Next Article in Journal
Thermochemical Conversion Processes as a Path for Sustainability of the Tire Industry: Carbon Black Recovery Potential in a Circular Economy Approach
Previous Article in Journal
Defect Properties of Li2NiGe3O8
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Adsorption of Lead from Aqueous Solution by Biochar: A Review

1
Tianjin Key Laboratory of Biomass Waste Utilization, School of Environmental Science and Engineering, Tianjin University, Tianjin 300072, China
2
School of Mechanical Engineering, Tianjin University of Commerce, Tianjin 300134, China
3
School of Science, Tibet University, Lhasa 850012, China
4
Tianjin Engineering Research Center for Organic Wastes Safe Disposal and Energy Utilization, School of Environmental Science and Engineering, Tianjin University, Tianjin 300072, China
*
Author to whom correspondence should be addressed.
Clean Technol. 2022, 4(3), 629-652; https://doi.org/10.3390/cleantechnol4030039
Submission received: 7 June 2022 / Revised: 29 June 2022 / Accepted: 30 June 2022 / Published: 8 July 2022

Abstract

:
Industrialization has resulted in the discharge of a certain amount of lead (Pb) from industrial sources causing damage risk to water quality and human health. Adsorption is an effective technique to remove Pb, and biochar has been widely studied owing to its advantages of low cost and high adsorption capacity. This review summarizes the influence of raw materials and modification methods on the adsorption capability of biochar. The adsorption isotherms and kinetics of biochar were summarized, and the main Pb removal mechanisms were studied systematically. In addition, the challenges and future perspectives were discussed comprehensively. It is expected that the review could provide insightful fundamentals for the experimental research and practical applications of biochar.

1. Introduction

Since the last century, the intensification of industrialization has put great pressure on the environment [1]. Automobile maintenance, tetraethyl manufacturing [2], refining process, and battery manufacturing process have produced a large amount of toxic lead (Pb), which will pollute the water [3]. Pb is a toxic and harmful element with high molecular weight, and it has the most abundant global distribution among heavy metals [2,4,5]. Besides, Pb is capable of accumulating in the human body, leading to serious health problems including cancer, anemia, renal insufficiency, permanent brain damage [6], and extreme mutations [7]. Recently, Pb pollution has become a major problem of water pollution in developing countries [8]. Therefore, removing Pb from wastewater is an important and urgent task to ensure human life and health [9].
In recent decades, a variety of techniques have been developed to remove Pb from wastewater, such as ion-exchange [10], membrane processes [11], chemical precipitation [12], filtration [13], electrocoagulation [14], coagulation [15], and adsorption [16,17,18,19]. Among these methods, adsorption is one of the most common removal technologies and has attracted wide attention owing to its advantages of low cost, simple process, and high removal efficiency [20,21,22]. Many adsorbents have been used to remove heavy metals from aqueous solutions. Traditional adsorbents may be unpopular due to their high production and regeneration costs [23], while lower-cost biochar is an effective material for removing Pb from water and has received extensive attention [24]. Adsorption features have been very recently exploited not only for environmental remediation by simple adsorption but also for sensing and catalytic applications [25,26,27].
Biochar, a porous carbonaceous material, is produced by the pyrolysis of biomass in an oxygen-limited environment [28]. Considering that biochar originates from the ‘Terra Preta’ soils in the Amazon region, the application of biochar for soil improvement was extensively investigated. Recently, biochars have been widely used in practical environments as air filters, fuels, building materials, and adsorbents [29]. Notably, biochar is proved to be a promising adsorbent for Pb removal in wastewater [30]. Common materials for biochar production include crop residues [31,32,33], wood, animal manure [34], and sewage sludge [35]. The adsorption capacity of biochar with different raw materials (sludge, cow dung, corn stalk, and willow) for heavy metals was explored, and it was found that corn stalk biochar had the best adsorption effect, which might be attributed to its larger specific surface area and excellent pore structure [36]. Biochar has a porous structure, large specific surface area, and abundant surface functional groups, which enables it to efficiently adsorb Pb from aqueous solutions [37]. However, the adsorption capacity of pristine biochar may be lower than that of conventional adsorbents [38]. For instance, the maximum Pb adsorption capacity of pristine straw biochar produced at 500 °C was 165 mg·g−1 [39]. Therefore, chemical or physical modifications are usually carried out to enhance the adsorption capacity of biochar [40,41]. For example, the N-doped MgO-modified corn cob biochar had an adsorption capacity of 1429 mg·g−1 for Pb in aqueous solution [42]. In addition, the micropore area and external specific surface area of ball-milled bone biochar were significantly increased, and the Pb adsorption capacity of ball-milled biochar was up to 558.8 mg·g−1 [43]. Hence, the modified biochars have considerable adsorption capacity of Pb, which can promote the application potential of biochar in wastewater treatment. At present, most of the relevant reviews are related to the summary of multiple heavy metals removed by biochar in wastewater, and there are relatively few summaries of single heavy metal. Considering the continual emergence of scientific advancement, it is necessary to systematically summarize the recent development of the Pb adsorption in wastewater by the pristine and modified biochars.
This paper focuses on the adsorption of Pb in wastewater by biochars. Important aspects such as raw materials, modification methods, and removal mechanisms are summarized. The overall purpose of this review is to provide a comprehensive and systematic analysis of the latest research results on biochar for Pb removal. The specific tasks of this work are as follows: (1) Collect relevant data about the Pb adsorption by biochar; (2) Clarify the influence of raw materials and modification methods on the Pb adsorption of biochar; (3) Explore the kinetics and isotherm models of Pb adsorption by biochar; (4) Summarize the adsorption capacity and removal mechanism of biochar; (5) Point out the shortcomings of existing research and provide guidance for Pb adsorption by biochar.

