Next Article in Journal
Feasibility of Carbon Dioxide as Cushion Gas in Depleted Gas Reservoirs: An Experiment Study on CO2–CH4 Dispersion during Flow Alternation
Previous Article in Journal
Multi-Stage ANN Model for Optimizing the Configuration of External Lightning Protection and Grounding Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Engineered Biochar for Metal Recycling and Repurposed Applications

by
Mehedi Hasan
,
Soumik Chakma
,
Xunjia Liang
,
Shrikanta Sutradhar
,
Janusz Kozinski
and
Kang Kang
*
Biorefining Research Institute (BRI) and Department of Chemical Engineering, Lakehead University, Thunder Bay, ON P7B 5E1, Canada
*
Author to whom correspondence should be addressed.
Energies 2024, 17(18), 4674; https://doi.org/10.3390/en17184674
Submission received: 15 August 2024 / Revised: 14 September 2024 / Accepted: 18 September 2024 / Published: 20 September 2024
(This article belongs to the Special Issue Bioenergy for Biofuels: Upgrading from Renewable Resources)

Abstract

:
Heavy metal pollution is posing significant threats to the environment and human health. Engineered biochar, derived from various biomass sources through thermochemical processes, has emerged as a promising solution for metal pollutant remediation and metal recovery. This review explores the latest advancements in the preparation, characterization, and application of engineered biochar for metal adsorption, recycling, and utilization. It begins by discussing the significance of metal adsorption and providing an overview of biochar properties. The review examines the preparation and characterization techniques, emphasizing feedstock selection, thermochemical conversion methods, and surface modifications. Mechanisms of metal adsorption, such as physical and chemical adsorption, ion exchange, and surface complexation, are critically discussed. Moreover, factors influencing metal adsorption capacity, including biochar properties, metal characteristics, and environmental conditions, are critically analyzed. The efficacy of engineered biochar in adsorbing specific metals, including heavy metals, transition metals, and rare earth elements, is reviewed with recent studies and key findings. Furthermore, the recycling and regeneration of metal-loaded biochar are discussed, focusing on recycling and repurposed application techniques alongside challenges and economic considerations. Finally, future perspectives are provided for the enlightening of future research. This review is unique in addressing the potential of metal-adsorbed biochar as a novel precursor to produce catalytical and electrochemical materials.

1. Introduction

In this era of heavy industrialization, metal resources are playing an increasingly important role in our daily lives for various industrial applications, including electronics, construction, and energy production. The extraction and processing of metal-bearing minerals and the manufacturing, utilization, and recycling of metals are becoming ever more important. The rising levels of heavy metals and other metal-containing contaminants in the environment are alarming due to their potential negative impacts on ecosystems and human health [1]. It has long been known that heavy metals such as lead (Pb), cadmium (Cd), mercury (Hg), and chromium (Cr) are nonbiodegradable and persist in the environment, posing long-term risks [2]. Specifically, Pb exposure can cause neurological damage, especially in children, while cadmium and mercury are known carcinogens that can cause kidney damage and other health issues. Cr, in its hexavalent form, is highly toxic and carcinogenic. These metals can accumulate in living organisms, leading to bioaccumulation and biomagnification through the food chain, thereby affecting entire ecosystems [3].
With the fast evolution of the production of electric, such as cell phones, smart home appliances, the booming of electric vehicles, and the development of novel batteries and clean energy technologies (e.g., wind turbines, photovoltaics, energy efficient lighting), other than the heavy metals, rare earth metals (Li, Ce, Ga, etc.) are emerging as new potential sources of metal pollutions (Figure 1). Let alone the fact that industrial activities, mining operations, agricultural practices, and improper waste disposal contribute significantly to metal pollution. Addressing this challenge is crucial for sustainable development and environmental protection [4]. Moreover, the finite nature of metal resources necessitates efficient recycling and recovery methods to ensure their sustainable use. Metal adsorption using engineered biomass-based carbon materials offers a promising solution for pollution control, value-added biomass utilization, and the potential for recovery of precious metal resources. Carbon-based adsorbents are generally more environmentally friendly and can selectively remove metals from contaminated water, allowing for subsequent recycling and reuse of the adsorbed metals [5]. These dual benefits of pollution mitigation and value-added waste utilization underscore the importance of developing effective carbon-based adsorbents.
Biochar (BC) refers to the carbon-rich material produced from the thermochemical conversion technologies of biomass (i.e., pyrolysis, gasification, and conventional and microwave-assisted hydrothermal processes) and has gained significant attention as a versatile adsorbent for environmental applications [6,7]. The pyrolysis process, which involves the thermal decomposition of biomass in an oxygen-limited environment, transforms organic materials into BC with a high surface area and porous structure [8]. BC’s surface is abundant with functional groups such as hydroxyl, carboxyl, and phenolic groups, which facilitate the adsorption of metal ions [9]. Together, these properties, high thermal stability, and biocompatibility make BC an excellent candidate for environmentally friendly metal adsorption applications. BC can be engineered after production to enhance its adsorption capacity through various modification techniques. These modifications include physical activation (e.g., steam or CO2 activation) and chemical treatments (e.g., acid or base treatments) that increase the surface area, porosity, and functional group density of BC [10,11,12]. Despite its potential and the tremendous progress being made to evaluate different BC for various metal adsorption, the current understanding of BC’s metal adsorption mechanisms and optimization strategies for different metals remains incomplete and is updating very rapidly. Also, to the best of our knowledge of the recycling and regeneration of metal-loaded BC, emphasizing desorption and thermochemical recycling techniques are not previously reported in a summarized manner. Therefore, this review aims to fill these knowledge gaps by providing a comprehensive analysis of the recent advancements in engineered BC for metal adsorption.
The uniqueness of this review lies in its integrated approach, which not only summarizes and upgrades the latest trends in BC engineering research (synthesis, modification, characterization) but also addresses the potential of the metal-adsorbed BC as a novel precursor to produce catalytical and electrochemical materials. In conclusion, this review provides a thorough examination of the current state and prospects of engineered biochar for metal recycling and repurposed applications. By highlighting key findings, addressing knowledge gaps, and proposing future research directions, it aims to advance BC-based material research.

2. Preparation and Characterization of Engineered Biochar

2.1. Feedstock Selection

The properties of BC are heavily dependent on their sources of feedstock. Careful selection of biomass feedstocks and tailored preparation techniques are crucial for creating engineered BC for specific applications. Since being invented, biochar research has always had a focus on low cost, renewability, low toxicity, easy operation, and broad application prospects for heavy metal removal from wastewater [13,14,15]. A diverse array of biomass feedstocks is used in BC production, encompassing woody biomass, crops and residues, forestry crops and residues, municipal solid waste, industrial solid waste, animal residues, sewage sludge, algal biomass, seaweeds, and other lignocellulosic materials [16,17,18,19]. Woody biomass is characterized by low moisture, low ash content, high density, and high calorific value [3]. Waste biomass, including animal waste, industrial and agricultural solid wastes, and biosolids, is characterized by high ash content, high moisture, low density, and low calorific value [20,21]. The feedstock type influences BC’s yield, functional groups, surface area, porosity, elemental composition, carbon content, cation exchange capacity, ash composition, and mineral concentrations, affecting its adsorption potential [17,22,23,24]. For example, BC from woody biomass is found to be rich in carbon. In contrast, BC from animal manure has a high potassium content and bamboo-derived BC has a higher surface area compared to pig-manure BC when pyrolyzed at the same temperature [25]. Zhao et al. [26] attributed the enhanced properties of BC, such as high carbon content, good cation exchange capacity, and high carbon storage capability to the use of various feedstocks for preparation, such as animal manures, waste wood, crop residues, food waste, aquatic plants, and municipal waste.
In terms of lignocellulosic biomass, the effects of feedstock that precursor composition, including the relative contents of lignin, hemicellulose, cellulose, and inorganic constituents, will significantly influence the effect of BC characteristics. For example, high-lignin content materials such as spruce and pine wood produce BC with higher yields and a high fixed carbon content. Lignin-rich biomaterials i coconut shells and bamboo yield macroporous BC, while cellulose-rich materials like husks produce BC with a high amount of micropores [19]. Leng et al. [27] suggested that lignocellulosic biomass, particularly wood, and woody biomass, is ideal for producing porous BC due to its low ash content, high lignin content, and preservation of the original pore structure. Other important biomasses that have gained significant attention from researchers are algae, fish scales, shells, and kelp, which are excellent feedstocks for BC production due to their high growth rate and seasonal availability [17,28]. Macroalgae BC was found to have high nitrogen, hydrogen, and ash content, high electrical conductivity, low carbon content, and lower cation-exchange capacity compared to lignocellulosic BC [11,17,28].
Yining et al. [29] highlighted the significant impact of feedstock type on BC and hydrochar production, including production rate, thermal stability, and elemental composition. The authors emphasized the challenges of seasonal variation in biomass feedstock availability and the complexities in handling, transportation, storage, and sizing due to different biomasses’ diverse physical and chemical characteristics. The nature of the feedstock, particularly its moisture content, strongly impacts the pyrolysis process. High moisture content can reduce BC yield, while low moisture content makes BC production more energy-efficient and commercially viable [30,31]. Efficient pretreatment methods, including physical pretreatments like size reduction, drying, and densification, and chemical treatments such as acid or alkaline treatments, are essential to improve BC quality for certain applications, such as adsorbent for wastewater treatment and soil additives [18]. Therefore, when selecting biomass feedstock for BC production, it is essential to carefully evaluate its composition and morphological structure, which includes the original pore structure and thermal stability. The unique properties and compositions of each biomass feedstock will partially determine the quality and suitability of BC for heavy metal adsorption.

2.2. Thermochemical Conversion Processes for Biochar Production

Thermochemical conversion produces BC through processes like pyrolysis, hydrothermal carbonization (HTC), gasification, and torrefaction, yielding bio-oil or biosyngas, with BC being a high-carbon, porous material rich in functional groups [32]. Key operational parameters, such as heating rate, temperature, and residence time, must be carefully optimized, as they significantly impact the physicochemical properties of the resulting BC like pore structures, surface area, elemental compositions, and surface functional groups [33,34]. Biochar produced at medium and high pyrolysis temperatures (350–650 °C) increases the release of volatile matter, the surface area, and the carbon content while balancing nitrogen content and effectively removing heavy metals from wastewater. In contrast, lower pyrolysis temperatures (e.g., <300 °C) produce BC with higher yield and moderate stability, making it better suited for agricultural production and soil contamination remediation [35,36]. For instance, corn-straw-derived BC removed Cu2+ more efficiently at 800 °C than at 400 °C [37]. Surface functionalization during high-temperature pyrolysis improves reactivity and adsorption capabilities [38]. Wang and Wang [39] reported that BC produced at high temperatures is effective at removing Cr(VI). Microwave-assisted pyrolysis is a novel method that accelerates the process without requiring biomass shredding or drying. This technique uniformly transfers thermal energy to the biomass’s functional groups, producing BC with a larger surface area (BET up to 450–800 m2/g) and more functional groups compared to unmodified BC [40]. Kazemi et al. [40] investigated that microwave-engineered BC enhanced heavy metal removal, soil water-holding capacity, and cation exchange capacity. Copyrolysis, involving multiple biomass feedstocks, has recently garnered attention for its ability to influence the pore structure of BC differently than single biomass pyrolysis [18].
The conversion of lignocellulosic biomass into BC involves various reaction parameters and techniques. Cellulose decomposes through slow pyrolysis, which has a longer residence time and lower heating rates, and fast pyrolysis, leading to rapid volatilization and the formation of levoglucosan, which can dehydrate into hydroxymethyl furfural, bio-oil, or syngas. Then, hydroxymethyl furfural undergoes aromatization, condensation, and polymerization to regenerate solid BC. Moreover, hemicellulose decomposes similarly, forming oligosaccharides through decarboxylation and depolymerization at lower temperatures (220–315 °C). Lignin decomposes over a wide range of temperatures (150–900 °C) through complex processes involving β-O-4 bond cleavage and radical formation [30].
Another relatively new technique, HTC, involves dissolving biomass in water and operating temperatures between 180 and 250 °C in a closed reactor, achieving a high carbon yield of 30–50% [41]. Amalina et al. [19] reported that during HTC, char is produced through the depolymerization and cross-linking of intermediates. Alongside pyrolysis, lignin components that remain undissolved in the liquid phase are transformed into hydrochar. Longer hydrothermal treatment reduces polar functional groups in BC, favoring nonpolar ones [42]. Ponnusamy et al. [43] highlighted the disadvantages of HTC such as reduced porosity, limited surface area, thermal resistance, and a less aromatic structure. Biomass gasification involves four stages: drying, pyrolysis, partial oxidation, and reduction. The quantity and quality of the resulting BC depend on the effective control and optimization of these stages [44]. Another method, called torrefaction or low-temperature pyrolysis, produces BC through pyrolysis at lower temperatures (<300 °C), altering biomass properties such as energy density, particle size, surface area, heating time, and moisture content, and generally improves the credibility of the biomass for further processing [45]. Each thermochemical conversion process offers distinct advantages for BC production, contributing to the efficient utilization of biomass resources as heavy metal adsorbents.

2.3. Surface Modification and Functionalization Methods

Biochar is valued for its large specific surface area, highly developed porous structure, and diverse functional groups, including carboxylic, carbonyl, ester, hydroxyl, phenolic, pyridine-N, pyrrole-N, and quaternary-N, which contribute to its efficacy as an adsorbent. Recent research shows that modified BC has a higher adsorption capacity than unmodified BC, driving interest in developing high-performance BC materials through various modification techniques [17,46]. Engineered BC is a biochar that has been deliberately modified through physical, chemical, or biological processes to improve its properties for targeted applications, such as adsorption or soil enhancement. These modifications typically enhance key characteristics like porosity, pH, cation exchange capacity, and specific surface area, leading to superior heavy metal adsorption capabilities compared to untreated BC [27,46]. Kumar et al. [21] reviewed modified agrowaste BC materials, which showed a wide range of adsorption capacities and achieved over 99% removal of toxic heavy metals from aquatic systems. Figure 2 illustrates various modification techniques and their effects on the properties of the engineered BC.
Physical activation techniques like steam, ball milling, magnetization, and microwave modification focus on altering BC’s structure and surface properties [24,47]. Steam activation of BC creates new pores and enhances surface functional groups, whereas microwave modification increases both the surface area and functional groups, thereby improving the BC’s ability to adsorb heavy metals [24,48]. Using CO2 as a gas medium or adding chemicals like KHCO3, K2CO3, NaHCO3, or Na2CO3 during pyrolysis is a potentially more cost-effective activation route for creating BC-based materials with a hierarchical porous structure, unlike steam activation, which has higher costs [49,50]. Ball milling reduces particle size, increasing adsorption sites like carbon nanotubes but with better dispersion in water [22,51]. Magnetizing BC enhances its ease of separation, thereby lowering recovery costs. However, caution is needed as magnetization can cause structural deformation and reduction in metal adsorption capacities. Optimizing the timing of magnetization and conducting synthesis in alkaline media can improve its heavy metal removal efficiency [50,52]. Shang et al. [53] magnetized BC derived from herb residues with Fe3O4 nanoparticles, achieving efficient Cr(VI) adsorption and easy separation using a magnetic field. Similarly, Wang et al. [54] synthesized magnetic rice husk BC, reporting maximum adsorption capacities of 129 mg/g for Pb(II) and 118 mg/g for U(VI), significantly higher than unmodified BC. Microwave-assisted pyrolysis is increasingly popular for BC modification due to its rapid heating rate and lower energy consumption, making it more environmentally friendly than conventional methods. Studies have shown that microwave-modified BC has significantly improved microporosity, surface area, and functional group presence, enhancing its yield and adsorption capacity for contaminants like As5+, Zn2+, Cu2+, and silver [55,56]. Chemical activation of BC with oxidants, acids, or alkalis improves the specific surface area, porosity, polarity, charge characteristics, oxygen-containing functional groups, and active sites for the chemisorption of heavy metals [57,58,59]. However, inappropriate activator concentrations, such as high KOH levels, can hinder heavy metal stabilization by expanding the lignin structure inside BC. Therefore, moderate concentrations of activators are generally recommended [60]. Surface oxidation is widely used to introduce oxygen-containing functional groups to BC, enhancing its specific surface area, pore volume, active sites, and hydrophilicity. This improves its adsorption performance and application properties through reactions such as oxidation, hydroxylation, carboxylation, and aldehydation [61]. Key functional groups like carbonyl, hydroxyl, and carboxyl enhance metal adsorption via hydrogen bonding and electrostatic attraction [62]. Common oxidizing agents include H2O2, O3, KMnO4, and HNO3, which also boost BC’s hydrophilicity [63]. For instance, Fan et al. [64] found that HNO3/H2SO4 and NaOH/H2O2-modified reed BC increased oxygen functional groups, the oxygen-to-carbon ratio, and specific surface area and pore volume. Liu et al. [65] reported an increase in specific surface area from 2.68 m2/g to 86.8 m2/g and an increase in adsorption capacity for Cr(VI) and Cd2+ from 1.29 mg/g to 3.20 mg/g. Biological modification of BC involves producing BC from pretreated feedstocks via anaerobic digestion after biological engineering. This process generates BC with increased surface area, anion exchange capacity, cation exchange capacity, and altered pH values, making it effective for heavy metal adsorption processes in wastewater treatment [66]. BC can serve as a substrate for immobilizing microorganisms and enzymes, creating composite materials. Microbes on BC secrete organic acids and metabolites, introducing new functional groups (-COOH, -OH, -NH2) that enhance adsorption. Some microbes produce biogums, increasing porosity and adsorption capacity. This method uses fewer chemicals, has a milder preparation process, and consumes less energy, aligning with green and low-carbon principles [67,68].
Furthermore, coating BC with metal oxides or graphene can further increase its adsorption capacity and efficiency in removing heavy metals, while utilizing BC as a carrier for nanoparticles enhances their effectiveness in wastewater treatment applications [69,70]. Tho and coauthors [71] reported that modifying cassava root husk-derived BC with ZnO nanoparticles significantly improved its adsorption capacity for As(III), Cd(II), Pb(II), and Cr(VI). This study demonstrates that incorporating ZnO nanoparticles into BC enhances its ability to remove heavy metals, paving the way for more effective heavy metal removal strategies using modified BC. Despite the benefits of these modification techniques, further research is necessary to understand the effects of modification, the costs, and the potential for secondary pollution in real wastewater treatment applications.

