, C.-Y.; Kuo, T.-H.; Chiang, C.-S. Development of low-concentration mercury adsorbents from biohydrogen-generation agricultural residues using sulfur impregnation. Bioresour. Technol. 2011, 102, 7470–7477. [Google Scholar] [CrossRef] [PubMed]
  192. Park, J.-H.; Wang, J.J.; Zhou, B.; Mikhael, J.E.; DeLaune, R.D. Removing mercury from aqueous solution using sulfurized biochar and associated mechanisms. Environ. Pollut. 2019, 244, 627–635. [Google Scholar] [CrossRef] [PubMed]
  193. Jeon, C.; Solis, K.L.; An, H.-R.; Hong, Y.; Igalavithana, A.D.; Ok, Y.S. Sustainable removal of Hg (II) by sulfur-modified pine-needle biochar. J. Hazard. Mater. 2020, 388, 122048. [Google Scholar] [CrossRef] [PubMed]
  194. Liu, P.; Ptacek, C.J.; Elena, K.M.; Blowes, D.W.; Gould, W.D.; Finfrock, Y.Z.; Wang, A.O.; Landis, R.C. Evaluation of mercury stabilization mechanisms by sulfurized biochars determined using X-ray absorption spectroscopy. J. Hazard. Mater. 2018, 347, 114–122. [Google Scholar] [CrossRef] [PubMed]
  195. Krishnan, K.A.; Anirudhan, T. Removal of mercury (II) from aqueous solutions and chlor-alkali industry effluent by steam activated and sulphurised activated carbons prepared from bagasse pith: Kinetics and equilibrium studies. J. Hazard. Mater. 2002, 92, 161–183. [Google Scholar] [CrossRef] [PubMed]
  196. Yang, X.; Wan, Y.; Zheng, Y.; He, F.; Yu, Z.; Huang, J.; Wang, H.; Ok, Y.S.; Jiang, Y.; Gao, B. Surface functional groups of carbon-based adsorbents and their roles in the removal of heavy metals from aqueous solutions: A critical review. Chem. Eng. J. 2019, 366, 608–621. [Google Scholar] [CrossRef] [PubMed]
  197. O’Connor, D.; Peng, T.; Li, G.; Wang, S.; Duan, L.; Mulder, J.; Cornelissen, G.; Cheng, Z.; Yang, S.; Hou, D. Sulfur-modified rice husk biochar: A green method for the remediation of mercury contaminated soil. Sci. Total Environ. 2018, 621, 819–826. [Google Scholar] [CrossRef]
  198. Lefebvre, D.D.; Kelly, D.; Budd, K. Biotransformation of Hg (II) by cyanobacteria. Appl. Environ. Microbiol. 2007, 73, 243–249. [Google Scholar] [CrossRef]
  199. Shen, Y.; Chen, B. Sulfonated graphene nanosheets as a superb adsorbent for various environmental pollutants in water. Environ. Sci. Technol. 2015, 49, 7364–7372. [Google Scholar] [CrossRef]
  200. Sebastian, A.; Prasad, M.N.V. Cadmium minimization in rice. A review. Agron. Sustain. Dev. 2014, 34, 155–173. [Google Scholar] [CrossRef]
  201. Kubier, A.; Wilkin, R.T.; Pichler, T. Cadmium in soils and groundwater: A review. Appl. Geochem. 2019, 108, 104388. [Google Scholar] [CrossRef] [PubMed]
  202. Chen, D.; Wang, X.; Wang, X.; Feng, K.; Su, J.; Dong, J. The mechanism of cadmium sorption by sulphur-modified wheat straw biochar and its application cadmium-contaminated soil. Sci. Total Environ. 2020, 714, 136550. [Google Scholar] [CrossRef] [PubMed]
  203. Rajendran, M.; Shi, L.; Wu, C.; Li, W.; An, W.; Liu, Z.; Xue, S. Effect of sulfur and sulfur-iron modified biochar on cadmium availability and transfer in the soil–rice system. Chemosphere 2019, 222, 314–322. [Google Scholar] [CrossRef] [PubMed]
  204. Higashikawa, F.S.; Conz, R.F.; Colzato, M.; Cerri, C.E.P.; Alleoni, L.R.F. Effects of feedstock type and slow pyrolysis temperature in the production of biochars on the removal of cadmium and nickel from water. J. Clean. Prod. 2016, 137, 965–972. [Google Scholar] [CrossRef]