2. Feedstocks for Biochar Production

A large amount of solid waste, such as crop straw and sludge, has caused a series of environmental problems. The preparation of biochar can achieve the sustainable recycling of resources. According to the existing raw material data of biochar used for Pb adsorption, the raw materials were mainly divided into the following six categories: wood materials, agricultural wastes, animal residue, sludge, fruit peel, and other wastes. The proportions of the six raw materials are approximately 23.01%, 46.02%, 7.08%, 8.05%, 5.31%, and 9.73% (Figure 1).
Biochars prepared from different raw materials have different adsorption capabilities. The adsorption capacity of different raw material-derived biochars for Pb is presented in Figure 2. As for wood materials, the KMnO4-modified hickory wood biochar obtained a maximum Pb adsorption capacity of 153.1 mg·g−1 [44]. In addition, pine sawdust biochar had an adsorption capacity up to 606.00 mg·g−1 [45]. Therefore, wood could be made into biochar as an effective material for heavy metal adsorption.
As for agricultural wastes, Medulla tetrapanacis biochar had a high efficiency for the sorption of Pb (1031.23 mg·g−1) [46]. Similarly, corncob-to-xylose residue biochar could efficiently adsorb Pb in aqueous solution with a maximum adsorption capacity of 1429.00 mg·g−1 [42]. There was one study that showed that the excellent adsorption capacity of tobacco stem biochar was 2047.00 mg·g−1 according to the Langmuir model [47]. Agricultural waste biochar has attracted more attention than other raw materials (Figure 1). It is a key material for biochar research and industrial application.
As for animal residue, the biochar prepared from cow bone and optimized by ball milling technology had an adsorption capacity of 558.88 mg·g−1 [43]. At the same time, biochar derived from cow manure showed a maximum adsorption capacity of about 230.00 mg·g−1. The animal residue biochar has a certain adsorption capacity, but the relative research is not enough.
As for sludge, coagulation sludge biochar had a maximum adsorption capacity of 450.58 mg·g−1 [48]. As for fruit peel, there were six related papers, and the Pb adsorption capacities were, respectively, 86.96 mg·g−1 [49], 247.10 mg·g−1 [50], 134.00 mg·g−1 [51], 88.70 mg·g−1 [52], 742.00 mg·g−1 [53], and 359.00 mg·g−1 [54]. It could be seen that the fruit peel was a suitable raw material for biochar production. The adsorption capacities of different biochars are shown in Figure 2. In addition, the different adsorption capacities are ascribed not only to the nature of the feedstock but also to the pyrolysis conditions and modification techniques.

3. Modification Methodology

Many experimental results showed that the unmodified biochar had a weak adsorption capacity for heavy metals [55]. Thus, the biochar could be modified to further improve the adsorption capacity. The surface functional groups, pore structure, specific surface area, cation exchange capacity, and other physicochemical properties could be greatly improved after modification [31].
As shown in Figure 3, there are three main methods of biochar modification: biological modification, physical modification, and chemical modification. There are many studies on chemical modification, and it has a relatively good modification effect. Therefore, the chemical modification method will be analyzed and summarized, and the physical and biological modification methods will be introduced briefly.