2.4. Characterization Techniques of Biochar Materials

Comprehensive characterization is crucial to decode BC properties and to evaluate BC’s suitability for specific applications. Key properties such as surface area, charge, porosity, pH, mineral content, and binding sites are essential to understand, as they influence BC’s effectiveness in adsorption mechanisms. Techniques such as scanning electron microscopy (SEM), Fourier-transform infrared spectroscopy (FTIR), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), nuclear magnetic resonance spectroscopy (NMR), BET analysis, thermogravimetric analysis (TGA), proximate and ultimate analysis, and Raman spectroscopy are used to examine BC’s structure, surface functional groups, and elemental composition (Figure 3) [27,72].
FTIR is used to analyze BC’s surface functional groups, which may change significantly when the chars are produced under various temperatures [73]. XPS is used to ascertain the valence states of chemical elements by examining the energy distribution of emitted photoelectrons. SEM is used to characterize the surface structures of BC, providing detailed images of its microporous and mesoporous distributions and pore configurations [74]. SEM and TEM are useful for elaborating surface morphology before and after adsorption, with TEM offering higher resolution and the ability to assess the electronic state of BC. STEM, like SEM and TEM but with higher resolution, provides atomic-scale BC characterization. SEM combined with EDX (SEM-EDX) is employed to analyze the composition of BC and identify surface components. However, SEM-EDX is incompatible with organic pollutants. EDS, often paired with these microscopes, identifies the molar fraction of surface elements [34].
XRD is widely used to determine the crystallinity and crystal structure of BC. The diffractogram shows attributes of amorphous nebulous material formed above 350 °C [17,72]. The nanocrystals produced exhibit a crystalline structure, reflected in sharp and robust XRD peaks, with a particle diameter increasing over time. Thermal analysis, particularly TGA, is used to determine the physicochemical properties of materials as the temperature rises [75]. TGA examines the thermal decomposition/oxidation behavior of various samples and can also help in identifying the ignitability of BC and BC mixtures [76]. It analyzes the synergistic activity between components, aiding in understanding the heating process and sample features [77]. In a typical TGA test, the heating starts at room temperature and increases to 1000 °C, with typical rates between 10 and 20 °C/min [78]. NMR spectroscopy is used to determine the structural composition of BC by analyzing the resonance frequencies of atomic nuclei using a magnetic field and RF pulses. It identifies carbon functional groups, aromatic ring formation, and overall BC molecule structure [73,79]. NMR distinguishes aliphatic and aromatic hydrocarbons and compares the stability and carbonization of different biochars. However, ferromagnetic minerals in BC and low signal-to-noise ratios in high-temperature pyrolysis products can interfere with NMR signals [80].
Surface area, total pore volume, and pore diameter were determined using the Brunauer–Emmett–Teller (BET) method with a surface area and porosity analyzer. In this method, nitrogen adsorption isotherms were utilized to measure the surface area of BC by assessing the multilayer adsorption capacity at varying nitrogen partial pressures [19].
Proximate analysis, which includes measuring moisture content, fixed carbon, and volatile matter, follows different standard methods like British Standards (BS), Chinese Standards, ASTM (American Society for Testing and Materials), and ISO (International Organization for Standardization). Residual matter was calculated by the difference. Electrical conductivity (EC) and pH were measured in a 1:10 DI water suspension. Cation exchange capacity (CEC) was determined using the ammonium acetate extraction method. Elemental analysis quantifies elements like nitrogen (N), hydrogen (H), carbon (C), and sulfur (S) using sophisticated instruments based on complete combustion in oxygen. Atomic ratios of elements like O/C and H/C were calculated to indicate the polarity and aromaticity of the BC materials [81]. These comprehensive characterization techniques can provide valuable data to confirm whether the characterized BC is optimized for specific applications.
Figure 3. Overview of engineered biochar characterization techniques [73,81,82].
Figure 3. Overview of engineered biochar characterization techniques [73,81,82].
Energies 17 04674 g003

3. Mechanisms of Metal Adsorption on Biochar

Various mechanisms, such as physical adsorption, chemical adsorption, and other potential processes, contribute to metal adsorption by engineered BC (Figure 4).

3.1. Physical Adsorption

Physical adsorption, known as physisorption, occurs when metal ions adhere to the BC surface through Van der Waals forces [83]. This process is typically weak and reversible, with adsorption and desorption continuously competing at varying rates depending on the specific conditions of the system [84]. This process largely depends on the BC’s surface area and pore structure, with higher surface areas and well-developed micro-, meso-, and macropores enhancing adsorption capacity. BC activation/modification processes can increase these properties, improving adsorption efficiency [10,27]. Liu et al. [85] prepared BC from rice straw biogas residue modified with KOH, which demonstrated that KOH modification significantly enhanced the specific surface area (from 273.26 to 2372.51 m2/g) and pore volume (from 0.24 to 1.20 cm3/g), resulting in the creation of BC with numerous adsorption sites and a high adsorption capacity (209.65 mg/g) of mercury.

3.2. Chemical Adsorption

Chemical adsorption, or chemisorption, involves the formation of stronger bonds between metal ions and functional groups on the BC surface. This process is typically irreversible and significantly depends on the BC’s chemical composition and surface properties. Various mechanisms can contribute to chemical adsorption, including ion exchange, complexation, electrostatic attraction, and precipitation.

3.2.1. Ion Exchange

Ion exchange is a process in which metal ions in the solution replace ions originating from the BC surface. This mechanism is facilitated by the presence of exchangeable cations like H+, Na+, K+, Ca2+, and Mg2+ on BC. The efficiency of ion exchange depends on the BC’s cation exchange capacity (CEC), which is influenced by its surface functional groups and the degree of activation. A higher cation exchange capacity (CEC) correlates with an increased rate of adsorption [86]. Wu et al. [87] showed that MgCl2 modification of coconut shell BC significantly enhanced the release of Mg2+ ions into the solution, leading to a substantial increase in adsorption capacities (qe) for Pb (192.36 mg/g) and Cd (105.46 mg/g) compared to unmodified BC (Pb 3.93 mg/g and Cd 1.78 mg/g), thus improving its ion exchange capacity [87]. In aqueous solutions, metal ions like Cd2+ and Pb2+ replace exchangeable cations such as H+ and Na+ on the surface of BC. These metal ions then attach to functional groups, including carboxyl (-COOH) and hydroxyl (-OH) groups. Studies have shown that BC produced from palm leaves, rich in oxygen-containing functional groups, effectively exchanges ions with heavy metals like chromium (Cr3+) and lead (Pb2+), reducing their concentration in contaminated water [88]. Another study demonstrated that ion exchange and precipitation were the predominant mechanisms for the adsorption of heavy metals (Pb, Cd, and Ni), as these metals effectively exchanged ions with the mineral constituents of the BC [89].

3.2.2. Complexation

Complexation involves the formation of stable, often covalent, bonds between metal ions and functional groups on the BC surface [90]. This mechanism significantly enhances the stability and retention of metal ions in BC. Metal ions (e.g., Cu2+, Fe3+) interact with specific functional groups such as carboxyl, hydroxyl, amino (-NH2), and thiol (-SH) groups to form metal-ligand complexes. For instance, Pb(II) and Cd(II) have been shown to form strong complexes with oxygen, nitrogen, and sulfur functional groups on BC derived from A. compressus and reed straw, leading to the efficient removal of these metals [90,91]. Similarly, BC modified with FeCl3 has demonstrated effective complexation with As(V) by forming complex compounds with iron ions [92]. The stability of these complexes depends on the metal ions and the specific functional groups present and thus effectively removes heavy metals from contaminated aqueous solutions.

3.2.3. Electrostatic Attraction

Electrostatic attraction occurs when metal ions are attracted to oppositely charged sites on the BC surface. This mechanism is particularly effective for adsorbing cationic metals and is influenced by the surface charge of the BC, which is determined by its pH and functional groups. Negatively charged sites on the BC surface, such as deprotonated carboxyl and hydroxyl groups, attract positively charged metal ions like Ni2+ and Zn2. As an example, a study by Chen et al. [93] indicated that Zeta potential (ZP) values, which indicate the electrostatic potential of biomaterials in various pH solutions, significantly decreased between pH 2 and 4 and stabilized around −30 mV at pH 4–10, suggesting that the negatively charged surfaces of pomelo-peel-derived BC facilitate the adsorption of Pb(II) and Cu(II) through electrostatic attraction. The authors also reported that the BC surface was negatively charged and contained carboxyl and hydroxyl functional groups, suggesting that electrostatic attraction was the dominant mechanism in removing those pollutants. Similarly, Li et al. [94] demonstrated that the adsorbents exhibited a negative charge when the pH of the solution exceeded the point of zero charge (pHpzc), facilitating the electrostatic binding of acetate-based novel BC to Pb(II) and Cu(II).

3.2.4. Precipitation

Precipitation is a mechanism by which BC adsorbs heavy metals. It involves the transformation of dissolved metal ions into insoluble solid phases on the BC surface. This process is influenced by the chemical composition of the BC, including the presence of mineral components and functional groups that can induce nucleation and growth of metal precipitates. BC has a high mineral content, such as calcium, magnesium, potassium, phosphorus, etc. [95]. BC’s high mineral content, mainly when derived from feedstocks rich in inorganic components, provides a favorable environment for precipitation reactions. These minerals can react with metal ions to form insoluble hydroxides, carbonates, or phosphates, which are then deposited onto the BC surface. For example, BC derived from coconut shells demonstrated that anions (e.g., CO32−, PO43−, and OH) released from the BC reacted with metal ions (Pb, Cd) to form mineral precipitates such as PbCO3, Pb3(CO3)2(OH)2, 2PbCO3·Pb(OH)2, Pb(H2PO4)2, CdCO3, Cd3(PO4)2, Cd(OH)2, and Cd3SiO5 [87]. The surface functional groups of BC, such as carboxyl (-COOH) and hydroxyl (-OH) groups, also play a significant role in the precipitation mechanism. Cr(III) reacts with (-OH) of BC and forms Cr(OH)3 precipitation and thus helps immobilize Cr(III) [96]. These groups can bind metal ions, forming stable metal-organic complexes that act as precursors for further precipitation. Studies have shown that Mg-loaded BC can enhance the precipitation of Cd(II) and Cu(II) by forming stable CdCO3 and Cu2Cl(OH)3 complexes, respectively [97].

3.2.5. Other Mechanisms

Biochar adsorbs heavy metals through several mechanisms beyond these dominant mechanisms. Redox reactions involve the reduction in metal ions facilitated by BC’s surface functional groups and redox-active components. For example, the redox reaction between oxygen-containing functional groups on the BC surface and Cr(VI) significantly contributes to adsorption, where these functional groups donate electrons to reduce Cr(VI) to Cr(III), which can then be released into the solution through ion exchange with K+ in the BC [98]. π-π interactions occur when metal ions coordinate with the aromatic rings of BC. Liu et al. [99] demonstrated that BC derived from banana stems and leaves effectively adsorbed Pb and Cd, as indicated by FTIR analysis showing reduced -CH, C=C, and C=O groups postadsorption, suggesting that these metal ions interacted with the π electrons during the adsorption process. Hydrogen bonding involves metal ions forming bonds with hydroxyl and carboxyl groups in BC, enhancing adsorption. A study showed that As(V) was adsorbed onto the BC surface via hydrogen bonding with the -COOH groups present on the BC [100]. These mechanisms collectively contribute to the effective removal and stabilization of heavy metals in contaminated environments.

4. Factors Influencing Metal Adsorption Capacity of Biochar

Various factors, including the properties of BC, the characteristics of the metal ions, and the environmental conditions, influence the ability of BC to adsorb heavy metals. Fundamental BC properties include surface area, porosity, functional groups, and cation exchange capacity, which collectively determine its affinity for metal ions [101]. Additionally, the ionic radius, charge, and redox potential of the metal ions significantly impact their adsorption behavior [102,103,104]. Environmental conditions, including pH, temperature, competing ions, etc., further influence the interaction between BC and metal ions, thereby modulating the overall adsorption capacity. A summary of the factors influencing metal adsorption capacity is shown in Table 1.

4.1. Effect of Biochar Properties on Metal Adsorption

The efficacy of BC as an adsorbent for metal ions is profoundly affected by its inherent properties. These properties determine the interaction between BC and metal ions, affecting overall adsorption capacity.

4.1.1. Surface Area and Porosity

The surface area and porosity of BC are fundamental determinants of its adsorption capacity. An increased surface area offers a greater number of active sites available for metal ion binding, while higher porosity, particularly the presence of micropores and mesopores, enhances accessibility and facilitates ion transport for adsorption [105,106]. For instance, BC derived from bamboo and modified with guanidine phosphate demonstrated an increase in surface area from 196 to 458 m2/g, which markedly enhanced its adsorption capacities for Pb(II) (166.2 mg/g), Cu(II) (81.7 mg/g), and Cd(II) (60.3 mg/g) [107]. This modification not only increased the surface area but also enhanced the porosity, thereby providing more pathways for metal ions to reach the active sites on the BC.

4.1.2. Surface Functional Groups

Functional groups on BC surfaces significantly influence their metal adsorption capacity through mechanisms such as complexation, ion exchange, and electrostatic interactions. Carboxyl (-COOH) and hydroxyl (-OH) groups are particularly influential in enhancing metal adsorption. These functional groups can form strong complexes with metal ions, facilitating their removal from aqueous solutions. Zeng et al. [108] demonstrated that oxygen-containing functional groups (COOH and -OH) on BC surfaces act as electron donors, reducing Cr(VI) to Cr(III) and resulting in the formation of Cr precipitates and complexes. Amino (-NH2) groups are particularly beneficial for removing heavy metals such as Pb2+ and Cd2+ due to their chelating properties, which enable the formation of strong metal-ligand complexes [109,110]. Phosphate groups (-PO43−) on BC surfaces are advantageous for removing metals like Pb2+ and Cd2+ through precipitation reactions. Deng et al. [111] demonstrated that incorporating chloro-phosphate into BC significantly enhances the adsorption efficiency of heavy metals (Pb2+, Cd2+, and Cu2+) due to the specific precipitation mechanisms, forming Pb5(PO4)3Cl, Cd5(PO4)3Cl, and Cu3(PO4)2 precipitates.

4.2. Effect of Metal Characteristics

The characteristics of metal ions, including ionic radius, charge, speciation, and oxidation state, significantly influence their adsorption capacity on biochar. Ionic radius and charge are crucial determinants of metal ion interactions with BC surfaces. Metal ions with smaller ionic radii and higher charges typically exhibit stronger adsorption due to the higher charge density, which enhances electrostatic interactions and complexation with functional groups in BC. For example, in a study, Cr(III) demonstrated more effective adsorption onto Fe3O4-loaded BC due to its smaller ionic radius (0.0615 nm) compared to that of Cu(II) (0.073 nm) and Zn(II) (0.074 nm) [102].
The speciation and oxidation state of metal ions also play a vital role in adsorption. Metal ions in different oxidation states exhibit varying affinities for BC surfaces. Cr(VI) and Cr(III), for example, display distinct adsorption behaviors; Cr(VI) is often adsorbed through reduction to Cr(III) on BC surfaces, followed by complexation or precipitation. Additionally, the speciation of metal ions, which is dependent upon pH and the presence of other ligands, influences their availability and reactivity toward BC [112]. For example, the speciation of As and Cr significantly affects their adsorption capacity, with As(V) and Cr(III) generally exhibiting higher adsorption efficiencies compared to As(III) and Cr(VI) due to differences in their chemical stability and reactivity [100,113].

4.3. Environmental Conditions

4.3.1. Effect of pH

The pH of the medium significantly impacts the metal adsorption capacity of BC by affecting both the surface charge of BC and the speciation of metal ions. The point of zero charge (pHpzc) is crucial in determining the surface charge of BC; at pH values above the pHpzc, the BC surface becomes negatively charged, which enhances the electrostatic attraction of positively charged metal ions. Conversely, at pH values below the pHpzc, the BC surface is positively charged, reducing the adsorption capacity for cationic metals due to electrostatic repulsion. For instance, Biswal et al. [114] showed that polymer-modified BC exhibited enhanced Cr(VI) adsorption at low pH (pH 2) due to the protonation of surface functional groups, resulting in strong electrostatic attraction with anionic Cr(VI) species and its reduction to Cr(III). In contrast, the adsorption capacity for cationic metals such as Cu2+ increases with higher pH, as shown by Zhang et al. [115], where cow manure BC achieved peak adsorption at the pH of 5. However, a further increase in pH reduced the adsorption slightly due to the formation of metal hydroxides. This trend is further supported by studies on the removal of contaminants like textile dyes, where higher pH (7) levels enhanced the adsorption due to increased negative charges on BC surfaces, facilitating the attraction of positively charged dyes [116].

4.3.2. Effect of Temperature

Temperature significantly influences the adsorption capacity of BC for heavy metals, with most studies indicating an endothermic process [117,118,119]. The adsorption rate and capacity significantly increase with rising temperatures due to the higher kinetic energy of adsorbate molecules, which enhances the ion transfer rate [120]. For instance, Zhang et al. [118] demonstrated that the adsorption capacity of Pb(II) and Cd(II) onto magnetized activated carbon (MAC) derived from rape straw powder increased significantly as the temperature rose from 24.8 to 44.85 °C. Researchers also concluded that, as temperature increased, the adsorption spontaneity rose (ΔG0), with positive enthalpy changes (ΔH0) for Pb(II) and Cd(II) adsorption by MAC (18.02 kJ/mol and 15.32 kJ/mol, respectively), indicating an endothermic process favoring higher temperatures. Similarly, Xu et al. [117] found that higher temperatures (39.85 °C) improved the adsorption capacity (160 mg/g) of U(VI) by BC derived from water hyacinth.
However, some studies suggest an exothermic process, where adsorption decreases with increasing temperature. Harsha et al. [121] reported a decline in Cu(II) adsorption by H2SO4-modified Annona reticulata seed-based BC as temperature increased from 26.85 to 56.85 °C. He proposed that increased temperature enhances cation movement and desorption, with optimal cation separation by BC observed at 30 °C, indicating that lower temperatures are more favorable for removal. Singh et al. [122] observed that Cr(VI) adsorption efficiency decreased with rising temperature for both BCs, indicating an exothermic process. The optimal temperature for Cr(VI) adsorption onto bacterial BC was found to be 24.85 °C, as higher temperatures reduce adsorption forces and efficiency due to increased molecular separation and decreased solution viscosity. These results indicate that the adsorption process can exhibit either endothermic or exothermic behavior, depending on the specific interactions between BC and metal ions.