  205. Pearson, R.G. Hard and soft acids and bases. J. Am. Chem. Soc. 1963, 85, 3533–3539. [Google Scholar] [CrossRef]
  206. Huang, M.; Wang, X.; Liu, C.; Fang, G.; Gao, J.; Wang, Y.; Zhou, D. Facile ball milling preparation of sulfur-doped carbon as peroxymonosulfate activator for efficient removal of organic pollutants. J. Environ. Chem. Eng. 2021, 9, 106536. [Google Scholar] [CrossRef]
  207. Kiciński, W.; Szala, M.; Bystrzejewski, M. Sulfur-doped porous carbons: Synthesis and applications. Carbon 2014, 68, 1–32. [Google Scholar] [CrossRef]
  208. Jin, Z.; Li, Y.; Dong, H.; Xiao, S.; Xiao, J.; Chu, D.; Hou, X.; Xiang, S.; Dong, Q.; Li, L. A comparative study on the activation of persulfate by mackinawite@ biochar and pyrite@ biochar for sulfamethazine degradation: The role of different natural iron-sulfur minerals doping. Chem. Eng. J. 2022, 448, 137620. [Google Scholar] [CrossRef]
  209. Maliutina, K.; Tahmasebi, A.; Yu, J. Pressurized entrained-flow pyrolysis of microalgae: Enhanced production of hydrogen and nitrogen-containing compounds. Bioresour. Technol. 2018, 256, 160–169. [Google Scholar] [CrossRef]
  210. Chen, W.; Chen, Y.; Yang, H.; Xia, M.; Li, K.; Chen, X.; Chen, H. Co-pyrolysis of lignocellulosic biomass and microalgae: Products characteristics and interaction effect. Bioresour. Technol. 2017, 245, 860–868. [Google Scholar] [CrossRef]
  211. Chen, W.; Chen, Y.; Yang, H.; Li, K.; Chen, X.; Chen, H. Investigation on biomass nitrogen-enriched pyrolysis: Influence of temperature. Bioresour. Technol. 2018, 249, 247–253. [Google Scholar] [CrossRef] [PubMed]
  212. Liu, L.; Tan, Z.; Ye, Z. Transformation and transport mechanism of nitrogenous compounds in a biochar “preparation–returning to the field” process studied by employing an isotope tracer method. ACS Sustain. Chem. Eng. 2018, 6, 1780–1791. [Google Scholar] [CrossRef]
  213. Yu, J.; Maliutina, K.; Tahmasebi, A. A review on the production of nitrogen-containing compounds from microalgal biomass via pyrolysis. Bioresour. Technol. 2018, 270, 689–701. [Google Scholar] [CrossRef] [PubMed]
  214. Lian, F.; Cui, G.; Liu, Z.; Duo, L.; Zhang, G.; Xing, B. One-step synthesis of a novel N-doped microporous biochar derived from crop straws with high dye adsorption capacity. J. Environ. Manag. 2016, 176, 61–68. [Google Scholar] [CrossRef]
  215. Yuan, S.; Tan, Z.; Huang, Q. Migration and transformation mechanism of nitrogen in the biomass–biochar–plant transport process. Renew. Sustain. Energy Rev. 2018, 85, 1–13. [Google Scholar] [CrossRef]
  216. Wang, T.; Arbestain, M.C.; Hedley, M.; Bishop, P. Chemical and bioassay characterisation of nitrogen availability in biochar produced from dairy manure and biosolids. Org. Geochem. 2012, 51, 45–54. [Google Scholar] [CrossRef]
  217. Lomnicki, S.; Truong, H.; Vejerano, E.; Dellinger, B. Copper oxide-based model of persistent free radical formation on combustion-derived particulate matter. Environ. Sci. Technol. 2008, 42, 4982–4988. [Google Scholar] [CrossRef] [PubMed]
  218. Vejerano, E.; Lomnicki, S.M.; Dellinger, B. Formation and stabilization of combustion-generated, environmentally persistent radicals on Ni (II) O supported on a silica surface. Environ. Sci. Technol. 2012, 46, 9406–9411. [Google Scholar] [CrossRef] [PubMed]
  219. Huang, D.; Luo, H.; Zhang, C.; Zeng, G.; Lai, C.; Cheng, M.; Wang, R.; Deng, R.; Xue, W.; Gong, X. Nonnegligible role of biomass types and its compositions on the formation of persistent free radicals in biochar: Insight into the influences on Fenton-like process. Chem. Eng. J. 2019, 361, 353–363. [Google Scholar] [CrossRef]
  220. Deng, R.; Luo, H.; Huang, D.; Zhang, C. Biochar-mediated Fenton-like reaction for the degradation of sulfamethazine: Role of environmentally persistent free radicals. Chemosphere 2020, 255, 126975. [Google Scholar] [CrossRef]
  221. Jia, H.; Zhao, S.; Nulaji, G.; Tao, K.; Wang, F.; Sharma, V.K.; Wang, C. Environmentally persistent free radicals in soils of past coking sites: Distribution and stabilization. Environ. Sci. Technol. 2017, 51, 6000–6008. [Google Scholar] [CrossRef] [PubMed]
  222. Qin, Y.; Li, G.; Gao, Y.; Zhang, L.; Ok, Y.S.; An, T. Persistent free radicals in carbon-based materials on transformation of refractory organic contaminants (ROCs) in water: A critical review. Water Res. 2018, 137, 130–143. [Google Scholar] [CrossRef] [PubMed]
  223. Ruan, X.; Sun, Y.; Du, W.; Tang, Y.; Liu, Q.; Zhang, Z.; Doherty, W.; Frost, R.L.; Qian, G.; Tsang, D.C. Formation, characteristics, and applications of environmentally persistent free radicals in biochars: A review. Bioresour. Technol. 2019, 281, 457–468. [Google Scholar] [CrossRef] [PubMed]
  224. Fang, G.; Liu, C.; Gao, J.; Dionysiou, D.D.; Zhou, D. Manipulation of persistent free radicals in biochar to activate persulfate for contaminant degradation. Environ. Sci. Technol. 2015, 49, 5645–5653. [Google Scholar] [CrossRef]
  225. Fang, G.; Gao, J.; Liu, C.; Dionysiou, D.D.; Wang, Y.; Zhou, D. Key role of persistent free radicals in hydrogen peroxide activation by biochar: Implications to organic contaminant degradation. Environ. Sci. Technol. 2014, 48, 1902–1910. [Google Scholar] [CrossRef] [PubMed]
  226. Yu, J.; Zhu, Z.; Zhang, H.; Chen, T.; Qiu, Y.; Xu, Z.; Yin, D. Efficient removal of several estrogens in water by Fe-hydrochar composite and related interactive effect mechanism of H2O2 and iron with persistent free radicals from hydrochar of pinewood. Sci. Total Environ. 2019, 658, 1013–1022. [Google Scholar] [CrossRef] [PubMed]
  227. Wang, H.; Guo, W.; Yin, R.; Du, J.; Wu, Q.; Luo, H.; Liu, B.; Sseguya, F.; Ren, N. Biochar-induced Fe (III) reduction for persulfate activation in sulfamethoxazole degradation: Insight into the electron transfer, radical oxidation and degradation pathways. Chem. Eng. J. 2019, 362, 561–569. [Google Scholar] [CrossRef]
  228. He, W.; Zhu, Y.; Zeng, G.; Zhang, Y.; Wang, Y.; Zhang, M.; Long, H.; Tang, W. Efficient removal of perfluorooctanoic acid by persulfate advanced oxidative degradation: Inherent roles of iron-porphyrin and persistent free radicals. Chem. Eng. J. 2020, 392, 123640. [Google Scholar] [CrossRef]
  229. Wang, Y.; Gu, X.; Huang, Y.; Ding, Z.; Chen, Y.; Hu, X. Insight into biomass feedstock on formation of biochar-bound environmentally persistent free radicals under different pyrolysis temperatures. RSC Adv. 2022, 12, 19318–19326. [Google Scholar] [CrossRef]