4.3.3. Effect of Competing Ions

In wastewater or natural water environments, the presence of various competing ions such as Na+, K+, Ca2+, Mg2+, Cl, NO3, and SO42− can significantly influence the adsorption capacity [123]. These competing ions coexist in the solution and their concentrations can influence the adsorption efficiency. These competing ions have high bonding affinities toward BC and occupy or block a greater number of adsorption sites on the BC surface, resulting in a substantial decrease in its adsorption capacity [124].
For instance, Wang et al. [125] investigated the effects of various ions (K+, Fe3+, Ca2+, and Mg2+) on the adsorption of Cr(VI) by modified BC. The authors found that the presence of K+ ions reduced the adsorption capacity of BC. Additionally, Ca2+ and Mg2+ had little impact on the adsorption. The competitive adsorption process is attributed to the intrinsic ions occupying the available adsorption sites on the BC surface, thereby diminishing the removal efficiency of heavy metal ions.
Table 1. Summary of factors influencing metal adsorption capacity.
Table 1. Summary of factors influencing metal adsorption capacity.
FactorsDescriptionImpact on Metal AdsorptionExampleReference
Surface areaTotal area available for metal ion binding on the BC surfaceA larger surface area increases the number of active sites, enhancing adsorptionIncreased surface area from 273.26 to 2372.51 m2/g improved Hg adsorption capacity[85]
PorosityMicropores and mesopores in BC affect the accessibility of adsorption sitesHigher porosity improves accessibility and adsorption capacityAlkali treatment of BC from Douglas fir enlarged micropores to mesopores, enhancing adsorption The KOH-activated BC effectively adsorbed Cr(VI), Pb(II), and Cd(II) with capacities of 127.2, 140.0, and 29.0 mg/g, respectively[126]
Functional groupsThe chemical groups on BC that interact with metal ionsDifferent functional groups enhance interaction with specific metalsCr(III) immobilized through complexation with the -COOH group of the modified BC[127]
Cation exchange capacity (CEC)BC’s ability to exchange cations with metal ionsHigher CEC improves the BC’s ability to adsorb metal ionsCEC of all BC samples increased by 1.33 to 2.40 times post-modification, enhancing the immobilization efficiency for Cd(II) and Cu(II)[59]
Ionic radiusThe size of the metal ion affects how easily it can fit into adsorption sites on BCSmaller ionic radius may enhance adsorption due to better site fitCr(III) exhibited higher adsorption efficiency due to its smaller ionic radius (0.0615 nm) compared to Cu(II) (0.073 nm) and Zn(II) (0.074 nm)[102]
ChargeThe electrical charge of metal ions influencing electrostatic interactions with BCA higher charge can lead to stronger electrostatic attractionThe crayfish shell BC possessed a higher positive charge, resulting in stronger electrostatic interactions with As[128]
SpeciationThe chemical form of the metal ion in solution, which affects its interaction with BCDifferent species may have varying adsorption efficienciesThe speciation of arsenic significantly influences its removal efficiency; As(V) is more readily adsorbed compared to As(III) due to its greater stability and inertness[100]
Oxidation stateThe oxidation state of the metal ion, which affects its reactivity and interaction with functional groups on BCDifferent oxidation states influence the adsorption mechanismsCr(VI) reduction to Cr(III) impacts its adsorption on BC[113]
pH of the mediumThe acidity or alkalinity of the solution affects the ionization of functional groups and metal speciationpH changes can enhance or reduce adsorption depending on the metal typeCationic metals such as Cu2+ increase with higher pH, while Cr(VI) adsorption decreases[114,115]
TemperatureThe thermal condition of the environment affecting adsorption kinetics and capacityIncreased temperature can enhance or reduce adsorption depending on the processHigher temperatures improved the adsorption capacity (160 mg/g) of U(VI) by BC[117]
Competing ionsThe presence of other ions in the solution that can compete with metal ions for adsorption sitesA high concentration of competing ions can decrease metal adsorptionAnions (Cl, SO42−, PO43−, and NO3−) in wastewater competed with Cr(VI) for adsorption sites, leading to a reduction in adsorption capacity[127]

5. Engineered Biochar for the Adsorption of Specific Types of Metals

5.1. Adsorption of Heavy Metals

Heavy metals such as Pb, Cd, Hg, and Cr are persistent pollutants that can contaminate soil, water, and air, posing severe risks to ecosystems and human health [129]. This underscores the critical necessity for implementing effective treatment technologies to safeguard environmental sustainability and public health. There are several methods for removing heavy metals from various sources, which can be divided into three main categories: physical, chemical, and biological. These include precipitation/coagulation, ion exchange, adsorption, membrane filtration, electrochemical treatment, bioremediation, chelation, phytoremediation, chemical oxidation/reduction, and acid/base treatment [130]. Each method has specific advantages and applications in environmental cleanup. BC offers a multifaceted approach to heavy metal remediation with several advantages. It is not only more cost-effective compared to technologies like membrane filtration and electrochemical treatment due to lower operational and maintenance costs but it also can be regenerated and reused multiple times by desorption of heavy metals using appropriate eluents, extending its lifecycle and reducing waste generation [131]. Moreover, once heavy metals are adsorbed onto BC, they are generally immobilized and less prone to leaching or release back into the environment, providing long-term effectiveness [132]. Table 2 presents the adsorption capacities of different engineered BC for heavy metal removal.
Biochar from rice husks and sunflower husks was obtained using stepwise pyrolysis, with temperatures reaching up to 750 °C, resulting in 34% and 32% BC yields, respectively. Sorption isotherm experiments demonstrated high adsorption capacities for heavy metals (Cu, Zn, Pb) due to the BC’s porosity and specific surface area [133]. Moreover, BC derived from secondary sedimentation (SS) and kitchen waste (KW) were evaluated for their heavy metal (HM) adsorption capacities, with KW-derived BC, notably KO-700, demonstrating significant efficacy in removing copper (Cu) and lead (Pb) from aqueous solutions (Cu: 1.25 mg/L after 72 h; Pb: 1.04 mg/L after four h). Characterization emphasized BC properties such as specific surface area, pore volume, and surface functional groups, elucidating their role in adsorption mechanisms, including electrostatic attraction and chemical precipitation [134]. Another study compared BC from waste biomass and plastic char from PET, PE, and PVC and confirmed the efficiency of these chars in removing heavy metals. Biochars showed the highest sorption capacities for heavy metals: SC 11 absorbed Ni (99.01%), Cu (91.60%), Cd (99.24%), Fe (10.95%), Pb (95.52%), and BC 11 absorbed Cr (99.06%). For plastic chars, PV 11 absorbed Ni (43.32%) and Cu (90.96%), while PT 11 absorbed Cd (28.31%), Fe (4.0%), Pb (70.07%), and Cr (99.01%) [135]. Biochars had higher metal adsorption capacities than plastic chars, even without activation. Derived sustainably from biomass like rice and sunflower husks, BC exhibits high adsorption capacities for metals such as Cu, Zn, and Pb due to its porous structure. It offers cost-effective and reusable capabilities, immobilizing metals effectively to prevent environmental leaching. Challenges include variability in BC properties and the need for optimized production processes. Despite these, BC demonstrates superior efficacy over plastic-derived pyrolysis chars for metal adsorption, highlighting its potential as a sustainable solution for environmental remediation efforts.
Table 2. Summary of heavy metals adsorption capacities of different engineered biochars.
Table 2. Summary of heavy metals adsorption capacities of different engineered biochars.
Biochar SourcesTemperature (°C)ProcedurePollutantAdsorption Capacity (mg/g)Sorption MechanismReference
Rice straw100–700Stepwise pyrolysis, including drying, grinding, and slow pyrolysis under anaerobic conditions. The BC was sonicated to form colloidsCr(III)2.4–14.1Physical adsorption due to high specific surface area and functional groups[136]
Eupatorium adenophorum500Mixed with magnesium nitrate and ferric chloride, dried, and then pyrolyzed in an oxygen-free environmentPb(II)252.70Precipitation and cation exchange[137]
Eupatorium adenophorumj500Mixed with magnesium nitrate and ferric chloride, dried, and then pyrolyzed in an oxygen-free environmentCd(II)156.60Precipitation and cation exchange[137]
Sargassum hemiphyllum700Dried, ground, sieved, pyrolyzed in muffle furnace 10 °C·min−1Cu(II)75–120Physical sorption, ion exchange, surface complexation, surface precipitation, electrostatic interaction, metal–π interaction[138]
HFO-BC (BC with Fe(III) and NaOH treatment)-Treated with 1 mol/L FeCl3 solution, stirred in 5% NaOH solution, washed, driedCd(II), Cu(II)29.9, 34.1Inner-sphere complexation, migration from solution to adsorbent surface, pore diffusion[139]
Sulfonated BC (SBC)180Carbonization and sulfonation with A. compressus, H2SO4 treatment, washing, and dryingPb(II), Cd(II) 191.07, 85.76Complexation, ion exchange, electrostatic interaction[90]
Nix-MnO2/BC-Hybrid BC precursor from CS and RH biomasses pyrolysis; surface-functionalized with Ni-MnO2 nanorods via ultrasonication, hydrothermal treatment, and calcinationLi+89.2Physisorption via electrostatic attractive forces; monolayer adsorption according to Langmuir model[140]
Magnetic EP BC400, 800Enteromorpha prolifera biomass (126 g) treated with 120 mL FeCl3 (2 mol/L) solution, stirred at 80 °C for 2 h. Ferric hydroxide-coated biomass separated via centrifugation, dried at 80 °C overnight. Pyrolysis at 400 or 800 °C under N2 with heating rate of 5 °C/min and N2 flow rate of 300 mL/min. Washed with 1 M HCl and distilled water, dried at 80 °CCr(VI)95.23Enhanced adsorption due to increased surface polarity and specific surface; likely includes physical adsorption due to magnetic modification[141]
Coffee ground650SCG/FO mixtures pyrolyzed in a quartz tube furnace with CO2 atmosphere. Mixtures of SCG and FO in mass ratios SCG1/FO0.33, SCG1/FO0.5, SCG1/FO1, and SCG1/FO2. Pyrolysis at 10 °C/min heating rate to 650 °C, held for 60 min, cooled to 25 ± 2 °CSb(V), Cd(II), Ni(II)Sb(V): 7.0 (single mode), 9.3 (multiple mode); Cd(II): 12.7; Ni(II): 16.8Ternary complexation on iron oxides surface; competition for adsorption sites in multiple mode reduces Cd(II) and Ni(II) uptake but enhances Sb(V) removal[142]
Porous BC600Corn straw powder soaked in 25% ZnCl2 for 20 h, dried, pyrolyzed at 600 °C for 1.5 h with N2, mixed with Fe(NO3) and Zn(NO3)2, pH adjusted to 11, stirred, red-brown precipitate formedCe(IV)453.518Chemisorption via surface active sites fits quasisecondary kinetic model better; adsorption influenced by temperature and is heat absorbing[143]
Palm tree fronds650Carbonization at 650 °C under inert N2 atmosphere. Oxidation with 8 M HNO3Eu3+123Reaction equation: Proposed two-step mechanism: Cation exchange of carboxylic protons with Eu3+, followed by complexation of Eu3+ with carboxylate groups[144]
Orange peel600Orange peel cut, washed, oven-dried at 60 °C, heated at 600 °C under oxygen-limited conditions for 2 h, ground and sieved to <0.15 mmLa(III)55.57High adsorption capacity; mesoporous structure with well-developed porosity and surface functional groups[145]
Corn cob600A mixture of 6.5 g KCl/NaCl (1:1) and 3 g corncob powder was dissolved in 30 mL DI water and soaked for 12 h at 50 °C. In the final hour, 1 g of Na2S, Na2S2O3, Na2SO3, or Na2SO4 was added. The precursors were then filtered, dried at 80 °C for 12 hPb(II), Cu(II)421.8 for Pb(II), 185 for CuIon exchange, electrostatic interaction, cation–π interaction[146]
Apple wood700-Cr(VI)5Electrostatic interactions, ion exchange, complexation[147]
Food waste300Food waste was mixed, oven-dried (105 °C), ground (10-mesh sieve), and stored in a sealed container at room temperatureCd(II), Pb(II)Cd 16.86, Pb 12.41Surface complexation, cation–π interaction, precipitation, electrostatic interaction[148]
Sewage sludge–coconut fiber600CopyrolysisCu(II), Zn(II), Ni(II), Cd(II)-Electrostatic interaction, complexation, π–π interaction, H-bond, pore filling, crystal lattices[149]
Modified pineapple pulp400-Cu(II)41.9Ion exchange, complexation, electrostatic attraction[150]

5.2. Adsorption of Transition Metals

Transition metals such as copper (Cu), zinc (Zn), nickel (Ni), and cobalt (Co) are indispensable in various industrial processes and essential for biological functions. However, their accumulation in the environment beyond permissible levels poses significant hazards. These metals are persistent pollutants that can contaminate soil, water, and air, damaging ecosystems and human health [151]. For instance, Cu and Zn are essential micronutrients but excess levels can disrupt aquatic life and soil microbial communities. Ni and Co, used in alloys and batteries, are toxic at elevated concentrations, impacting organisms’ growth and reproduction. Engineered BC has emerged as a promising solution due to its unique properties. Biochar, derived from biomass through pyrolysis, offers a highly porous structure with a large surface area and functional groups such as carboxyl, hydroxyl, and phenolic moieties [19]. Recent research has stated that BC production methods encompass various factors, including characterization, stability, and potential applications in promoting regenerative economic sustainability [30]. The resulting BC loaded with Fe-Mn binary oxides demonstrated a maximum adsorption capacity of 95.7 mg/g, which surpassed those loaded with individual oxides and pristine BC, showing potential for effective water remediation applications [152]. In exploring the pH-dependent adsorption behaviors of As(V) and Ni(II) using Fe-BC composites produced under N2 and CO2 environments, it was found that adsorption of As(V) increased significantly with the final pH from 3.0 to 5.8, reaching up to 20.1 mg/g for Fe-C-N2 and 18.7 mg/g for Fe-C-CO2, but decreased linearly thereafter. This trend is attributed to electrostatic repulsion between As(V) ions and the surface of magnetite and Fe as the pH exceeded 5.8 [153]. Continued research and innovation in BC technology are crucial for optimizing its effectiveness and sustainability in mitigating the adverse impacts of transition metal pollution on ecosystems and human well-being.

5.3. Adsorption of Rare Earth Metals

Rare earth elements (REEs), first discovered in 1794 by Finnish scholar John Gadolin, are defined by the IUPAC as comprising 15 lanthanides and two transition elements with similar chemical properties and they exhibit “lanthanide shrinkage” due to their decreasing atomic radius with increasing atomic numbers. Rare earth metals also can be found in secondary sources such as industrial byproducts (e.g., coal ash), electronic waste, and other residues [4]. Due to their unique chemical and physical properties, REEs are essential in various industries, including electronics, green energy, and agriculture. However, their environmental impacts and toxicological effects remain under-researched and poorly understood [154]. Studies have evaluated the profitability of various extraction methods from specific secondary sources. Understanding both the quality (concentration) and quantity of REEs in these sources is crucial; for instance, coal ash can contain concentrations ranging from 0.9% to 1.3% of total REEs. Techno-economic analysis (TEA) plays a pivotal role in assessing the economic feasibility of scaling up these extraction processes from lab to industrial scales [4]. There is another study that extensively reviews environmental considerations in rare earth mining, emphasizing techniques such as selective precipitation, solvent extraction, and ion exchange for thorium and uranium removal. These methods, particularly solvent extraction using tertiary and primary amines, offer effective separation capabilities. However, challenges remain in scaling up applications for commercial use and addressing similar adsorption behaviors between REEs and uranium [155].
According to the study [156], BC-FeNPs exhibited significant advantages in adsorbing rare earth elements (REEs) compared to FeNPs and BC alone. The adsorption efficiencies for REEs increased over time to a maximum, contrasting with the unchanged efficiency for Zn(II). At equilibrium, BC-FeNPs achieved a maximum adsorption efficiency of 88.4% for Lu(III), far surpassing the 8.7% efficiency observed for Zn(II). This demonstrated BC-FeNPs’ superior selectivity for REEs over Zn(II) in mine wastewater [156]. In another study [157], the structural analysis of BC before and after adsorption experiments revealed significant changes under scanning electron microscopy, utilizing an accelerating voltage of 10.0 kV and a working distance of 15.55 mm for enhanced resolution and depth. This study also employed infrared spectroscopy to examine surface functional groups of BC, highlighting key spectral features in the range of 450 to 4000 cm−1. Kinetic studies on the adsorption of rare earth elements using pseudo-second order and intraparticle diffusion models indicated that resonant vibratory mixing (RVM) enhanced adsorption efficiency more effectively than stirring, as evidenced by higher K2 values and better model fits.
Recent advancements in rare earth element (REEs) research emphasize their pivotal role in electronics, green energy, and agricultural technologies. Notably, innovative adsorption technologies, such as BC-supported iron nanoparticles (BC-FeNPs), have demonstrated superior REE adsorption capabilities compared to traditional methods, underpinned by sophisticated analytical techniques to elucidate structural changes and surface interactions during adsorption processes.

6. Recycling and Regeneration of Metal-Loaded Biochar

6.1. Chemical Regeneration Techniques

After a certain period of use, BC’s adsorption capacity diminishes as its active sites become saturated with heavy metals, necessitating regeneration. Additionally, heavy-metal-adsorbed BC, when discarded, particularly in acidic conditions, can release these metals as secondary pollutants, posing a long-term environmental hazard. Spent BC can often be regenerated through multiple adsorption-desorption cycles without significantly altering its physicochemical properties. Understanding the desorption process is crucial for evaluating BC’s adsorption efficiency and recycling potential. Effective desorption methods include thermal, solvent, chemical, microwave, and supercritical fluid treatments, with acids and complexing agents being particularly efficient [55,70].
Common chemical agents for regeneration include inorganic acids (HCl, HNO3, H2SO4), organic acids (formic acid, acetic acid), and alkaline solutions (NaOH, KOH). These agents, also known as stripping agents, are usually applied at lower concentrations (0.1–0.2 M) to avoid altering BC’s properties [152]. Additionally, complexing agents like ethylenediaminetetraacetic acid (EDTA) and salts such as NaCl, KNO3, NaNO3, and Ca (NO3)2 are employed for desorption [114,158]. Several studies have demonstrated the effectiveness of regenerating BC through various techniques. For instance, HCl has shown a desorption efficiency of 10–93%, outperforming HNO3 and NaOH due to the Cl ion’s similar ionic radius to metal ions [159]. Another study found EDTA-2Na to be the most effective desorbing agent for lead-loaded magnetic BC, with a desorption efficiency of 91.1%. Furthermore, microwave irradiation combined with chemical reagents like NaOH improved desorption rates, achieving up to 73% efficiency [160].
Poonam et al. [161] found that bagasse agricultural waste, modified through various physicochemical methods, was used to remove lead ions (Pb2+) from industrial wastewater. After adsorption, the saturated BC was treated with 0.1 M HNO3, achieving approximately 90% regeneration. Similarly, magnetic nanocomposites made from camel bone BC demonstrated efficient desorption of heavy metals like cadmium (Cd2+), cobalt (Co2+), and lead (Pb2+) using 0.01 M HCl [162]. Furthermore, nano zinc/bagasse BC showed an increased adsorption capacity for chromium (Cr6+) at 102.66 mg/g and maintained high heavy metal adsorption ability after six desorption cycles [163]. Additionally, water hyacinth BC exhibited significant regeneration potential following removing cadmium and lead when desorbing with 0.5 mol/L HCl [164]. The effectiveness of chemical-reagent-based regeneration methods depends on factors such as surface functional groups, pore structure, and changes in active sites. While effective, these methods can produce secondary contamination due to the potential leaching of sorbed contaminants. Among the various methods, acidic solutions are particularly effective at desorbing heavy metal ions [114].