  230. Zhong, D.; Zhang, Y.; Wang, L.; Chen, J.; Jiang, Y.; Tsang, D.C.; Zhao, Z.; Ren, S.; Liu, Z.; Crittenden, J.C. Mechanistic insights into adsorption and reduction of hexavalent chromium from water using magnetic biochar composite: Key roles of Fe3O4 and persistent free radicals. Environ. Pollut. 2018, 243, 1302–1309. [Google Scholar] [CrossRef]
  231. Wang, R.-Z.; Huang, D.-L.; Liu, Y.-G.; Zhang, C.; Lai, C.; Wang, X.; Zeng, G.-M.; Gong, X.-M.; Duan, A.; Zhang, Q. Recent advances in biochar-based catalysts: Properties, applications and mechanisms for pollution remediation. Chem. Eng. J. 2019, 371, 380–403. [Google Scholar] [CrossRef]
  232. Zhu, K.; Wang, X.; Geng, M.; Chen, D.; Lin, H.; Zhang, H. Catalytic oxidation of clofibric acid by peroxydisulfate activated with wood-based biochar: Effect of biochar pyrolysis temperature, performance and mechanism. Chem. Eng. J. 2019, 374, 1253–1263. [Google Scholar] [CrossRef]
  233. Tang, L.; Yu, J.; Pang, Y.; Zeng, G.; Deng, Y.; Wang, J.; Ren, X.; Ye, S.; Peng, B.; Feng, H. Sustainable efficient adsorbent: Alkali-acid modified magnetic biochar derived from sewage sludge for aqueous organic contaminant removal. Chem. Eng. J. 2018, 336, 160–169. [Google Scholar] [CrossRef]
  234. Chen, W.; Duan, L.; Wang, L.; Zhu, D. Adsorption of hydroxyl-and amino-substituted aromatics to carbon nanotubes. Environ. Sci. Technol. 2008, 42, 6862–6868. [Google Scholar] [CrossRef]
  235. Tang, J.; Lv, H.; Gong, Y.; Huang, Y. Preparation and characterization of a novel graphene/biochar composite for aqueous phenanthrene and mercury removal. Bioresour. Technol. 2015, 196, 355–363. [Google Scholar] [CrossRef] [PubMed]
  236. Zhou, Y.; He, Y.; He, Y.; Liu, X.; Xu, B.; Yu, J.; Dai, C.; Huang, A.; Pang, Y.; Luo, L. Analyses of tetracycline adsorption on alkali-acid modified magnetic biochar: Site energy distribution consideration. Sci. Total Environ. 2019, 650, 2260–2266. [Google Scholar] [CrossRef] [PubMed]
  237. Chen, J.; Bai, X.; Yuan, Y.; Zhang, Y.; Sun, J. Printing and dyeing sludge derived biochar for activation of peroxymonosulfate to remove aqueous organic pollutants: Activation mechanisms and environmental safety assessment. Chem. Eng. J. 2022, 446, 136942. [Google Scholar] [CrossRef]
  238. Xu, L.; Fu, B.; Sun, Y.; Jin, P.; Bai, X.; Jin, X.; Shi, X.; Wang, Y.; Nie, S. Degradation of organic pollutants by Fe/N co-doped biochar via peroxymonosulfate activation: Synthesis, performance, mechanism and its potential for practical application. Chem. Eng. J. 2020, 400, 125870. [Google Scholar] [CrossRef]
  239. Wei, J.; Liu, Y.; Li, J.; Zhu, Y.; Yu, H.; Peng, Y. Adsorption and co-adsorption of tetracycline and doxycycline by one-step synthesized iron loaded sludge biochar. Chemosphere 2019, 236, 124254. [Google Scholar] [CrossRef]
  240. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Johir, M.A.H.; Sornalingam, K. Single and competitive sorption properties and mechanism of functionalized biochar for removing sulfonamide antibiotics from water. Chem. Eng. J. 2017, 311, 348–358. [Google Scholar] [CrossRef]
  241. 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]