6.2. Thermochemical Recycling Techniques

Regeneration using mineral acids, alkalis, and inorganic salts extends the BC’s service life. However, the adsorption capacity of regenerated BC decreases with each cycle, and waste BC disposal remains an issue. Additionally, wastewater generated during BC regeneration requires proper treatment [165]. Incineration is a prevalent thermochemical method for processing spent BC, but it does not regenerate BC. This process decreases both the mass and volume of the spent BC while facilitating thermal energy recovery. In contrast, pyrolysis is seen as a more effective method for the safe treatment and regeneration of exhausted BC [114]. Cui et al. [165] found that pyrolysis at 500 °C effectively removed Cd from the hyperaccumulator, producing more valuable BC than incineration-derived ash. The authors also investigated the pyrolysis of exhausted BC adsorbent at 300–900 °C, examining Cd behavior and the physicochemical properties and environmental applications of the regenerated BC [165]. Increasing pyrolysis temperatures greatly enhanced the vaporization of Cd, with almost no Cd in BC regenerated at 700–900 °C. The regenerated BC exhibited higher pH, ash content, and carbon content but lower hydrogen and oxygen contents compared to raw hydrochar. The toxicity and mobility of Cd were significantly reduced after pyrolysis and the regenerated BC showed a much higher Cd adsorption capacity (26.05–30.24 mg/g) compared to raw hydrochar (6.70 mg/g). Surface complexation with oxygen-containing functional groups was the primary Cd adsorption mechanism for hydrochar while precipitation with carbonates dominated Cd removal in regenerated BC. Yang et al. [166] stabilized spent rice straw BC loaded with cobalt and nickel through hydrothermal carbonization, creating a stable structure with a low leaching rate, which was reused as a catalyst for pollutant degradation. Pan et al. [167] used spent magnetic BC with adsorbed Cu2+ to improve peroxymonosulfate activation and norfloxacin degradation, enhancing the BC’s properties. Chen et al. [168] converted spent BC into heterostructured electrocatalysts, showing good performance in oxygen evolution reactions. The advanced oxidation process for removing organic pollutants can also utilize regenerated BC. Mer et al. [169] used BC loaded with lead and nickel to generate reactive hydroxyl radicals. Their findings suggest that the effectiveness of the advanced oxidation process relies on the structure of the carbon material and the type and amount of metal complexes it contains. Anton Zubrik et al. [170] found that microwave irradiation significantly enhances the desorption rate of As5+ from magnetic BC. These findings demonstrate that thermochemical processes are an effective technique for safely disposing of exhausted BC adsorbents and regenerating valuable BC.

6.3. Limitations in Recycling and Environmental Considerations

Disposing of or reusing heavy-metal-loaded BC presents a significant challenge, with no effective large-scale recycling strategy currently available. Stabilizing heavy metals on spent BC is difficult and improper treatment can result in secondary pollution [70]. Proper management and reuse are crucial to prevent secondary pollution when regenerating spent BC is not feasible. A resource utilization strategy can repurpose heavy-metal-laden spent BC, aligning with circular economy principles by transforming waste into valuable products. This approach creates economic opportunities and promotes sustainable resource use [169]. The reuse options for spent BC, as described in the subsequent section, include applications in energy storage devices, supercapacitors, catalysts, and construction materials. These applications align with circular economy principles and contribute to the reduction of carbon footprints by capturing CO2 from the atmosphere [114]. Therefore, developing convenient and economical recycling technologies reduces the production costs of adsorbents and minimizes waste generated from metal-loaded adsorbents. However, challenges like loss and structural degradation during recycling limit their effective reuse over multiple cycles in laboratory-scale studies [171]. Incineration offers waste-to-energy benefits, reducing the mass and volume of spent adsorbents while recovering thermal energy but it is not capable of regenerating BC. Landfilling, although a last option, may require pretreatments to reduce heavy metal content to comply with local regulations or USEPA standards for hazardous waste [114]. Thus, advancing technologies for the recovery, regeneration, and recycling of adsorbents is crucial for scaling up adsorption engineering applications.

7. Applications of Metal-Loaded Biochar

7.1. Soil Amendment and Fertilization

The application of metal-loaded BC in soil amendment and fertilization offers promising benefits for enhancing soil properties and promoting agricultural productivity. BC, a carbon-rich material from biomass pyrolysis, is widely recognized for enhancing soil structure, nutrient retention, and plant growth, serving primarily as a soil amendment to improve agricultural conditions [172]. When loaded with metal ions through adsorption processes, BC can also address soil contamination issues and provide additional nutrients for crops. Metal-loaded BC can be used as a soil amendment to address deficiencies in essential trace elements and improve soil fertility. For example, iron-loaded BC can enhance soil fertility and nutrient absorption efficiency, effectively delivering nutrients to plants [173]. The introduction of metal-loaded BC into soil can enhance nutrient availability, improve soil pH, and increase soil cation exchange capacity, thereby fostering better crop yields.
Furthermore, BC amended with heavy metals can aid in the remediation of contaminated soils. For instance, BC loaded with Pb and Cd has been employed to immobilize these toxic metals in polluted soils, reducing their bioavailability and mitigating potential risks to plant and human health [174]. This immobilization process occurs through various mechanisms, including adsorption and chemical stabilization, thereby decreasing mobility and toxicity. Studies have demonstrated that metal-loaded BC can also contribute to improved soil fertility. For example, BC loaded with Cu and Zn enhanced bean production and micronutrient uptake, particularly Zn, outperforming traditional fertilizers [175]. BC contains a significant concentration of trace elements such as Fe, Cu, B, Zn, and Mn, which can enhance plant nutrient uptake and contribute to disease resistance mechanisms [176]. Similarly, the use of metal-loaded BC has been reported to improve soil quality and water-holding capacity, which can be particularly beneficial in drought-prone areas [177].

7.2. Catalytic Applications

Biochar has gained significant attention as a versatile support for metal-loaded catalysts, presenting numerous innovative applications across environmental, energy, and chemical sectors. The high surface area, porosity, thermal stability, and diverse functional groups of BC make it an excellent candidate for dispersing and stabilizing metal nanoparticles, enhancing their catalytic efficiency [178]. In environmental remediation, metal-loaded BC catalysts are effectively used in water treatment processes to adsorb and degrade pollutants such as heavy metals, textile dyes, and organic contaminants [97,179,180,181,182]. For energy production, these catalysts play a critical role in biofuel synthesis by improving the conversion efficiency of biomass into valuable fuels and chemicals through processes like pyrolysis, gasification, and hydrothermal liquefaction. Additionally, metal-loaded BC catalysts facilitate the production of hydrogen via catalytic reforming of bio-oil or methane, contributing to cleaner energy solutions [183,184]. In chemical synthesis, they are employed to enhance the production of fine chemicals and pharmaceuticals due to their high catalytic activity and selectivity. For instance, alkaline earth metals and transition metals like iron (Fe), nickel (Ni), cobalt (Co), and copper (Cu) loaded BC are favored for their cost-effectiveness and catalytic properties. Bimetallic and multimetallic combinations of BC further improve catalytic performance through synergistic effects. Preparation methods such as impregnation, coprecipitation, and the sol-gel process ensure uniform metal distribution and enhanced interaction with the BC matrix. Overall, metal-loaded BC catalysts represent a sustainable and efficient approach to addressing contemporary industrial and environmental challenges.

7.2.1. Application for Wastewater Treatment

Metal-loaded BC catalysts have emerged as a highly effective solution for wastewater treatment, leveraging the unique properties of BC to enhance the removal of inorganic and organic contaminants. These catalysts excel in adsorbing and degrading a wide range of pollutants, including heavy metals, dyes, pharmaceuticals, and organic contaminants, through processes like adsorption, reduction, and oxidation [185,186,187,188]. For example, iron-loaded BC can effectively remove arsenic and lead from wastewater [189,190] while copper-loaded BC excels in breaking down complex organic compounds [185]. Additionally, the catalytic properties of metal-loaded BC enable advanced oxidation processes (AOPs), such as Fenton-like reactions, which generate reactive oxygen species to degrade persistent pollutants [180,182]. The use of metal-loaded BC in wastewater treatment not only improves water quality but also offers a sustainable approach by utilizing waste biomass for BC production and reducing reliance on traditional, more hazardous chemical treatments. This innovative application highlights the potential of metal-loaded BC catalysts to address pressing environmental challenges and enhance the efficiency of wastewater purification systems.
The list of recent studies on metal-loaded BC is presented in Table 3 for the removal and degradation of inorganic and organic pollutants from wastewater. Ouyang et al. and Wu, Jia et al. investigated the degradation of tetracycline (TC) and bisphenol A (BPA) using MoO2-loaded BC as a cocatalyst to enhance Fe3+/Fe2+ cycling, achieving removal efficiencies of 90.9% for TC and over 96% for BPA within 60 min [188,191]. In both studies, Mo- and iron-loaded BC were prepared through impregnation with iron salts, pyrolysis at 600–800 °C, and subsequent autoclaving at 180 °C with ammonium molybdate. However, these multistage processes could increase operational costs. In another separate study, Coloaded BC was prepared from cherry kernels for efficient removal of BPA with the aid of peroxymonosulfate (PMS) and persulfate (PDS) [186]. The optimal concentrations of PMS and PDS were found to be 0.25 g L−1 and 1.0 g L−1, respectively, when using Coloaded BC. Increasing the temperature from 15 to 45 °C improved the BPA removal efficiency, which was highest at pH 7 and lowest at pH 3. The removal efficiency of BPA remained above 90% after five cycles, indicating economic viability. Transition metals, carbon materials, heat, UV radiation, and ultrasound can activate PDS, while PMS activation by Fe2+ for organic contaminant removal can be enhanced by the active MoO2 surface [181,192]. A recent study highlighted the superior performance of Cu-loaded BC (BC-Cu, 1:4) in activating PMS [180]. The prepared BC-Cu composite achieved a remarkable 99.99% removal rate of ciprofloxacin, with an initial concentration of 20 mg·L−1 within 40 min. Additionally, the BC-Cu sample demonstrated significant activity across a wide pH range (4–12) and exhibited an “anti-anion interference” capability [180]. Another unconventional modified BC was prepared in recent years by autoclaving of corn stalks in sulfuric acid followed by copyrolysis with KOH (1:1 with the hydro char) and finally wet impregnation of the BC with FeCl3·6H2O at room temperature in an inert environment. The produced hydrophilic BC had excellent removal efficiency of Pb2+ (99.8%) due to the synergetic effect with the zero-valent iron (Fe0) [193].
Table 3. Recent studies of the application of metal-loaded biochar for wastewater treatment.
Table 3. Recent studies of the application of metal-loaded biochar for wastewater treatment.
Raw MaterialsMetal CompoundsProcessRemarksReferences
CalamusMoO2-enhanced Fe2+/Fe3+Impregnation and followed by pyrolysis, 600 °CDegradation of tetracycline, 90.9%[188]
Rice strawMoO2-enhanced Fe2+/Fe3+Impregnation and followed by pyrolysis, 800 °CDegradation of BPA, >96%[191]
Cherry kernelCoCl2·6H2OImpregnation and followed by pyrolysis, 800 °CDegradation of BPA, ~100%[186]
Rice strawCuCl2·2H2OImpregnation and followed by pyrolysis, 800 °CDegradation of ciprofloxacin, ~100%[180]
Rice strawFeCl3·6H2O/LaCl3·7H2OPyrolysis, 450 °CAdsorption of phosphorus, 52 mg P/g[182]
Corn stalksKOH, FeCl3·6H2OAutoclave at 200 °C, Copyrolysis (800 °C) with KOH and impregnation in FeCl3·6H2O solution480.9 mg g−1 of Pb2+adsorption capacity[193]
Sewage sludgeZero valent ironCopyrolysis 500 °C83.4% degradation of sulfamethoxazole[194]
Municipal sludgeFeCl3·6H2OImpregnation and followed by pyrolysis, 400 °C99.8% degradation of thiamethoxam[195]
Moreover, iron-loaded BC has been studied for Fenton-like catalytic reactions for the degradation of organic pollutants in wastewater [194,195]. The Fenton reaction, which involves the combination of Fe2+ and H2O2, has attracted increased attention due to its potent free radical oxidation capabilities and environmental benefits. Iron-loaded magnetic biochar has been identified as an effective catalyst for activating hydrogen peroxide (H2O2) to generate hydroxyl radicals (OH). Both radical quenching experiments and electron paramagnetic resonance exposure have confirmed the generation of OH and its significant role in the oxidative degradation of thiamethoxam [195]. Previous research indicates that an acidic pH optimizes Fenton activity, with pH levels above 4 hindering the Fe2+/Fe3+ cycle due to the formation of inactive iron complexes. Additionally, increasing the catalytic dosage to 1.5 g/L and raising the adsorption batch temperature to 45 °C have demonstrated superior degradation capabilities for sulfamethoxazole [194].
Phosphorus (P) removal from wastewater has also been a focus of recent research. It was found that raw BC has a limited capacity for P removal, necessitating further modification. Previous studies identified La(OH)3 as an effective compound for P removal, achieving nearly 100% removal efficiency [182,196]. Building on this, Sun, Zhang, et al. developed a novel rice-straw-based BC, which was subsequently mixed with FeCl3·6H2O and LaCl3·7H2O salt solutions to create a composite granule with montmorillonite, demonstrating efficient phosphorus removal [182].

7.2.2. Applications for Catalytic Bio-Oil Production

Transitional metal ion-loaded BC has demonstrated significant potential in enhancing the production of bio-oil and biochemicals, offering an innovative and sustainable approach to biomass conversion [197,198,199,200]. These metals are well regarded for their catalytic properties in various thermochemical processes. When loaded onto BC, these metal ions can significantly improve the efficiency of biomass conversion techniques like pyrolysis, gasification, and hydrothermal liquefaction. For instance, nickel-loaded BC serves as an effective catalyst in the pyrolysis process, enhancing the yield and quality of bio-oil by promoting the breakdown of complex organic polymers such as lignin and cellulose, etc. into simpler, more valuable hydrocarbons [198,199]. Similarly, iron and Mn-loaded BC catalysts facilitate the gasification process, aiding in the production of syngas, which can be further processed into a variety of biochemicals and fuels [201,202].
The presence of these metal ions not only accelerates the reaction rates but also aids in the selective formation of desired products, reducing the formation of unwanted by-products such as tar. The synergistic effect between BC and transitional metals enhances catalytic stability and longevity, making these catalysts more effective and economically viable for large-scale applications. A few recent studies on the production of biofuel by using metal-loaded BC as a catalyst are listed in Table 4.
The superior catalytic activity and notable recyclability of iron enhance its industrial application value compared to other metals. During the thermochemical conversion of biomass, the produced bio-oil is enriched with phenolic/aromatic compounds. Lignin has been identified as the primary source of phenolic compounds in bio-oil [203]. Cellulose undergoes dehydration, decarboxylation, decarboxylation, and cyclization reactions, leading to the formation of cyclic oxygen-containing compounds such as furan derivatives, which subsequently transform into phenolic compounds [204,205]. Specifically, anhydro sugars derived from cellulose are converted into furans through dehydration and structural rearrangements. Additionally, C5 compounds are produced via dehydration and decarboxylation processes that release gases like carbon monoxide, methane, and water. These reactions occur at the catalytic sites of BC through mechanisms such as Diels–Alder reactions, decarboxylation, and oligomerization, eventually reforming into phenols [203]. Hemicellulose, upon aromatization, can also be converted into phenolic compounds [206]. In this context, iron-loaded cellulose BC effectively facilitates the transformation of aldehydes and ketones present in corncob bio-oil, significantly enhancing the phenolic compound content [200,203]. Nickel-based catalysts are extensively employed in the hydrogenolysis of lignin and its model compounds due to their exceptional hydrogenation capabilities and cost-effectiveness [197,198,199]. These catalysts, often supported on materials like BC, can selectively cleave the C-O linkages in β-O-4 and α-O-4 bonds within lignin and model compounds such as diphenyl ether, resulting in the formation of phenolic compounds [198]. Nickel-supported activated carbon (AC) has demonstrated significant efficacy in breaking the C-O bonds of diaryl ethers, yielding high concentrations of aromatic and phenolic compounds. In a previous study, BC derived from various carbon sources, including cellulose, lignin, and hemicellulose, was used to support a Ni catalyst for the catalytic hydrogenolysis of lignin model compounds (e.g., diphenyl ether) [199]. The study indicated that Ni/AC derived from cellulose exhibited the highest catalytic activity in converting lignin-derived compounds into aromatic products. The uniform pore sizes and structural characteristics of the cellulose-derived BC were identified as critical factors enhancing the dispersion of Ni while the sulfur-free nature of cellulose contributed to increased catalytic activity of the Ni/AC [199].
The industrial synthesis of biodiesel typically involves the homogeneous transesterification of vegetable oils and animal fats, catalyzed by strong acids and bases such as KOH, NaOH, and H2SO4 [207]. However, these strong acids and bases are generally avoided due to their corrosive nature. As an alternative, CaO has been employed in biodiesel production, owing to its basicity and low cost. Nevertheless, the stability of CaO can diminish due to its transformation into Ca(OH)2 and CaCO3, which reduces catalytic activity [207]. In this context, BC can serve as an effective carrier for CaO, helping to maintain catalyst activity during the transesterification reaction. Recent studies have developed efficient Ca-BC composite catalysts using rice husk and municipal sludge as carriers. These materials are rich in silica, which can form Ca-O-Si bonds on the BC surface, enhancing catalyst stability [207,208]. Postactivation with KOH significantly increases the porosity of BC, resulting in a high biodiesel yield of 93–94% due to the catalyst’s high basicity. Comparisons between the newly synthesized catalyst and pure CaO revealed that rice husk BC is an effective carrier, with the Ca-O-Si bond formation contributing to both catalytic activity and stability [209].
Table 4. Recent studies of the application of metal-loaded biochar for bio-oil production.
Table 4. Recent studies of the application of metal-loaded biochar for bio-oil production.
Feedstock for BCProcess for BC CatalystMetal Ion SourcesFeedstock for BiofuelProcess for BiofuelRemarksReferences
Rice huskMicrowave-assisted catalytic pyrolysis, 550 °CFe(NO3)3, impregnation and calcined with BCCorn cobMW-assisted torrefactionIncreased phenol compounds in bio-oil up to 60%[200]
Cellulose powderPyrolysis, 600 °CNi(NO3)2·6H2O, impregnation and calcined with cellulose powder BC (CPB)Lignin model compoundsHydrogenolysis, 140 °CIncreased yield of phenol (12.5%) and cyclohexanol (80.7%)[199]
Peanut shellMicrowave-assisted Pyrolysis, 550 °CFeCl2·4H2O and FeCl3 (1:1 ratio), impregnation and calcined, 450 °CPeanut shellPyrolysisTotal bio-oil 24.3% where 27% aliphatic and 18% aromatic hydrocarbon selectivity[201]
Rice huskPyrolysis, 700 °CCaO (30 wt%), impregnation and calcined (700 °C) with rice husk BCPalm oilTransesterification reaction, 65 °CBiodiesel yields up to 93.4%[209]
Municipal sludgePyrolysis, 800 °CCaO (20 wt%), impregnation and calcined (700 °C) with the BCPalm oilTransesterification reaction, 65 °CBiodiesel yields up to 93.8%[207]
Peanut shellPyrolysis, 700 °CFeCl3·4H2O, MnCl2·4H2O, NaCl; impregnation and calcined with BCPeanut shellPyrolysis, 440–660 °C8% phenolic selectivity over Fe3+[202]
Pine tree needlesImpregnation with ZnCl2 (1:1 biomass to ZnCl2 weight ratio) and Pyrolysis, 700 °CNi(NO3)2·6H2O, Impregnation and calcined with the modified BC, 500 °CPine tree needlesPyrolysis, 550 °CIncreased the aromatic selectivity to ~36%[197]

7.3. Electrochemical Applications

By increasing the porosity, surface area, graphitization, or doping with heteroatoms, BC can be tuned into a suitable for energy storage applications in supercapacitors, batteries, and hydrogen storage [210]. BC derived from manure and sewage sludge, after microwave oxidation to reduce carbon content and increase oxygen content, showed enhanced stability in specific capacitance, power density, and charge–discharge capacities when used as supercapacitor materials loaded with Ni2+, forming NiO and NiOOH [211]. Spent BC has been used to create supercapacitor electrodes, addressing the demand for green energy. For instance, cobalt-doped BC exhibited remarkable electrochemical properties, making it suitable for energy storage [212]. Deping Li et al. [213] converted waste Ni-laden BC into pseudocapacitive materials via hydrothermal treatment. The optimized temperature (90–180 °C) and Ni content (100 and 500 mg/g) revealed that nickel can amplify the graphite layer and improve the stability of BC-based pseudocapacitive material during hydrothermal treatment. Rawat et al. [210] reported that optimizing BC for energy storage in batteries and supercapacitors requires a careful selection of additives, conducting polymers, solvents, and electrolytes to enhance overall device performance and justify costs. Comprehensive performance evaluation and testing are necessary to develop better alternatives to traditional batteries and accelerate research and development. During this review, we were not able to find lots of reports in this area but with the rapid increase in the demand for energy storage devices, we believe this is an interesting area that is worth exploring.

8. Future Research Directions

Most studies use pristine BC for heavy metal removal from groundwater and wastewater sources. However, further research is needed to evaluate the application of engineered BC under diverse conditions, accounting for all potential influencing factors in water and wastewater. Future studies should focus on enhancing BC adsorption capacity by incorporating functional groups and novel substances to improve its physical and chemical properties for targeted heavy metal removal. Current research on biomass-based BC often uses synthetic wastewater containing a limited number of metals. Investigating real wastewater with multiple metals is crucial to understanding adsorption interactions and competitive effects. Additionally, exploring BC’s efficacy in treating acid mine drainage and combining it with other treatments could optimize heavy metal remediation [30].
BC-based heavy metal removal has predominantly been investigated at the laboratory scale. While many studies have demonstrated effective removal, the transition to the industrial application has been limited due to a lack of engagement from entrepreneurs and industry personnel. Collaborating with industry and large-scale pilot trials is essential to advance BC as a cost-effective and efficient adsorbent to assess its stability and performance under real-world conditions. Sustainable adsorption technologies require low-cost adsorbents. Comparative cost-benefit analyses between biomass-based BC and conventional adsorbents along with comprehensive life cycle assessments are necessary to evaluate environmental sustainability. Employing artificial neural networks can model and predict BC performance under varying conditions and hopefully enhance adsorption efficiency [18,21]. Specifically, artificial neural networks (ANNs) are powerful tools for modeling and predicting BC performance by analyzing complex data patterns. They can optimize BC properties for enhanced adsorption efficiency by training on historical data to understand relationships between BC characteristics and performance under various conditions. Once trained, ANNs can predict how different conditions affect performance, facilitating the development of more efficient BCs and guiding improvements based on complex, nonlinear interactions that are challenging to model with traditional methods.
Furthermore, challenges related to the regeneration, reuse, and proper disposal of spent BC must be addressed. There is limited research on regenerating spent BC to extend its lifespan and reduce operational costs. Future studies should focus on developing efficient, low-cost, and environmentally friendly regeneration techniques. Spent BC with heavy metals cannot be directly disposed of due to the risk of secondary pollution. There is limited information on the toxic effects of disposing of metal-impregnated BC, necessitating the development of effective stabilization and desorption methods. Eco-friendly decontamination methods using natural micro-organisms or plants should be developed [70,114].
Additionally, novel methods for reusing spent BC for energy recovery or converting it into value-added and energy storage should be explored to support circular economy principles. Converting spent BC into electrode materials for supercapacitors and catalysts supports sustainable energy solutions. This highlights the need for further exploration of the relationship between biomass waste precursors and carbonaceous electrochemical materials [114].

9. Conclusions

The use of engineered BC derived from waste biomass for heavy metal adsorption offers significant environmental benefits, such as waste reduction and improved sustainability. This review examines the current progress in developing engineered BC for effective heavy metal remediation and the repurposed applications of metal-loaded BC. However, challenges persist in terms of scalability, regeneration, disposal, and system compatibility. Future research should focus on green functionalization methods and evaluate their environmental and economic feasibility. Overcoming these challenges will allow engineered BC to play a more significant role in creating a cleaner and more sustainable future.