  242. Zhou, Y.; Liu, X.; Xiang, Y.; Wang, P.; Zhang, J.; Zhang, F.; Wei, J.; Luo, L.; Lei, M.; Tang, L. Modification of biochar derived from sawdust and its application in removal of tetracycline and copper from aqueous solution: Adsorption mechanism and modelling. Bioresour. Technol. 2017, 245, 266–273. [Google Scholar] [CrossRef] [PubMed]
  243. Li, T.; He, Y.; Peng, X. Efficient removal of tetrabromobisphenol A (TBBPA) using sewage sludge-derived biochar: Adsorptive effect and mechanism. Chemosphere 2020, 251, 126370. [Google Scholar] [CrossRef] [PubMed]
  244. Duan, X.; Sun, H.; Ao, Z.; Zhou, L.; Wang, G.; Wang, S. Unveiling the active sites of graphene-catalyzed peroxymonosulfate activation. Carbon 2016, 107, 371–378. [Google Scholar] [CrossRef]
  245. Geng, A.; Xu, L.; Gan, L.; Mei, C.; Wang, L.; Fang, X.; Li, M.; Pan, M.; Han, S.; Cui, J. Using wood flour waste to produce biochar as the support to enhance the visible-light photocatalytic performance of BiOBr for organic and inorganic contaminants removal. Chemosphere 2020, 250, 126291. [Google Scholar] [CrossRef] [PubMed]
  246. Yu, J.; Tang, L.; Pang, Y.; Zeng, G.; Feng, H.; Zou, J.; Wang, J.; Feng, C.; Zhu, X.; Ouyang, X. 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]
  247. Ouyang, D.; Chen, Y.; Yan, J.; Qian, L.; Han, L.; Chen, M. Activation mechanism of peroxymonosulfate by biochar for catalytic degradation of 1, 4-dioxane: Important role of biochar defect structures. Chem. Eng. J. 2019, 370, 614–624. [Google Scholar] [CrossRef]
  248. Kohantorabi, M.; Moussavi, G.; Giannakis, S. A review of the innovations in metal-and carbon-based catalysts explored for heterogeneous peroxymonosulfate (PMS) activation, with focus on radical vs. non-radical degradation pathways of organic contaminants. Chem. Eng. J. 2021, 411, 127957. [Google Scholar] [CrossRef]
  249. 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]
  250. Duan, X.; Sun, H.; Tade, M.; Wang, S. Metal-free activation of persulfate by cubic mesoporous carbons for catalytic oxidation via radical and nonradical processes. Catal. Today 2018, 307, 140–146. [Google Scholar] [CrossRef]
  251. Zhang, H.; Tang, L.; Wang, J.; Yu, J.; Feng, H.; Lu, Y.; Chen, Y.; Liu, Y.; Wang, J.; Xie, Q. Enhanced surface activation process of persulfate by modified bagasse biochar for degradation of phenol in water and soil: Active sites and electron transfer mechanism. Colloids Surf. A Physicochem. Eng. Asp. 2020, 599, 124904. [Google Scholar] [CrossRef]
  252. 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] [PubMed]
  253. Nosaka, Y.; Nosaka, A.Y. Generation and detection of reactive oxygen species in photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef] [PubMed]
  254. Wang, J.; Shen, M.; Gong, Q.; Wang, X.; Cai, J.; Wang, S.; Chen, Z. One-step preparation of ZVI-sludge derived biochar without external source of iron and its application on persulfate activation. Sci. Total Environ. 2020, 714, 136728. [Google Scholar] [CrossRef] [PubMed]
  255. Park, Y.; Kim, C.; Kim, M.; Kim, S.; Choi, W. Ambient-temperature catalytic degradation of aromatic compounds on iron oxide nanorods supported on carbon nanofiber sheet. Appl. Catal. B Environ. 2019, 259, 118066. [Google Scholar] [CrossRef]