Author Contributions

M.H.: conceptualization; methodology; writing—original draft; writing—review and editing. S.C.: writing—original draft; writing—review and editing. X.L.: writing—original draft; writing—review and editing. S.S.: writing—original draft; writing—review and editing. K.K.: conceptualization; supervision; validation; funding acquisition. J.K.: supervision; validation; funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thank the NSERC (Natural Sciences and Engineering Research Council of Canada) for supporting this study through the Discovery Grant (RGPIN-2023-03289: Kang Kang) and Alliance International Catalyst Grants (Janusz Kozinski and Kang Kang).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Briffa, J.; Sinagra, E.; Blundell, R. Heavy metal pollution in the environment and their toxicological effects on humans. Heliyon 2020, 6, e04691. [Google Scholar] [CrossRef] [PubMed]
  2. Emenike, E.C.; Iwuozor, K.O.; Anidiobi, S.U. Heavy metal pollution in aquaculture: Sources, impacts and mitigation techniques. Biol. Trace Elem. Res. 2021, 200, 4476–4492. [Google Scholar] [CrossRef] [PubMed]
  3. Yaashikaa, P.; Kumar, P.S.; Jeevanantham, S.; Saravanan, R. A review on bioremediation approach for heavy metal detoxification and accumulation in plants. Environ. Pollut. 2022, 301, 119035. [Google Scholar] [CrossRef] [PubMed]
  4. Gaustad, G.; Williams, E.; Leader, A. Rare earth metals from secondary sources: Review of potential supply from waste and byproducts. Resour. Conserv. Recycl. 2021, 167, 105213. [Google Scholar] [CrossRef]
  5. Duan, C.; Ma, T.; Wang, J.; Zhou, Y. Removal of heavy metals from aqueous solution using carbon-based adsorbents: A review. J. Water Process Eng. 2020, 37, 101339. [Google Scholar] [CrossRef]
  6. Patra, B.R.; Mukherjee, A.; Nanda, S.; Dalai, A.K. Biochar production, activation and adsorptive applications: A review. Environ. Chem. Lett. 2021, 19, 2237–2259. [Google Scholar] [CrossRef]
  7. Li, Y.; Gupta, R.; Zhang, Q.; You, S. Review of biochar production via crop residue pyrolysis: Development and perspectives. Bioresour. Technol. 2023, 369, 128423. [Google Scholar] [CrossRef]
  8. Tan, S.; Zhou, G.; Yang, Q.; Ge, S.; Liu, J.; Cheng, Y.W.; Yek, P.N.Y.; Mahari, W.A.W.; Kong, S.H.; Chang, J.-S. Utilization of current pyrolysis technology to convert biomass and manure waste into biochar for soil remediation: A review. Sci. Total Environ. 2023, 864, 160990. [Google Scholar] [CrossRef]
  9. Liao, W.; Zhang, X.; Ke, S.; Shao, J.; Yang, H.; Zhang, S.; Chen, H. Effect of different biomass species and pyrolysis temperatures on heavy metal adsorption, stability and economy of biochar. Ind. Crops Prod. 2022, 186, 115238. [Google Scholar] [CrossRef]
  10. Panwar, N.L.; Pawar, A. Influence of activation conditions on the physicochemical properties of activated biochar: A review. Biomass Convers. Biorefinery 2020, 12, 925–947. [Google Scholar] [CrossRef]
  11. Anto, S.; Sudhakar, M.; Ahamed, T.S.; Samuel, M.S.; Mathimani, T.; Brindhadevi, K.; Pugazhendhi, A. Activation strategies for biochar to use as an efficient catalyst in various applications. Fuel 2021, 285, 119205. [Google Scholar] [CrossRef]
  12. Xu, Z.; He, M.; Xu, X.; Cao, X.; Tsang, D.C. Impacts of different activation processes on the carbon stability of biochar for oxidation resistance. Bioresour. Technol. 2021, 338, 125555. [Google Scholar] [CrossRef] [PubMed]
  13. Pradhan, S.; Parthasarathy, P.; Mackey, H.R.; Al-Ansari, T.; McKay, G. Food waste biochar: A sustainable solution for agriculture application and soil–water remediation. Carbon Res. 2024, 3, 41. [Google Scholar] [CrossRef]
  14. Afraz, M.; Muhammad, F.; Nisar, J.; Shah, A.; Munir, S.; Ali, G.; Ahmad, A. Production of value added products from biomass waste by pyrolysis: An updated review. Waste Manag. Bull. 2024, 1, 30–40. [Google Scholar] [CrossRef]
  15. Islam, M.S.; Kwak, J.-H.; Nzediegwu, C.; Wang, S.; Palansuriya, K.; Kwon, E.E.; Naeth, M.A.; El-Din, M.G.; Ok, Y.S.; Chang, S.X. Biochar heavy metal removal in aqueous solution depends on feedstock type and pyrolysis purging gas. Environ. Pollut. 2021, 281, 117094. [Google Scholar] [CrossRef]
  16. Hashem, M.A.; Hasan, M.; Momen, M.A.; Payel, S.; Nur-A-Tomal, M.S. Water hyacinth biochar for trivalent chromium adsorption from tannery wastewater. Environ. Sustain. Indic. 2020, 5, 100022. [Google Scholar] [CrossRef]
  17. Singh, A.; Sharma, R.; Pant, D.; Malaviya, P. Engineered algal biochar for contaminant remediation and electrochemical applications. Sci. Total Environ. 2021, 774, 145676. [Google Scholar] [CrossRef]
  18. Adeniyi, A.G.; Iwuozor, K.O.; Emenike, E.C.; Ajala, O.J.; Ogunniyi, S.; Muritala, K.B. Thermochemical co-conversion of biomass-plastic waste to biochar: A review. Green Chem. Eng. 2024, 5, 31–49. [Google Scholar] [CrossRef]
  19. Amalina, F.; Razak, A.S.A.; Krishnan, S.; Sulaiman, H.; Zularisam, A.W.; Nasrullah, M. Biochar production techniques utilizing biomass waste-derived materials and environmental applications—A review. J. Hazard. Mater. Adv. 2022, 7, 100134. [Google Scholar] [CrossRef]
  20. 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]
  21. Kumar, P.S.; Gayathri, R.; Rathi, B.S. A review on adsorptive separation of toxic metals from aquatic system using biochar produced from agro-waste. Chemosphere 2021, 285, 131438. [Google Scholar] [CrossRef]
  22. Kumar, M.; Xiong, X.; Wan, Z.; Sun, Y.; Tsang, D.C.W.; Gupta, J.; Gao, B.; Cao, X.; Tang, J.; Ok, Y.S. Ball milling as a mechanochemical technology for fabrication of novel biochar nanomaterials. Bioresour. Technol. 2020, 312, 123613. [Google Scholar] [CrossRef] [PubMed]
  23. Zuhara, S.; Pradhan, S.; McKay, G. Pyrolysis of biosolids with waste cardboard: Effect of operating parameters, feedstock size and blending ratio. Int. J. Environ. Sci. Technol. 2024, 21, 617–630. [Google Scholar] [CrossRef]
  24. Masud, M.A.A.; Shin, W.S.; Sarker, A.; Septian, A.; Das, K.; Deepo, D.M.; Iqbal, M.A.; Islam, A.R.M.T.; Malafaia, G. A critical review of sustainable application of biochar for green remediation: Research uncertainty and future directions. Sci. Total Environ. 2023, 904, 166813. [Google Scholar] [CrossRef]
  25. Yao, Y.; Gao, B.; Fang, J.; Zhang, M.; Chen, H.; Zhou, Y.; Creamer, A.E.; Sun, Y.; Yang, L. Characterization and environmental applications of clay–biochar composites. Chem. Eng. J. 2014, 242, 136–143. [Google Scholar] [CrossRef]
  26. Zhao, L.; Cao, X.; Mašek, O.; Zimmerman, A. Heterogeneity of biochar properties as a function of feedstock sources and production temperatures. J. Hazard. Mater. 2013, 256–257, 1–9. [Google Scholar] [CrossRef]
  27. Leng, L.; Xiong, Q.; Yang, L.; Li, H.; Zhou, Y.; Zhang, W.; Jiang, S.; Li, H.; Huang, H. An overview on engineering the surface area and porosity of biochar. Sci. Total Environ. 2021, 763, 144204. [Google Scholar] [CrossRef]
  28. Chen, Y.-D.; Liu, F.; Ren, N.-Q.; Ho, S.-H. Revolutions in algal biochar for different applications: State-of-the-art techniques and future scenarios. Chin. Chem. Lett. 2020, 31, 2591–2602. [Google Scholar] [CrossRef]
  29. Sun, Y.; Gao, B.; Yao, Y.; Fang, J.; Zhang, M.; Zhou, Y.; Chen, H.; Yang, L. Effects of feedstock type, production method, and pyrolysis temperature on biochar and hydrochar properties. Chem. Eng. J. 2014, 240, 574–578. [Google Scholar] [CrossRef]
  30. Amalina, F.; Razak, A.S.A.; Krishnan, S.; Zularisam, A.W.; Nasrullah, M. A comprehensive assessment of the method for producing biochar, its characterization, stability, and potential applications in regenerative economic sustainability—A review. Clean. Mater. 2022, 3, 100045. [Google Scholar] [CrossRef]
  31. Vijayaraghavan, K. Recent advancements in biochar preparation, feedstocks, modification, characterization and future applications. Environ. Technol. Rev. 2019, 8, 47–64. [Google Scholar] [CrossRef]
  32. Chandra, S.; Bhattacharya, J. Influence of temperature and duration of pyrolysis on the property heterogeneity of rice straw biochar and optimization of pyrolysis conditions for its application in soils. J. Clean. Prod. 2019, 215, 1123–1139. [Google Scholar] [CrossRef]
  33. Gale, M.; Nguyen, T.; Moreno, M.; Gilliard-AbdulAziz, K.L. Physiochemical Properties of Biochar and Activated Carbon from Biomass Residue: Influence of Process Conditions to Adsorbent Properties. ACS Omega 2021, 6, 10224–10233. [Google Scholar] [CrossRef] [PubMed]
  34. Jawad, A.H.; Razuan, R.; Appaturi, J.N.; Wilson, L.D. Adsorption and mechanism study for methylene blue dye removal with carbonized watermelon (Citrullus lanatus) rind prepared via one-step liquid phase H2SO4 activation. Surf. Interfaces 2019, 16, 76–84. [Google Scholar] [CrossRef]
  35. Mariyam, S.; Alherbawi, M.; Pradhan, S.; Al-Ansari, T.; McKay, G. Biochar yield prediction using response surface methodology: Effect of fixed carbon and pyrolysis operating conditions. Biomass Convers. Biorefinery 2023, 1–14. [Google Scholar] [CrossRef]
  36. Pradhan, S.; Abdelaal, A.H.; Mroue, K.; Al-Ansari, T.; Mackey, H.R.; McKay, G. Biochar from vegetable wastes: Agro-environmental characterization. Biochar 2020, 2, 439–453. [Google Scholar] [CrossRef]
  37. Carolin, C.F.; Kumar, P.S.; Saravanan, A.; Joshiba, G.J.; Naushad, M. Efficient techniques for the removal of toxic heavy metals from aquatic environment: A review. J. Environ. Chem. Eng. 2017, 5, 2782–2799. [Google Scholar] [CrossRef]
  38. Jiang, T.; Wang, B.; Gao, B.; Cheng, N.; Feng, Q.; Chen, M.; Wang, S. Degradation of organic pollutants from water by biochar-assisted advanced oxidation processes: Mechanisms and applications. J. Hazard. Mater. 2023, 442, 130075. [Google Scholar] [CrossRef]
  39. Wang, J.; Wang, S. Preparation, modification and environmental application of biochar: A review. J. Clean. Prod. 2019, 227, 1002–1022. [Google Scholar] [CrossRef]
  40. Kazemi Shariat Panahi, H.; Dehhaghi, M.; Ok, Y.S.; Nizami, A.-S.; Khoshnevisan, B.; Mussatto, S.I.; Aghbashlo, M.; Tabatabaei, M.; Lam, S.S. A comprehensive review of engineered biochar: Production, characteristics, and environmental applications. J. Clean. Prod. 2020, 270, 122462. [Google Scholar] [CrossRef]
  41. Brown, A.E.; Adams, J.M.M.; Grasham, O.R.; Camargo-Valero, M.A.; Ross, A.B. An Assessment of Different Integration Strategies of Hydrothermal Carbonisation and Anaerobic Digestion of Water Hyacinth. Energies 2020, 13, 5983. [Google Scholar] [CrossRef]
  42. Chen, W.-H.; Hoang, A.T.; Nižetić, S.; Pandey, A.; Cheng, C.K.; Luque, R.; Ong, H.C.; Thomas, S.; Nguyen, X.P. Biomass-derived biochar: From production to application in removing heavy metal-contaminated water. Process Saf. Environ. Prot. 2022, 160, 704–733. [Google Scholar] [CrossRef]
  43. Ponnusamy, V.K.; Nagappan, S.; Bhosale, R.R.; Lay, C.-H.; Duc Nguyen, D.; Pugazhendhi, A.; Chang, S.W.; Kumar, G. Review on sustainable production of biochar through hydrothermal liquefaction: Physico-chemical properties and applications. Bioresour. Technol. 2020, 310, 123414. [Google Scholar] [CrossRef]
  44. Sakhiya, A.K.; Anand, A.; Kaushal, P. Production, activation, and applications of biochar in recent times. Biochar 2020, 2, 253–285. [Google Scholar] [CrossRef]
  45. Hassan, M.; Liu, Y.; Naidu, R.; Parikh, S.J.; Du, J.; Qi, F.; Willett, I.R. Influences of feedstock sources and pyrolysis temperature on the properties of biochar and functionality as adsorbents: A meta-analysis. Sci. Total Environ. 2020, 744, 140714. [Google Scholar] [CrossRef]
  46. Medeiros, D.C.C.d.S.; Nzediegwu, C.; Benally, C.; Messele, S.A.; Kwak, J.-H.; Naeth, M.A.; Ok, Y.S.; Chang, S.X.; Gamal El-Din, M. Pristine and engineered biochar for the removal of contaminants co-existing in several types of industrial wastewaters: A critical review. Sci. Total Environ. 2022, 809, 151120. [Google Scholar] [CrossRef]
  47. Qiu, B.; Shao, Q.; Shi, J.; Yang, C.; Chu, H. Application of biochar for the adsorption of organic pollutants from wastewater: Modification strategies, mechanisms and challenges. Sep. Purif. Technol. 2022, 300, 121925. [Google Scholar] [CrossRef]
  48. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Chen, M. Progress in the preparation and application of modified biochar for improved contaminant removal from water and wastewater. Bioresour. Technol. 2016, 214, 836–851. [Google Scholar] [CrossRef]
  49. Kim, Y.; Oh, J.-I.; Vithanage, M.; Park, Y.-K.; Lee, J.; Kwon, E.E. Modification of biochar properties using CO2. Chem. Eng. J. 2019, 372, 383–389. [Google Scholar] [CrossRef]
  50. Wan, S.; Qiu, L.; Tang, G.; Chen, W.; Li, Y.; Gao, B.; He, F. Ultrafast sequestration of cadmium and lead from water by manganese oxide supported on a macro-mesoporous biochar. Chem. Eng. J. 2020, 387, 124095. [Google Scholar] [CrossRef]
  51. Lyu, H.; Gao, B.; He, F.; Zimmerman, A.R.; Ding, C.; Huang, H.; Tang, J. Effects of ball milling on the physicochemical and sorptive properties of biochar: Experimental observations and governing mechanisms. Environ. Pollut. 2018, 233, 54–63. [Google Scholar] [CrossRef] [PubMed]
  52. Ye, S.; Cheng, M.; Zeng, G.; Tan, X.; Wu, H.; Liang, J.; Shen, M.; Song, B.; Liu, J.; Yang, H.; et al. Insights into catalytic removal and separation of attached metals from natural-aged microplastics by magnetic biochar activating oxidation process. Water Res. 2020, 179, 115876. [Google Scholar] [CrossRef] [PubMed]
  53. Shang, J.; Pi, J.; Zong, M.; Wang, Y.; Li, W.; Liao, Q. Chromium removal using magnetic biochar derived from herb-residue. J. Taiwan Inst. Chem. Eng. 2016, 68, 289–294. [Google Scholar] [CrossRef]
  54. Wang, S.; Guo, W.; Gao, F.; Wang, Y.; Gao, Y. Lead and uranium sorptive removal from aqueous solution using magnetic and nonmagnetic fast pyrolysis rice husk biochars. RSC Adv. 2018, 8, 13205–13217. [Google Scholar] [CrossRef]
  55. Wang, L.; Wang, Y.; Ma, F.; Tankpa, V.; Bai, S.; Guo, X.; Wang, X. Mechanisms and reutilization of modified biochar used for removal of heavy metals from wastewater: A review. Sci. Total Environ. 2019, 668, 1298–1309. [Google Scholar] [CrossRef]
  56. Antunes, E.; Jacob, M.V.; Brodie, G.; Schneider, P.A. Silver removal from aqueous solution by biochar produced from biosolids via microwave pyrolysis. J. Environ. Manag. 2017, 203, 264–272. [Google Scholar] [CrossRef]
  57. Chen, M.; He, F.; Hu, D.; Bao, C.; Huang, Q. Broadened operating pH range for adsorption/reduction of aqueous Cr(VI) using biochar from directly treated jute (Corchorus capsularis L.) fibers by H3PO4. Chem. Eng. J. 2020, 381, 122739. [Google Scholar] [CrossRef]
  58. Liang, H.; Sun, R.; Song, B.; Sun, Q.; Peng, P.; She, D. Preparation of nitrogen-doped porous carbon material by a hydrothermal-activation two-step method and its high-efficiency adsorption of Cr(VI). J. Hazard. Mater. 2020, 387, 121987. [Google Scholar] [CrossRef]
  59. Zhang, H.; Shao, J.; Zhang, S.; Zhang, X.; Chen, H. Effect of phosphorus-modified biochars on immobilization of Cu (II), Cd (II), and As (V) in paddy soil. J. Hazard. Mater. 2020, 390, 121349. [Google Scholar] [CrossRef]
  60. Wang, L.; Bolan, N.S.; Tsang, D.C.W.; Hou, D. Green immobilization of toxic metals using alkaline enhanced rice husk biochar: Effects of pyrolysis temperature and KOH concentration. Sci. Total Environ. 2020, 720, 137584. [Google Scholar] [CrossRef]
  61. Bian, S.; Xu, S.; Yin, Z.; Liu, S.; Li, J.; Xu, S.; Zhang, Y. An Efficient Strategy for Enhancing the Adsorption Capabilities of Biochar via Sequential KMnO4-Promoted Oxidative Pyrolysis and H2O2 Oxidation. Sustainability 2021, 13, 2641. [Google Scholar] [CrossRef]
  62. Zhang, Y.; Zheng, Y.; Yang, Y.; Huang, J.; Zimmerman, A.R.; Chen, H.; Hu, X.; Gao, B. Mechanisms and adsorption capacities of hydrogen peroxide modified ball milled biochar for the removal of methylene blue from aqueous solutions. Bioresour. Technol. 2021, 337, 125432. [Google Scholar] [CrossRef] [PubMed]
  63. Xu, S.; Li, J.; Yin, Z.; Liu, S.; Bian, S.; Zhang, Y. A highly efficient strategy for enhancing the adsorptive and magnetic capabilities of biochar using Fenton oxidation. Bioresour. Technol. 2020, 315, 123797. [Google Scholar] [CrossRef] [PubMed]
  64. Fan, Y.; Wang, B.; Yuan, S.; Wu, X.; Chen, J.; Wang, L. Adsorptive removal of chloramphenicol from wastewater by NaOH modified bamboo charcoal. Bioresour. Technol. 2010, 101, 7661–7664. [Google Scholar] [CrossRef] [PubMed]
  65. Liu, Y.; Ma, S.; Chen, J. A novel pyro-hydrochar via sequential carbonization of biomass waste: Preparation, characterization and adsorption capacity. J. Clean. Prod. 2018, 176, 187–195. [Google Scholar] [CrossRef]
  66. Yao, Y.; Zhang, Y.; Gao, B.; Chen, R.; Wu, F. Removal of sulfamethoxazole (SMX) and sulfapyridine (SPY) from aqueous solutions by biochars derived from anaerobically digested bagasse. Environ. Sci. Pollut. Res. 2018, 25, 25659–25667. [Google Scholar] [CrossRef]
  67. Jiang, Y.; Yang, F.; Dai, M.; Ali, I.; Shen, X.; Hou, X.; Alhewairini, S.S.; Peng, C.; Naz, I. Application of microbial immobilization technology for remediation of Cr(VI) contamination: A review. Chemosphere 2022, 286, 131721. [Google Scholar] [CrossRef]
  68. Lin, M.; Li, F.; Li, X.; Rong, X.; Oh, K. Biochar-clay, biochar-microorganism and biochar-enzyme composites for environmental remediation: A review. Environ. Chem. Lett. 2023, 21, 1837–1862. [Google Scholar] [CrossRef]
  69. Inyang, M.; Gao, B.; Zimmerman, A.; Zhou, Y.; Cao, X. Sorption and cosorption of lead and sulfapyridine on carbon nanotube-modified biochars. Environ. Sci. Pollut. Res. 2015, 22, 1868–1876. [Google Scholar] [CrossRef]
  70. Phiri, Z.; Moja, N.T.; Nkambule, T.T.I.; de Kock, L.-A. Utilization of biochar for remediation of heavy metals in aqueous environments: A review and bibliometric analysis. Heliyon 2024, 10, e25785. [Google Scholar] [CrossRef]
  71. Tho, P.T.; Van, H.T.; Nguyen, L.H.; Hoang, T.K.; Ha Tran, T.N.; Nguyen, T.T.; Hanh Nguyen, T.B.; Nguyen, V.Q.; Le Sy, H.; Thai, V.N.; et al. Enhanced simultaneous adsorption of As(iii), Cd(ii), Pb(ii) and Cr(vi) ions from aqueous solution using cassava root husk-derived biochar loaded with ZnO nanoparticles. RSC Adv. 2021, 11, 18881–18897. [Google Scholar] [CrossRef] [PubMed]
  72. Singh, J.K.; Chaurasia, B.; Dubey, A.; Faneite Noguera, A.M.; Gupta, A.; Kothari, R.; Upadhyaya, C.P.; Kumar, A.; Hashem, A.; Alqarawi, A.A.; et al. Biological Characterization and Instrumental Analytical Comparison of Two Biorefining Pretreatments for Water Hyacinth (Eichhornia crassipes) Biomass Hydrolysis. Sustainability 2021, 13, 245. [Google Scholar] [CrossRef]
  73. Yaashikaa, P.R.; Kumar, P.S.; Varjani, S.; Saravanan, A. A critical review on the biochar production techniques, characterization, stability and applications for circular bioeconomy. Biotechnol. Rep. 2020, 28, e00570. [Google Scholar] [CrossRef]
  74. Waqas, M.; Aburiazaiza, A.S.; Miandad, R.; Rehan, M.; Barakat, M.A.; Nizami, A.S. Development of biochar as fuel and catalyst in energy recovery technologies. J. Clean. Prod. 2018, 188, 477–488. [Google Scholar] [CrossRef]
  75. Chen, S.; Qin, C.; Wang, T.; Chen, F.; Li, X.; Hou, H.; Zhou, M. Study on the adsorption of dyestuffs with different properties by sludge-rice husk biochar: Adsorption capacity, isotherm, kinetic, thermodynamics and mechanism. J. Mol. Liq. 2019, 285, 62–74. [Google Scholar] [CrossRef]
  76. Labiadh, L.; Kamali, A.R. Textural, structural and morphological evolution of mesoporous 3D graphene saturated with methyl orange dye during thermal regeneration. Diam. Relat. Mater. 2020, 103, 107698. [Google Scholar] [CrossRef]
  77. Santoso, E.; Ediati, R.; Kusumawati, Y.; Bahruji, H.; Sulistiono, D.O.; Prasetyoko, D. Review on recent advances of carbon based adsorbent for methylene blue removal from waste water. Mater. Today Chem. 2020, 16, 100233. [Google Scholar] [CrossRef]
  78. Hamzah, N.; Tokimatsu, K.; Yoshikawa, K. Solid Fuel from Oil Palm Biomass Residues and Municipal Solid Waste by Hydrothermal Treatment for Electrical Power Generation in Malaysia: A Review. Sustainability 2019, 11, 1060. [Google Scholar] [CrossRef]
  79. Sun, Y.; Xiong, X.; He, M.; Xu, Z.; Hou, D.; Zhang, W.; Ok, Y.S.; Rinklebe, J.; Wang, L.; Tsang, D.C.W. Roles of biochar-derived dissolved organic matter in soil amendment and environmental remediation: A critical review. Chem. Eng. J. 2021, 424, 130387. [Google Scholar] [CrossRef]
  80. Chiappero, M.; Norouzi, O.; Hu, M.; Demichelis, F.; Berruti, F.; Di Maria, F.; Mašek, O.; Fiore, S. Review of biochar role as additive in anaerobic digestion processes. Renew. Sustain. Energy Rev. 2020, 131, 110037. [Google Scholar] [CrossRef]
  81. Świechowski, K.; Rasaq, W.A.; Stegenta-Dąbrowska, S.; Białowiec, A. Characterization of Engineered Biochar: Proximate Analyses, Ultimate Analyses, Physicochemical Analyses, Surface Analyses, and Molecular Analyses. In Engineered Biochar: Fundamentals, Preparation, Characterization and Applications; Ramola, S., Mohan, D., Masek, O., Méndez, A., Tsubota, T., Eds.; Springer Nature: Singapore, 2022; pp. 127–148. [Google Scholar]
  82. Liu, M.; Almatrafi, E.; Zhang, Y.; Xu, P.; Song, B.; Zhou, C.; Zeng, G.; Zhu, Y. A critical review of biochar-based materials for the remediation of heavy metal contaminated environment: Applications and practical evaluations. Sci. Total Environ. 2022, 806, 150531. [Google Scholar] [CrossRef] [PubMed]
  83. Agboola, O.D.; Benson, N.U. Physisorption and chemisorption mechanisms influencing micro (nano) plastics-organic chemical contaminants interactions: A review. Front. Environ. Sci. 2021, 9, 678574. [Google Scholar] [CrossRef]
  84. Heinrich, P.; Hanslik, L.; Kämmer, N.; Braunbeck, T. The tox is in the detail: Technical fundamentals for designing, performing, and interpreting experiments on toxicity of microplastics and associated substances. Environ. Sci. Pollut. Res. 2020, 27, 22292–22318. [Google Scholar] [CrossRef] [PubMed]
  85. Liu, Z.; Zhen, F.; Zhang, Q.; Qian, X.; Li, W.; Sun, Y.; Zhang, L.; Qu, B. Nanoporous biochar with high specific surface area based on rice straw digestion residue for efficient adsorption of mercury ion from water. Bioresour. Technol. 2022, 359, 127471. [Google Scholar] [CrossRef]
  86. Ambaye, T.; Vaccari, M.; van Hullebusch, E.D.; Amrane, A.; Rtimi, S. Mechanisms and adsorption capacities of biochar for the removal of organic and inorganic pollutants from industrial wastewater. Int. J. Environ. Sci. Technol. 2021, 18, 3273–3294. [Google Scholar] [CrossRef]
  87. Wu, J.; Wang, T.; Wang, J.; Zhang, Y.; Pan, W.-P. A novel modified method for the efficient removal of Pb and Cd from wastewater by biochar: Enhanced the ion exchange and precipitation capacity. Sci. Total Environ. 2021, 754, 142150. [Google Scholar] [CrossRef]
  88. Almanassra, I.W.; Chatla, A.; Zakaria, Y.; Kochkodan, V.; Shanableh, A.; Laoui, T.; Atieh, M.A. Palm leaves based biochar: Advanced material characterization and heavy metal adsorption study. Biomass Convers. Biorefinery 2024, 14, 14811–14830. [Google Scholar] [CrossRef]
  89. Ye, Q.; Li, Q.; Li, X. Removal of heavy metals from wastewater using biochars: Adsorption and mechanisms. Environ. Pollut. Bioavailab. 2022, 34, 385–394. [Google Scholar] [CrossRef]
  90. Yu, W.; Hu, J.; Yu, Y.; Ma, D.; Gong, W.; Qiu, H.; Hu, Z.; Gao, H.-w. Facile preparation of sulfonated biochar for highly efficient removal of toxic Pb(II) and Cd(II) from wastewater. Sci. Total Environ. 2021, 750, 141545. [Google Scholar] [CrossRef]
  91. Wang, H.; Chen, Q.; Xia, H.; Liu, R.; Zhang, Y. Enhanced complexation and electrostatic attraction through fabrication of amino- or hydroxyl-functionalized Fe/Ni-biochar composite for the adsorption of Pb(II) and Cd(II). Sep. Purif. Technol. 2024, 328, 125074. [Google Scholar] [CrossRef]
  92. Chen, C.-K.; Chen, J.-J.; Nguyen, N.-T.; Le, T.-T.; Nguyen, N.-C.; Chang, C.-T. Specifically designed magnetic biochar from waste wood for arsenic removal. Sustain. Environ. Res. 2021, 31, 29. [Google Scholar] [CrossRef]
  93. Chen, Y.; Liu, Y.; Li, Y.; Chen, Y.; Wu, Y.; Li, H.; Wang, S.; Peng, Z.; Xu, R.; Zeng, Z. Novel magnetic pomelo peel biochar for enhancing pb(ii) and cu(ii) adsorption: Performance and mechanism. Water Air Soil Pollut. 2020, 231, 404. [Google Scholar] [CrossRef]
  94. Li, Y.; Wang, S.; Ouyang, X.F.; Dang, Z.; Yin, H. Acetate anions intercalated Fe/Mg-layered double hydroxides modified biochar for efficient adsorption of anionic and cationic heavy metal ions from polluted water. Chemosphere 2024, 362, 142652. [Google Scholar] [CrossRef]
  95. Vijayaraghavan, K. The importance of mineral ingredients in biochar production, properties and applications. Crit. Rev. Environ. Sci. Technol. 2021, 51, 113–139. [Google Scholar] [CrossRef]
  96. Frolova, L.; Kharytonov, M.; Klimkina, I.; Kovrov, O.; Koveria, A. Investigation of the adsorption of ions chromium by mean biochar from coniferous trees. Appl. Nanosci. 2022, 12, 1123–1129. [Google Scholar] [CrossRef]
  97. Li, A.Y.; Deng, H.; Jiang, Y.H.; Ye, C.H.; Yu, B.G.; Zhou, X.L.; Ma, A.Y. Superefficient removal of heavy metals from wastewater by Mg-loaded biochars: Adsorption characteristics and removal mechanisms. Langmuir 2020, 36, 9160–9174. [Google Scholar] [CrossRef]
  98. Guo, X.; Liu, A.; Lu, J.; Niu, X.; Jiang, M.; Ma, Y.; Liu, X.; Li, M. Adsorption mechanism of hexavalent chromium on biochar: Kinetic, thermodynamic, and characterization studies. ACS Omega 2020, 5, 27323–27331. [Google Scholar] [CrossRef]
  99. Liu, X.; Li, G.; Chen, C.; Zhang, X.; Zhou, K.; Long, X. Banana stem and leaf biochar as an effective adsorbent for cadmium and lead in aqueous solution. Sci. Rep. 2022, 12, 1584. [Google Scholar] [CrossRef]
  100. Wang, Y.-P.; Liu, Y.-L.; Tian, S.-Q.; Yang, J.-J.; Wang, L.; Ma, J. Straw biochar enhanced removal of heavy metal by ferrate. J. Hazard. Mater. 2021, 416, 126128. [Google Scholar] [CrossRef] [PubMed]
  101. Xie, Y.; Wang, L.; Li, H.; Westholm, L.J.; Carvalho, L.; Thorin, E.; Yu, Z.; Yu, X.; Skreiberg, Ø. A critical review on production, modification and utilization of biochar. J. Anal. Appl. Pyrolysis 2022, 161, 105405. [Google Scholar] [CrossRef]
  102. Xie, J.; Lin, R.; Liang, Z.; Zhao, Z.; Yang, C.; Cui, F. Effect of cations on the enhanced adsorption of cationic dye in Fe3O4-loaded biochar and mechanism. J. Environ. Chem. Eng. 2021, 9, 105744. [Google Scholar] [CrossRef]
  103. Tan, Z.; Yuan, S.; Hong, M.; Zhang, L.; Huang, Q. Mechanism of negative surface charge formation on biochar and its effect on the fixation of soil Cd. J. Hazard. Mater. 2020, 384, 121370. [Google Scholar] [CrossRef] [PubMed]
  104. Chacón, F.J.; Sánchez-Monedero, M.A.; Lezama, L.; Cayuela, M.L. Enhancing biochar redox properties through feedstock selection, metal preloading and post-pyrolysis treatments. Chem. Eng. J. 2020, 395, 125100. [Google Scholar] [CrossRef]
  105. 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]
  106. Dinh, V.C.; Hou, C.-H.; Dao, T.N. O, N-doped porous biochar by air oxidation for enhancing heavy metal removal: The role of O, N functional groups. Chemosphere 2022, 293, 133622. [Google Scholar] [CrossRef]
  107. Meng, X.; Hu, R. Nitrogen/phosphorus enriched biochar with enhanced porosity activated by guanidine phosphate for efficient passivation of Pb(II), Cu(II) and Cd(II). J. Mol. Liq. 2021, 323, 115071. [Google Scholar] [CrossRef]
  108. Zeng, B.; Xu, W.; Khan, S.B.; Wang, Y.; Zhang, J.; Yang, J.; Su, X.; Lin, Z. Preparation of sludge biochar rich in carboxyl/hydroxyl groups by quenching process and its excellent adsorption performance for Cr(VI). Chemosphere 2021, 285, 131439. [Google Scholar] [CrossRef]
  109. Pan, J.; Deng, H.; Du, Z.; Tian, K.; Zhang, J. Design of nitrogen-phosphorus-doped biochar and its lead adsorption performance. Environ. Sci. Pollut. Res. 2022, 29, 28984–28994. [Google Scholar] [CrossRef]
  110. Huang, H.; Wang, Y.; Zhang, Y.; Niu, Z.; Li, X. Amino-functionalized graphene oxide for Cr(VI), Cu(II), Pb(II) and Cd(II) removal from industrial wastewater. Open Chem. 2020, 18, 97–107. [Google Scholar] [CrossRef]
  111. Deng, R.; Huang, D.; Wan, J.; Xue, W.; Lei, L.; Wen, X.; Liu, X.; Chen, S.; Yang, Y.; Li, Z.; et al. Chloro-phosphate impregnated biochar prepared by co-precipitation for the lead, cadmium and copper synergic scavenging from aqueous solution. Bioresour. Technol. 2019, 293, 122102. [Google Scholar] [CrossRef]
  112. Li, H.; Dong, X.; da Silva, E.B.; de Oliveira, L.M.; Chen, Y.; Ma, L.Q. Mechanisms of metal sorption by biochars: Biochar characteristics and modifications. Chemosphere 2017, 178, 466–478. [Google Scholar] [CrossRef]
  113. Qiu, Y.; Zhang, Q.; Gao, B.; Li, M.; Fan, Z.; Sang, W.; Hao, H.; Wei, X. Removal mechanisms of Cr(VI) and Cr(III) by biochar supported nanosized zero-valent iron: Synergy of adsorption, reduction and transformation. Environ. Pollut. 2020, 265, 115018. [Google Scholar] [CrossRef]
  114. Biswal, B.K.; Balasubramanian, R. Use of biochar as a low-cost adsorbent for removal of heavy metals from water and wastewater: A review. J. Environ. Chem. Eng. 2023, 11, 110986. [Google Scholar] [CrossRef]
  115. Zhang, P.; Zhang, X.; Yuan, X.; Xie, R.; Han, L. Characteristics, adsorption behaviors, Cu(II) adsorption mechanisms by cow manure biochar derived at various pyrolysis temperatures. Bioresour. Technol. 2021, 331, 125013. [Google Scholar] [CrossRef] [PubMed]
  116. Paredes-Quevedo, L.C.; González-Caicedo, C.; Torres-Luna, J.A.; Carriazo, J.G. Removal of a Textile Azo-Dye (Basic Red 46) in Water by Efficient Adsorption on a Natural Clay. Water Air Soil Pollut. 2021, 232, 4. [Google Scholar] [CrossRef]
  117. Xu, Z.; Xing, Y.; Ren, A.; Ma, D.; Li, Y.; Hu, S. Study on adsorption properties of water hyacinth-derived biochar for uranium (VI). J. Radioanal. Nucl. Chem. 2020, 324, 1317–1327. [Google Scholar] [CrossRef]
  118. Zhang, Z.; Wang, T.; Zhang, H.; Liu, Y.; Xing, B. Adsorption of Pb(II) and Cd(II) by magnetic activated carbon and its mechanism. Sci. Total Environ. 2021, 757, 143910. [Google Scholar] [CrossRef]
  119. Yang, Z.; Liu, X.; Zhang, M.; Liu, L.; Xu, X.; Xian, J.; Cheng, Z. Effect of temperature and duration of pyrolysis on spent tea leaves biochar: Physiochemical properties and Cd(II) adsorption capacity. Water Sci. Technol. 2020, 81, 2533–2544. [Google Scholar] [CrossRef]
  120. Budnyak, T.M.; Błachnio, M.; Slabon, A.; Jaworski, A.; Tertykh, V.A.; Deryło-Marczewska, A.; Marczewski, A.W. Chitosan deposited onto fumed silica surface as sustainable hybrid biosorbent for Acid Orange 8 dye capture: Effect of temperature in adsorption equilibrium and kinetics. J. Phys. Chem. C 2020, 124, 15312–15323. [Google Scholar] [CrossRef]
  121. Harsha Vardhan, K.; Kumar, P.S.; Panda, R.C. Adsorption of copper ions from polluted water using biochar derived from waste renewable resources: Static and dynamic analysis. Int. J. Environ. Anal. Chem. 2022, 102, 4067–4088. [Google Scholar] [CrossRef]
  122. Singh, S.; Anil, A.G.; Naik, T.S.; Basavaraju, U.; Khasnabis, S.; Nath, B.; Kumar, V.; Subramanian, S.; Singh, J.; Ramamurthy, P.C. Mechanism and kinetics of Cr(VI) adsorption on biochar derived from Citrobacter freundii under different pyrolysis temperatures. J. Water Process Eng. 2022, 47, 102723. [Google Scholar] [CrossRef]
  123. Yang, H.I.; Lou, K.; Rajapaksha, A.U.; Ok, Y.S.; Anyia, A.O.; Chang, S.X. Adsorption of ammonium in aqueous solutions by pine sawdust and wheat straw biochars. Environ. Sci. Pollut. Res. 2018, 25, 25638–25647. [Google Scholar] [CrossRef]
  124. Zhang, M.; Song, G.; Gelardi, D.L.; Huang, L.; Khan, E.; Mašek, O.; Parikh, S.J.; Ok, Y.S. Evaluating biochar and its modifications for the removal of ammonium, nitrate, and phosphate in water. Water Res. 2020, 186, 116303. [Google Scholar] [CrossRef]
  125. Wang, X.; Zhao, Y.; Deng, J.; Zhou, Y.; Yuan, S. Study on the influence mechanism of mineral components in biochar on the adsorption of Cr(VI). Fuel 2023, 340, 127631. [Google Scholar] [CrossRef]
  126. Herath, A.; Layne, C.A.; Perez, F.; Hassan, E.I.B.; Pittman, C.U.; Mlsna, T.E. KOH-activated high surface area Douglas Fir biochar for adsorbing aqueous Cr(VI), Pb(II) and Cd(II). Chemosphere 2021, 269, 128409. [Google Scholar] [CrossRef] [PubMed]
  127. Guo, N.; Lv, X.; Yang, Q.; Xu, X.; Song, H. Effective removal of hexavalent chromium from aqueous solution by ZnCl2 modified biochar: Effects and response sequence of the functional groups. J. Mol. Liq. 2021, 334, 116149. [Google Scholar] [CrossRef]
  128. Sun, T.; Pei, P.; Sun, Y.; Xu, Y.; Jia, H. Performance and mechanism of As(III/V) removal from aqueous solution by novel positively charged animal-derived biochar. Sep. Purif. Technol. 2022, 290, 120836. [Google Scholar] [CrossRef]
  129. Edo, G.I.; Samuel, P.O.; Oloni, G.O.; Ezekiel, G.O.; Ikpekoro, V.O.; Obasohan, P.; Ongulu, J.; Otunuya, C.F.; Opiti, A.R.; Ajakaye, R.S.; et al. Environmental persistence, bioaccumulation, and ecotoxicology of heavy metals. Chem. Ecol. 2024, 40, 322–349. [Google Scholar] [CrossRef]
  130. Qasem, N.A.A.; Mohammed, R.H.; Lawal, D.U. Removal of heavy metal ions from wastewater: A comprehensive and critical review. npj Clean Water 2021, 4, 36. [Google Scholar] [CrossRef]
  131. Alsawy, T.; Rashad, E.; El-Qelish, M.; Mohammed, R.H. A comprehensive review on the chemical regeneration of biochar adsorbent for sustainable wastewater treatment. npj Clean Water 2022, 5, 29. [Google Scholar] [CrossRef]
  132. Zhang, R.-H.; Xie, Y.; Zhou, G.; Li, Z.; Ye, A.; Huang, X.; Xie, Y.; Shi, L.; Cao, X.; Zhang, J.; et al. The effects of short-term, long-term, and reapplication of biochar on the remediation of heavy metal-contaminated soil. Ecotoxicol. Environ. Saf. 2022, 248, 114316. [Google Scholar] [CrossRef] [PubMed]
  133. Burachevskaya, M.; Minkina, T.; Bauer, T.; Lobzenko, I.; Fedorenko, A.; Mazarji, M.; Sushkova, S.; Mandzhieva, S.; Nazarenko, A.; Butova, V.; et al. Fabrication of biochar derived from different types of feedstocks as an efficient adsorbent for soil heavy metal removal. Sci. Rep. 2023, 13, 2020. [Google Scholar] [CrossRef] [PubMed]
  134. Wang, Y.; Wang, K.; Wang, X.; Zhao, Q.; Jiang, J.; Jiang, M. Effect of different production methods on physicochemical properties and adsorption capacities of biochar from sewage sludge and kitchen waste: Mechanism and correlation analysis. J. Hazard. Mater. 2024, 461, 132690. [Google Scholar] [CrossRef] [PubMed]
  135. Singh, E.; Kumar, A.; Mishra, R.; You, S.; Singh, L.; Kumar, S.; Kumar, R. Pyrolysis of waste biomass and plastics for production of biochar and its use for removal of heavy metals from aqueous solution. Bioresour. Technol. 2021, 320, 124278. [Google Scholar] [CrossRef] [PubMed]
  136. Qian, L.; Zhang, W.; Yan, J.; Han, L.; Chen, Y.; Ouyang, D.; Chen, M. Nanoscale zero-valent iron supported by biochars produced at different temperatures: Synthesis mechanism and effect on Cr(VI) removal. Environ. Pollut. 2017, 223, 153–160. [Google Scholar] [CrossRef]
  137. Cheng, S.; Meng, W.; Xing, B.; Shi, C.; Wang, Q.; Xia, D.; Nie, Y.; Yi, G.; Zhang, C.; Xia, H. Efficient removal of heavy metals from aqueous solutions by Mg/Fe bimetallic oxide-modified biochar: Experiments and DFT investigations. J. Clean. Prod. 2023, 403, 136821. [Google Scholar] [CrossRef]
  138. Truong, Q.-M.; Nguyen, T.-B.; Chen, W.-H.; Chen, C.-W.; Patel, A.K.; Bui, X.-T.; Singhania, R.R.; Dong, C.-D. Removal of heavy metals from aqueous solutions by high performance capacitive deionization process using biochar derived from Sargassum hemiphyllum. Bioresour. Technol. 2023, 370, 128524. [Google Scholar] [CrossRef]
  139. Li, Y.; Gao, L.; Lu, Z.; Wang, Y.; Wang, Y.; Wan, S. Enhanced Removal of Heavy Metals from Water by Hydrous Ferric Oxide-Modified Biochar. ACS Omega 2020, 5, 28702–28711. [Google Scholar] [CrossRef]
  140. Kamran, U.; Park, S.-J. Hybrid biochar supported transition metal doped MnO2 composites: Efficient contenders for lithium adsorption and recovery from aqueous solutions. Desalination 2022, 522, 115387. [Google Scholar] [CrossRef]
  141. Wang, Y.; Yang, Q.; Chen, J.; Yang, J.; Zhang, Y.; Chen, Y.; Li, X.; Du, W.; Liang, A.; Ho, S.-H.; et al. Adsorption behavior of Cr(VI) by magnetically modified Enteromorpha prolifera based biochar and the toxicity analysis. J. Hazard. Mater. 2020, 395, 122658. [Google Scholar] [CrossRef]
  142. Cho, D.-W.; Chon, C.-M.; Yim, G.-J.; Ryu, J.; Jo, H.; Kim, S.-J.; Jang, J.-Y.; Song, H. Adsorption of potentially harmful elements by metal-biochar prepared via Co-pyrolysis of coffee grounds and Nano Fe(III) oxides. Chemosphere 2023, 319, 136536. [Google Scholar] [CrossRef]
  143. Han, J.; Song, Y.; Li, H.; Wang, Y.; Zhang, L.; Sun, P.; Fan, J.; Li, Y. Preparation of novel magnetic porous biochar and its adsorption mechanism on cerium in rare earth wastewater. Ceram. Int. 2023, 49, 9901–9908. [Google Scholar] [CrossRef]
  144. Georgiou, E.; Ioannidis, I.; Pashalidis, I.; Viet, D.D.; Tsubota, T.; Kalderis, D. Europium removal from aqueous solutions by oxidized biochar prepared from waste palm tree fronds. Sustain. Chem. Environ. 2023, 4, 100040. [Google Scholar] [CrossRef]
  145. Liu, L.; Feng, B.; Rao, Y.Z.; Tian, C.S.; Gu, Q.X.; Huang, T.; Iqhrammullah, M. Development of Efficient Biochar Produced from Orange Peel for Effective La(III) and Y(III) Adsorption. Adsorpt. Sci. Technol. 2023, 2023, 5519783. [Google Scholar] [CrossRef]