  256. Yi, Q.; Ji, J.; Shen, B.; Dong, C.; Liu, J.; Zhang, J.; Xing, M. Singlet oxygen triggered by superoxide radicals in a molybdenum cocatalytic Fenton reaction with enhanced REDOX activity in the environment. Environ. Sci. Technol. 2019, 53, 9725–9733. [Google Scholar] [CrossRef]
  257. Zhang, P.; Tan, X.; Liu, S.; Liu, Y.; Zeng, G.; Ye, S.; Yin, Z.; Hu, X.; Liu, N. Catalytic degradation of estrogen by persulfate activated with iron-doped graphitic biochar: Process variables effects and matrix effects. Chem. Eng. J. 2019, 378, 122141. [Google Scholar] [CrossRef]
  258. Zhang, W.; Li, Y.; Fan, X.; Zhang, F.; Zhang, G.; Zhu, Y.-A.; Peng, W.; Wang, S.; Duan, X. Synergy of nitrogen doping and structural defects on hierarchically porous carbons toward catalytic oxidation via a non-radical pathway. Carbon 2019, 155, 268–278. [Google Scholar] [CrossRef]
  259. Cheng, X.; Guo, H.; Zhang, Y.; Wu, X.; Liu, Y. Non-photochemical production of singlet oxygen via activation of persulfate by carbon nanotubes. Water Res. 2017, 113, 80–88. [Google Scholar] [CrossRef]
  260. 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]
  261. Guo, Y.; Yan, L.; Li, X.; Yan, T.; Song, W.; Hou, T.; Tong, C.; Mu, J.; Xu, M. Goethite/biochar-activated peroxymonosulfate enhances tetracycline degradation: Inherent roles of radical and non-radical processes. Sci. Total Environ. 2021, 783, 147102. [Google Scholar] [CrossRef]
  262. Pang, K.; Sun, W.; Ye, F.; Yang, L.; Pu, M.; Yang, C.; Zhang, Q.; Niu, J. Sulfur-modified chitosan derived N, S-co-doped carbon as a bifunctional material for adsorption and catalytic degradation sulfamethoxazole by persulfate. J. Hazard. Mater. 2022, 424, 127270. [Google Scholar] [CrossRef]
  263. Gao, S.; Wang, Z.; Wang, H.; Jia, Y.; Xu, N.; Wang, X.; Wang, J.; Zhang, C.; Tian, T.; Shen, W. Peroxydisulfate activation using B-doped biochar for the degradation of oxytetracycline in water. Appl. Surf. Sci. 2022, 599, 153917. [Google Scholar] [CrossRef]
  264. Li, X.; Jia, Y.; Zhang, J.; Qin, Y.; Wu, Y.; Zhou, M.; Sun, J. Efficient removal of tetracycline by H2O2 activated with iron-doped biochar: Performance, mechanism, and degradation pathways. Chin. Chem. Lett. 2022, 33, 2105–2110. [Google Scholar] [CrossRef]
  265. Yin, Q.; Liu, M.; Ren, H. Biochar produced from the co-pyrolysis of sewage sludge and walnut shell for ammonium and phosphate adsorption from water. J. Environ. Manag. 2019, 249, 109410. [Google Scholar] [CrossRef] [PubMed]
  266. Duan, X.; Sun, H.; Wang, S. Metal-free carbocatalysis in advanced oxidation reactions. Acc. Chem. Res. 2018, 51, 678–687. [Google Scholar] [CrossRef] [PubMed]
  267. Kemmou, L.; Frontistis, Z.; Vakros, J.; Manariotis, I.D.; Mantzavinos, D. Degradation of antibiotic sulfamethoxazole by biochar-activated persulfate: Factors affecting the activation and degradation processes. Catal. Today 2018, 313, 128–133. [Google Scholar] [CrossRef]
  268. Xing, B.; Dong, J.; Yang, G.; Jiang, N.; Liu, X.; Yuan, J. An insight into N, S-codoped activated carbon for the catalytic persulfate oxidation of organic pollutions in water: Effect of surface functionalization. Appl. Catal. A Gen. 2020, 602, 117714. [Google Scholar] [CrossRef]
Figure 1. Preparation of biochar-based nanocomposites [81].