  146. Yin, M.; Bai, X.; Wu, D.; Li, F.; Jiang, K.; Ma, N.; Chen, Z.; Zhang, X.; Fang, L. Sulfur-functional group tunning on biochar through sodium thiosulfate modified molten salt process for efficient heavy metal adsorption. Chem. Eng. J. 2022, 433, 134441. [Google Scholar] [CrossRef]
  147. Liu, N.; Zhang, Y.; Xu, C.; Liu, P.; Lv, J.; Liu, Y.; Wang, Q. Removal mechanisms of aqueous Cr(VI) using apple wood biochar: A spectroscopic study. J. Hazard. Mater. 2020, 384, 121371. [Google Scholar] [CrossRef]
  148. Xing, Y.; Luo, X.; Liu, S.; Wan, W.; Huang, Q.; Chen, W. A novel eco-friendly recycling of food waste for preparing biofilm-attached biochar to remove Cd and Pb in wastewater. J. Clean. Prod. 2021, 311, 127514. [Google Scholar] [CrossRef]
  149. Yang, Y.; Luo, X.; Zhang, J.; Ma, X.; Sun, P.; Zhao, L. Sewage sludge–coconut fiber co-pyrolysis biochar: Mechanisms underlying synergistic heavy metal stabilization and ciprofloxacin adsorption. J. Clean. Prod. 2022, 375, 134149. [Google Scholar] [CrossRef]
  150. Charnkeitkong, P.; Sripiboon, S. Effects of Cupric Ion Adsorption onto the Modified Pineapple Pulp as a Biochar Adsorbent. Key Eng. Mater. 2024, 974, 57–62. [Google Scholar] [CrossRef]
  151. Jomova, K.; Makova, M.; Alomar, S.Y.; Alwasel, S.H.; Nepovimova, E.; Kuca, K.; Rhodes, C.J.; Valko, M. Essential metals in health and disease. Chem.-Biol. Interact. 2022, 367, 110173. [Google Scholar] [CrossRef]
  152. Gupta, S.; Sireesha, S.; Sreedhar, I.; Patel, C.M.; Anitha, K.L. Latest trends in heavy metal removal from wastewater by biochar based sorbents. J. Water Process Eng. 2020, 38, 101561. [Google Scholar] [CrossRef]
  153. Sigdel, A.; Park, J.; Kwak, H.; Park, P.K. Arsenic removal from aqueous solutions by adsorption onto hydrous iron oxide-impregnated alginate beads. J. Ind. Eng. Chem. 2016, 35, 277–286. [Google Scholar] [CrossRef]
  154. Talan, D.; Huang, Q. A review of environmental aspect of rare earth element extraction processes and solution purification techniques. Miner. Eng. 2022, 179, 107430. [Google Scholar] [CrossRef]
  155. Liu, W.-S.; Chen, Y.-Y.; Huot, H.; Liu, C.; Guo, M.-N.; Qiu, R.-L.; Morel, J.L.; Tang, Y.-T. Phytoextraction of rare earth elements from ion-adsorption mine tailings by Phytolacca americana: Effects of organic material and biochar amendment. J. Clean. Prod. 2020, 275, 122959. [Google Scholar] [CrossRef]
  156. Yang, Y.; Liu, W.; Weng, X.; Chen, Z.; Owens, G.; Chen, Z. Highly selective recovery of rare earth elements from mining wastewater using phyto-synthesized biochar dispersed iron nanoparticles. Sep. Purif. Technol. 2025, 353, 128491. [Google Scholar] [CrossRef]
  157. Adebayo, O.; LaDouceur, R.; Nasrullah, Z. Process Intensification for Rare Earth Elements Adsorption by Resonant Vibratory Mixing. Grad. Theses Non-Theses 2024, 320. [Google Scholar]
  158. Ifthikar, J.; Jiao, X.; Ngambia, A.; Wang, T.; Khan, A.; Jawad, A.; Xue, Q.; Liu, L.; Chen, Z. Facile One-Pot Synthesis of Sustainable Carboxymethyl Chitosan—Sewage Sludge Biochar for Effective Heavy Metal Chelation and Regeneration. Bioresour. Technol. 2018, 262, 22–31. [Google Scholar] [CrossRef]
  159. Fseha, Y.H.; Sizirici, B.; Yildiz, I.; Yavuz, C. Pristine biochar performance investigation to remove metals in primary and secondary treated municipal wastewater for groundwater recharge application. PLoS ONE 2022, 17, e0278315. [Google Scholar] [CrossRef]
  160. Wang, S.; Tang, Y.; Chen, C.; Wu, J.; Huang, Z.; Mo, Y.; Zhang, K.; Chen, J. Regeneration of magnetic biochar derived from eucalyptus leaf residue for lead(II) removal. Bioresour. Technol. 2015, 186, 360–364. [Google Scholar] [CrossRef]
  161. Poonam; Bharti, S.K.; Kumar, N. Kinetic study of lead (Pb2+) removal from battery manufacturing wastewater using bagasse biochar as biosorbent. Appl. Water Sci. 2018, 8, 119. [Google Scholar] [CrossRef]
  162. Alqadami, A.A.; Khan, M.A.; Otero, M.; Siddiqui, M.R.; Jeon, B.-H.; Batoo, K.M. A magnetic nanocomposite produced from camel bones for an efficient adsorption of toxic metals from water. J. Clean. Prod. 2018, 178, 293–304. [Google Scholar] [CrossRef]
  163. Gan, C.; Liu, Y.; Tan, X.; Wang, S.; Zeng, G.; Zheng, B.; Li, T.; Jiang, Z.; Liu, W. Effect of porous zinc–biochar nanocomposites on Cr(vi) adsorption from aqueous solution. RSC Adv. 2015, 5, 35107–35115. [Google Scholar] [CrossRef]
  164. Ding, Y.; Liu, Y.; Liu, S.; Li, Z.; Tan, X.; Huang, X.; Zeng, G.; Zhou, Y.; Zheng, B.; Cai, X. Competitive removal of Cd(ii) and Pb(ii) by biochars produced from water hyacinths: Performance and mechanism. RSC Adv. 2016, 6, 5223–5232. [Google Scholar] [CrossRef]
  165. Cui, X.; Wang, J.; Wang, X.; Du, G.; Khan, K.Y.; Yan, B.; Cheng, Z.; Chen, G. Pyrolysis of exhausted hydrochar sorbent for cadmium separation and biochar regeneration. Chemosphere 2022, 306, 135546. [Google Scholar] [CrossRef] [PubMed]
  166. Yang, Q.; Cui, P.; Liu, C.; Fang, G.; Huang, M.; Wang, Q.; Zhou, Y.; Hou, H.; Wang, Y. In situ stabilization of the adsorbed Co2+ and Ni2+ in rice straw biochar based on LDH and its reutilization in the activation of peroxymonosulfate. J. Hazard. Mater. 2021, 416, 126215. [Google Scholar] [CrossRef]
  167. Pan, J.; Gao, B.; Duan, P.; Guo, K.; Xu, X.; Yue, Q. Recycling exhausted magnetic biochar with adsorbed Cu2+ as a cost-effective permonosulfate activator for norfloxacin degradation: Cu contribution and mechanism. J. Hazard. Mater. 2021, 413, 125413. [Google Scholar] [CrossRef]
  168. Chen, Z.; Zheng, R.; Wei, W.; Wei, W.; Zou, W.; Li, J.; Ni, B.-J.; Chen, H. Recycling spent water treatment adsorbents for efficient electrocatalytic water oxidation reaction. Resour. Conserv. Recycl. 2022, 178, 106037. [Google Scholar] [CrossRef]
  169. Mer, K.; Sajjadi, B.; Egiebor, N.O.; Chen, W.-Y.; Mattern, D.L.; Tao, W. Enhanced degradation of organic contaminants using catalytic activity of carbonaceous structures: A strategy for the reuse of exhausted sorbents. J. Environ. Sci. 2021, 99, 267–273. [Google Scholar] [CrossRef]
  170. Zubrik, A.; Matik, M.; Mačingová, E.; Danková, Z.; Jáger, D.; Briančin, J.; Machala, L.; Pechoušek, J.; Hredzák, S. The use of microwave irradiation for preparation and fast-acting regeneration of magnetic biochars. Chem. Eng. Process.-Process Intensif. 2022, 178, 109016. [Google Scholar] [CrossRef]
  171. Wu, Y.; Wang, Z.; Yan, Y.; Zhou, Y.; Huma, B.; Tan, Z.; Zhou, T. Recovery and regeneration of water-hardened magnetic composite biochar sphere for the removal of multiple heavy metals in contaminated soils. J. Clean. Prod. 2024, 450, 141906. [Google Scholar] [CrossRef]
  172. Dai, Y.; Zheng, H.; Jiang, Z.; Xing, B. Combined effects of biochar properties and soil conditions on plant growth: A meta-analysis. Sci. Total Environ. 2020, 713, 136635. [Google Scholar] [CrossRef]
  173. Murtaza, G.; Ahmed, Z.; Eldin, S.M.; Ali, B.; Bawazeer, S.; Usman, M.; Iqbal, R.; Neupane, D.; Ullah, A.; Khan, A. Biochar-Soil-Plant interactions: A cross talk for sustainable agriculture under changing climate. Front. Environ. Sci. 2023, 11, 1059449. [Google Scholar] [CrossRef]
  174. Jiang, L.; Yi, X.; Xu, B.; Lai, K. Effect of wheat straw derived biochar on immobilization of Cd and Pb in single- and binary-metal contaminated soil. Hum. Ecol. Risk Assess. Int. J. 2020, 26, 2420–2433. [Google Scholar] [CrossRef]
  175. Carneiro, J.S.d.S.; Leite, D.A.d.C.; Castro, G.M.d.; Franca, J.R.; Botelho, L.; Soares, J.R.; Oliveira, J.E.d.; Melo, L.C.A. Biochar-graphene oxide composite is efficient to adsorb and deliver copper and zinc in tropical soil. J. Clean. Prod. 2022, 360, 132170. [Google Scholar] [CrossRef]
  176. Hou, J.; Pugazhendhi, A.; Sindhu, R.; Vinayak, V.; Thanh, N.C.; Brindhadevi, K.; Chi, N.T.L.; Yuan, D. An assessment of biochar as a potential amendment to enhance plant nutrient uptake. Environ. Res. 2022, 214, 113909. [Google Scholar] [CrossRef]
  177. Singh Yadav, S.P.; Bhandari, S.; Bhatta, D.; Poudel, A.; Bhattarai, S.; Yadav, P.; Ghimire, N.; Paudel, P.; Paudel, P.; Shrestha, J.; et al. Biochar application: A sustainable approach to improve soil health. J. Agric. Food Res. 2023, 11, 100498. [Google Scholar] [CrossRef]
  178. Yameen, M.Z.; Naqvi, S.R.; Juchelková, D.; Khan, M.N.A. Harnessing the power of functionalized biochar: Progress, challenges, and future perspectives in energy, water treatment, and environmental sustainability. Biochar 2024, 6, 25. [Google Scholar] [CrossRef]
  179. Zeghioud, H.; Fryda, L.; Djelal, H.; Assadi, A.; Kane, A. A comprehensive review of biochar in removal of organic pollutants from wastewater: Characterization, toxicity, activation/functionalization and influencing treatment factors. J. Water Process Eng. 2022, 47, 102801. [Google Scholar] [CrossRef]
  180. Guo, T.; Yang, Q.; Qiu, R.; Gao, J.; Shi, J.; Lei, X.; Zhao, Z. Efficient Degradation of Ciprofloxacin in Water over Copper-Loaded Biochar Using an Enhanced Non-Radical Pathway. Molecules 2023, 28, 8094. [Google Scholar] [CrossRef]
  181. Ji, J.; Aleisa, R.M.; Duan, H.; Zhang, J.; Yin, Y.; Xing, M. Metallic active sites on MoO2 (110) surface to catalyze advanced oxidation processes for efficient pollutant removal. Iscience 2020, 23, 100861. [Google Scholar] [CrossRef]
  182. Sun, E.; Zhang, Y.; Xiao, Q.; Li, H.; Qu, P.; Yong, C.; Wang, B.; Feng, Y.; Huang, H.; Yang, L.; et al. Formable porous biochar loaded with La-Fe(hydr)oxides/montmorillonite for efficient removal of phosphorus in wastewater: Process and mechanisms. Biochar 2022, 4, 53. [Google Scholar] [CrossRef]
  183. Wang, B.-S.; Cao, J.-P.; Zhao, X.-Y.; Bian, Y.; Song, C.; Zhao, Y.-P.; Fan, X.; Wei, X.-Y.; Takarada, T. Preparation of nickel-loaded on lignite char for catalytic gasification of biomass. Fuel Process. Technol. 2015, 136, 17–24. [Google Scholar] [CrossRef]
  184. Li, Y.; Williams, P.T. Catalytic Biochar and Refuse-Derived Char for the Steam Reforming of Waste Plastics Pyrolysis Volatiles for Hydrogen-Rich Syngas. Ind. Eng. Chem. Res. 2023, 62, 14335–14348. [Google Scholar] [CrossRef]
  185. Wu, B.; Jia, Y.; Xu, N.; Liao, L.; Zhang, C.; Wang, Z.; Shan, Y.; Feng, W.; Xue, H. MoO2-Enhanced Fe-Loaded Biochar Promotes Fe2+/Fe3+ Cycling for Activation of Peroxydisulfate to Degrade Organic Matter. Environ. Technol. Innov. 2024, 35, 103736. [Google Scholar] [CrossRef]
  186. Ouyang, D.; Wu, R.; Xu, Z.; Zhu, X.; Cai, Y.; Chen, R.; Zhu, C.; Barceló, D.; Zhang, H. Efficient degradation of Bisphenol A by Fe3+/Fe2+ cycle activating persulfate with the assistance of biochar-supported MoO2. Chem. Eng. J. 2023, 455, 140381. [Google Scholar] [CrossRef]
  187. Li, L.; Zhang, Y.; Yang, S.; Zhang, S.; Xu, Q.; Chen, P.; Du, Y.; Xing, Y. Cobalt-loaded cherry core biochar composite as an effective heterogeneous persulfate catalyst for bisphenol A degradation. RSC Adv. 2022, 12, 7284–7294. [Google Scholar] [CrossRef]
  188. Li, S.; Yang, F.; Li, J.; Cheng, K. Porous biochar-nanoscale zero-valent iron composites: Synthesis, characterization and application for lead ion removal. Sci. Total Environ. 2020, 746, 141037. [Google Scholar] [CrossRef]
  189. Liu, L.; Yu, R.; Zhao, S.; Cao, X.; Zhang, X.; Bai, S. Heterogeneous Fenton system driven by iron-loaded sludge biochar for sulfamethoxazole-containing wastewater treatment. J. Environ. Manag. 2023, 335, 117576. [Google Scholar] [CrossRef] [PubMed]
  190. Zhang, Z.; Yang, X.; Wang, J.; Zhuang, T.; Liu, S.; Dou, M.; Huo, K.; Zhou, Y.; Ding, G. Magnetic biochar pyrolyzed from municipal sludge for Fenton-like degradation of thiamethoxam: Characteristics and mechanism. J. Water Process Eng. 2023, 51, 103391. [Google Scholar] [CrossRef]
  191. Ahmad, M.; Islam, I.U.; Ahmad, M.; Rukh, S.; Ullah, I. Preparation of iron-modified biochar from rice straw and its application for the removal of lead (Pb+2) from lead-contaminated water by adsorption. Chem. Pap. 2022, 76, 3789–3808. [Google Scholar] [CrossRef]
  192. Xu, L.; Shu, Z.; Feng, L.; Zhou, J.; Li, T.; Zhao, Z.; Wang, W. Fresh biomass derived biochar with high-load zero-valent iron prepared in one step for efficient arsenic removal. J. Clean. Prod. 2022, 352, 131616. [Google Scholar] [CrossRef]
  193. Zhou, T.; Shi, C.; Wang, Y.; Wang, X.; Lei, Z.; Liu, X.; Wu, J.; Luo, F.; Wang, L. Progress of metal-loaded biochar-activated persulfate for degradation of emerging organic contaminants. Water Sci. Technol. 2024, 90, 824–843. [Google Scholar] [CrossRef] [PubMed]
  194. Zhao, Y.; Yuan, X.; Li, X.; Jiang, L.; Wang, H. Burgeoning prospects of biochar and its composite in persulfate-advanced oxidation process. J. Hazard. Mater. 2021, 409, 124893. [Google Scholar] [CrossRef]
  195. 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]
  196. Huang, X.; Wei, D.; Zhang, X.; Fan, D.; Sun, X.; Du, B.; Wei, Q. Synthesis of amino-functionalized magnetic aerobic granular sludge-biochar for Pb (II) removal: Adsorption performance and mechanism studies. Sci. Total Environ. 2019, 685, 681–689. [Google Scholar] [CrossRef] [PubMed]
  197. Gupta, S.; Patel, P.; Mondal, P. Catalytic Pyrolysis Using a Nickel-Functionalized Chemically Activated Biochar Catalyst: Insight into Process Kinetics, Products, and Mechanism. ACS Sustain. Chem. Eng. 2022, 10, 5770–5780. [Google Scholar] [CrossRef]
  198. He, J.; Zhao, C.; Lercher, J.A. Ni-catalyzed cleavage of aryl ethers in the aqueous phase. J. Am. Chem. Soc. 2012, 134, 20768–20775. [Google Scholar] [CrossRef]
  199. Xie, J.-X.; Cao, J.-P.; Jiang, W.; Zhao, X.-Y.; Zhao, L.; Zhang, C.; Bai, H.-C. Nickel loaded on biochar prepared from different carbon sources for selective hydrogenolysis of diphenyl ether. Fuel Process. Technol. 2022, 231, 107219. [Google Scholar] [CrossRef]
  200. Dai, L.; Zeng, Z.; Tian, X.; Jiang, L.; Yu, Z.; Wu, Q.; Wang, Y.; Liu, Y.; Ruan, R. Microwave-assisted catalytic pyrolysis of torrefied corn cob for phenol-rich bio-oil production over Fe modified bio-char catalyst. J. Anal. Appl. Pyrolysis 2019, 143, 104691. [Google Scholar] [CrossRef]
  201. Bu, Q.; Cao, M.; Wang, M.; Vasudevan, S.V.; Mao, H. Enhancement of bio-oil quality over self-derived bio-char catalyst via microwave catalytic pyrolysis of peanut shell. J. Anal. Appl. Pyrolysis 2022, 164, 105534. [Google Scholar] [CrossRef]
  202. Liu, Z.; Li, P.; Chang, C.; Wang, X.; Song, J.; Fang, S.; Pang, S. Influence of metal chloride modified biochar on products characteristics from biomass catalytic pyrolysis. Energy 2022, 250, 123776. [Google Scholar] [CrossRef]
  203. Zeng, Z.; Tian, X.; Wang, Y.; Cui, X.; Zhang, Q.; Dai, L.; Liu, Y.; Zou, R.; Chen, J.; Liu, J. Microwave-assisted catalytic pyrolysis of corn cobs with Fe-modified Choerospondias axillaris seed-based biochar catalyst for phenol-rich bio-oil. J. Anal. Appl. Pyrolysis 2021, 159, 105306. [Google Scholar] [CrossRef]
  204. Wang, W.; Wang, M.; Huang, J.; Tang, N.; Dang, Z.; Shi, Y.; Zhaohe, M. Microwave-assisted catalytic pyrolysis of cellulose for phenol-rich bio-oil production. J. Energy Inst. 2019, 92, 1997–2003. [Google Scholar] [CrossRef]
  205. Zhang, Y.; Lei, H.; Yang, Z.; Duan, D.; Villota, E.; Ruan, R. From glucose-based carbohydrates to phenol-rich bio-oils integrated with syngas production via catalytic pyrolysis over an activated carbon catalyst. Green Chem. 2018, 20, 3346–3358. [Google Scholar] [CrossRef]
  206. Hoang Pham, L.K.; Vi Tran, T.T.; Kongparakul, S.; Reubroycharoen, P.; Ding, M.; Guan, G.; Vo, D.-V.N.; Jaiyong, P.; Youngvises, N.; Samart, C. Data-driven prediction of biomass pyrolysis pathways toward phenolic and aromatic products. J. Environ. Chem. Eng. 2021, 9, 104836. [Google Scholar] [CrossRef]
  207. Wang, Y.; Li, D.; Zhao, D.; Fan, Y.; Bi, J.; Shan, R.; Yang, J.; Luo, B.; Yuan, H.; Ling, X.; et al. Calcium-Loaded Municipal Sludge-Biochar as an Efficient and Stable Catalyst for Biodiesel Production from Vegetable Oil. ACS Omega 2020, 5, 17471–17478. [Google Scholar] [CrossRef]
  208. Jahan, M.S.; Rahman, M.M.; Sutradhar, S.; Quaiyyum, M. Fractionation of rice straw for producing dissolving pulp in biorefinery concept. Nord. Pulp Pap. Res. J. 2015, 30, 562–567. [Google Scholar] [CrossRef]
  209. Zhao, C.; Yang, L.; Xing, S.; Luo, W.; Wang, Z.; Lv, P. Biodiesel production by a highly effective renewable catalyst from pyrolytic rice husk. J. Clean. Prod. 2018, 199, 772–780. [Google Scholar] [CrossRef]
  210. Rawat, S.; Wang, C.-T.; Lay, C.-H.; Hotha, S.; Bhaskar, T. Sustainable biochar for advanced electrochemical/energy storage applications. J. Energy Storage 2023, 63, 107115. [Google Scholar] [CrossRef]
  211. Wang, Y.; Zhang, Y.; Pei, L.; Ying, D.; Xu, X.; Zhao, L.; Jia, J.; Cao, X. Converting Ni-loaded biochars into supercapacitors: Implication on the reuse of exhausted carbonaceous sorbents. Sci. Rep. 2017, 7, 41523. [Google Scholar] [CrossRef]
  212. Sengottian, M.; Venkatachalam, C.D.; Ravichandran, S.R. Optimization of Alkali catalyzed hydrothermal carbonization of Prosopis juliflora woody biomass to biochar for copper and zinc adsorption and its application in supercapacitor. Int. J. Electrochem. Sci. 2022, 17, 220938. [Google Scholar] [CrossRef]
  213. Li, D.; Ma, J.; Xu, H.; Xu, X.; Qiu, H.; Cao, X.; Zhao, L. Recycling waste nickel-laden biochar to pseudo-capacitive material by hydrothermal treatment: Roles of nickel-carbon interaction. Carbon Res. 2022, 1, 16. [Google Scholar] [CrossRef]
Figure 1. Metal resources used in a smartphone.
Figure 1. Metal resources used in a smartphone.
Energies 17 04674 g001
Figure 2. Various modification techniques and their impacts on the properties of engineered biochar [17].
Figure 2. Various modification techniques and their impacts on the properties of engineered biochar [17].
Energies 17 04674 g002
Figure 4. Schematic representation of metal adsorption mechanism on engineered biochar.
Figure 4. Schematic representation of metal adsorption mechanism on engineered biochar.
Energies 17 04674 g004
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

Hasan, M.; Chakma, S.; Liang, X.; Sutradhar, S.; Kozinski, J.; Kang, K. Engineered Biochar for Metal Recycling and Repurposed Applications. Energies 2024, 17, 4674. https://doi.org/10.3390/en17184674

AMA Style

Hasan M, Chakma S, Liang X, Sutradhar S, Kozinski J, Kang K. Engineered Biochar for Metal Recycling and Repurposed Applications. Energies. 2024; 17(18):4674. https://doi.org/10.3390/en17184674

Chicago/Turabian Style

Hasan, Mehedi, Soumik Chakma, Xunjia Liang, Shrikanta Sutradhar, Janusz Kozinski, and Kang Kang. 2024. "Engineered Biochar for Metal Recycling and Repurposed Applications" Energies 17, no. 18: 4674. https://doi.org/10.3390/en17184674

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