Figure 1. Preparation of biochar-based nanocomposites [81].
Water 16 00875 g001
Figure 2. (a) The synthesis of CuO/BC, Fe3O4/BC, and ZnO/BC leads to the creation of lactones. (b) Understanding the activation mechanism of PS in metal/biochar nanocomposites [102].
Figure 2. (a) The synthesis of CuO/BC, Fe3O4/BC, and ZnO/BC leads to the creation of lactones. (b) Understanding the activation mechanism of PS in metal/biochar nanocomposites [102].
Water 16 00875 g002
Figure 3. Schematic representation delineating the synthesis of N-doped biochar for the degradation of tetracycline [144].
Figure 3. Schematic representation delineating the synthesis of N-doped biochar for the degradation of tetracycline [144].
Water 16 00875 g003
Figure 4. (a) The degradation process of sulfadimethoxine in Fenton-like systems and the significance of oxygen-centered PFRs [220]. (b) Investigating the degradation mechanism of 17 β-estradiol in a Fenton-like system employing hydrothermal carbon immobilized by iron (hydrogen) oxides as a catalyst and exploring the function of carbon-centered PFRs [226].
Figure 4. (a) The degradation process of sulfadimethoxine in Fenton-like systems and the significance of oxygen-centered PFRs [220]. (b) Investigating the degradation mechanism of 17 β-estradiol in a Fenton-like system employing hydrothermal carbon immobilized by iron (hydrogen) oxides as a catalyst and exploring the function of carbon-centered PFRs [226].
Water 16 00875 g004
Table 2. Application of biochar-based catalysts prepared by different methods in water treatment.
Table 2. Application of biochar-based catalysts prepared by different methods in water treatment.
MethodBiomassCatalystCharacteristics and AdvantagesRemoval PropertyReference
PyrolysisWheat strawCeO2Higher BET, average pore size and pore volume.The removal rate was high (95.8%), and the removal rate was still 87% after 5 cycles of experiment. [56]
WoodGraphene oxideHigher BET, porous structure and thermal stability.More effcient removal of organic pollutants through π-π EDA interaction and the maximum adsorption capacity was 30.78 mg·g−1. [57]
Waste walnut shellTiO2Strong interaction between BC and TiO2 effectively promotes the transfer of e in TiO2.The decolorization rate reached 96.88%, and the activity was still high after 5 catalytic experiments. [58]
Corn stoverZnO/ZnSHigher BET, porous structure and total pore volume.The maximum adsorption capacities were 135.8, 91.2 and 24.5 mg·g−1. [59]
Rice huskFe3O4presents a superior magnetic response.Displayed a preeminent adsorption performance for U(VI). [60]
Co-precipitationCorncobMgCl2 and CaCl2High surface area, mesoporous structure.The adsorption of P reached 326.63 mg·g−1. [61]
Kans grass straw (Saccharum spontaneum)FeSO4•7H2O and FeCl3•6H2OMGKB has ferromagnetism, high thermal stability, higher surface porosity, high surface area, and large total pore volume.When pH = 13.5, the adsorption capacities of As(III) and As(V) are 2.004 mg·g−1 and 3.132 mg·g−1. The adsorption effect is best at this time. [62]
Sugarcane baggase (SCB)FeCl3 and FeSO4It has a porous structure with a large number of carboxyl groups and negative charges on the surface.The adsorption capacities of Pb2+ and Cd2+ were 1.2 and 1.1 mmol·g−1, respectively. When C(Pb): C(Cd) is greater than or equal to 4:1, the magnetic adsorbent can selectively adsorb Pb2+. [63]
Peeled pine wood (Pinus massoniana)Hydrous-manganese oxide (HMO)High BET and porosity. The surface hydroxyl group increases and the pHPZC(zero charge point pH) of the carbon decreases.When pH = 5.00, the removal efficiency of lead (II) increased from 6.4 to 98.9%. [64]
Oak wood and oak barkFe2(SO4)3•nH2O (n: 6–9) and FeSO4High surface area, large porosity, high MS value.The adsorption capacities of MOWBC and MOBBC were Pb2+ 10.13 and 2.87 mg·g−1 and Pb2+: 30.2 and Cd2+: 7.4 mg·g−1. [65]
HydrothermalFresh olive waste/Carboxyl and carbonyl groups increased by about 300%.Removes approximately 100% of methylene blue and Congo red. The three adsorption cycles have a repeat utilization rate of about 80%. [66]
Pine wood/The optimum carboxylic acid content of hydrocarbon surface was obtained by oxyketone treatment.The maximum adsorption capacity of MB was 86.7 mg·g−1 and Pb(II) was 46.7 mg·g−1. [67]
wheat strawFeIron-modified hydrocarbons have higher voids, roughness and specific surface area.Under the conditions of initial pH 6, concentration of RhB 5 mg·L−1, concentration of RhB 1 g·L−1 and adsorption time 90 min, the optimal adsorption efficiency of fe modified hydrocarbons for RHB is 91%. [68]
Mg-doped grape pomace
Mg-doped corn cob
Mg-doped Miscanthus × giganteus
MgHydrogen bonding, π-π interactions, electrostatic interactions, and surface complexation all play a roleTheir adsorbability is 289.65 mg·g−1, 262.30 mg·g−1, and 232.48 mg·g−1. [69]
sugarcane bagasseNi, FeHigh BET and contains a large number of carboxyl and metal carboxyl groups, hydrogen bonds, π-π or π-eda interactions, surface complexation, and acid-base interactions.Qmax = 395.9 mg·g−1 for CV dye and 568.1 mg·g−1 for TCQmax. After 4 regeneration cycles, it has good recyclability. [70]
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

Kong, F.; Liu, J.; Xiang, Z.; Fan, W.; Liu, J.; Wang, J.; Wang, Y.; Wang, L.; Xi, B. Degradation of Water Pollutants by Biochar Combined with Advanced Oxidation: A Systematic Review. Water 2024, 16, 875. https://doi.org/10.3390/w16060875

AMA Style

Kong F, Liu J, Xiang Z, Fan W, Liu J, Wang J, Wang Y, Wang L, Xi B. Degradation of Water Pollutants by Biochar Combined with Advanced Oxidation: A Systematic Review. Water. 2024; 16(6):875. https://doi.org/10.3390/w16060875

Chicago/Turabian Style

Kong, Fanrong, Jin Liu, Zaixin Xiang, Wei Fan, Jiancong Liu, Jinsheng Wang, Yangyang Wang, Lei Wang, and Beidou Xi. 2024. "Degradation of Water Pollutants by Biochar Combined with Advanced Oxidation: A Systematic Review" Water 16, no. 6: 875. https://doi.org/10.3390/w16060875

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