3.1. Biological Modification

The well-developed pore structure and acid-base buffering capacity of biochar are beneficial to the growth of microorganisms. Biologically modified biochar is mainly used to load suitable microorganisms, which has achieved synergistic effects in both biochar adsorption and microbial adsorption, but suitable microorganisms need to be screened. Besides, anchoring proteins on the cell surface have metal-responsive motifs that can be specialized for metal adsorption and dissociation [56]. Wang et al. studied the Pb adsorption by biochar-loaded UV-mutant Bacillus subtilis, and a higher Pb adsorption capacity was achieved for the modified biochar [57]. The adsorption amount of Pb on the original biochar was about 150 mg/g, and the adsorption amount increased to 166.8 mg/g after loading bacteria, which might be attributed to the synergistic effect of biochar and bacteria. In addition, the microbial surface contains a large number of adsorption sites, which improves the adsorption effect of biochar on Pb [58]. However, considering that bacteria have strict requirements on reaction conditions, the Pb removal efficiency of bacterium-modified biochar would be affected during the treatment of industrial wastewater.

3.2. Physical Modification

Physical modification methods mainly aim to optimize the physical properties of biochar (pore size, specific surface area, etc.). Common physical modification methods include steam activation, air activation, and ball-milling. Steam activation is a relatively common physical modification method, which will increase the specific surface area and pore size of biochar after activation. For example, the steam-activated biochar had a higher adsorption capacity for Pb due to the large surface area [59]. Air oxidation also has a good modification effect. For instance, the Pb adsorption capacity increased from 2.50 to 44.00 mg·g−1 after the air-oxidized operation [60]. Furthermore, the biochar modified by ball milling has excellent heavy metal adsorption capacity, with the advantages of environmental friendliness and wide adaptability [61]. The bone-derived biochar showed a larger specific surface area and a better pore structure after ball-milling, which promoted the adsorption of Pb [43]. However, the separation of ball-milled biochar after Pb adsorption should be paid more attention during the practical application.

3.3. Chemical Modification

Chemical modification is mainly divided into the following categories: metal salt modification, acid modification, alkali modification, metal oxide modification, and organic matter modification.

3.3.1. Metal Salt Modification

The most used modifier is metal salt, which has an excellent modification effect on the pore size, the specific surface area, and the metal ions on the biochar surface. In addition, magnetic substances, such as iron, can enhance the magnetic properties of biochar, which helps to separate the biochar after adsorption.
The co-pyrolysis of magnesium-containing compounds, such as MgCl2, with biomass can increase the specific surface area, active sites, and surface charge of biochar [62]. For example, the adsorption capacity of MgCl2-modified biochar was 20 times higher than that of unmodified biochar [40]. The content of carboxyl, lactone, and phenol functional groups of modified biochar increased by 0.50, 0.86, and 0.06 mmol/g compared with the original biochar. The magnetization of biochar was a popular modification method. For example, magnetic cane biochar contained 18.40% oxygen, which could swell in water and allowed Pb to be adsorbed on the biochar [63]. Potassium permanganate (KMnO4) was a kind of metal salt with good modification effect owing to its strong oxidation property. KMnO4 modification promotes the formation of hierarchical structure, while increasing the specific surface area of biochar and optimizing the pore structure [64]. The Pb adsorption capacity of hickory wood biochar modified by KMnO4 was 153.10 mg·g−1, which was significantly higher than the original biochar, because the modified biochar had more surface oxygen-containing functional groups and larger surface area than the original biochar [44].

3.3.2. Acid Modification

Acid-modification is mainly used to increase the content of oxygen-containing functional groups to increase the adsorption ability of the biochar. Currently, phosphoric acid (H3PO4) and sulfuric acid (H2SO4) are the main acids for biochar modification.
H3PO4 could activate biochar to form −P=O and −P=OOH functional groups [65]. For example, orange peel biochar modified by H3PO4 had successfully introduced oxygen and phosphorus functional groups [52]. In addition, H3PO4 modification was beneficial to improve the phenolic group content and alkalinity of biochar, and the Pb adsorption capacity of the modified biochar was 5.41 times higher than unmodified biochar [66]. The carboxyl content on the surface of modified biochar increased significantly, which improved the adsorption capacity of biochar for heavy metals [67].
H2SO4 is corrosive and oxidizing, which contributes to the formation of pore- and oxygen-containing functional groups on biochar. H2SO4 will sulfonate with the aromatic hydrocarbons and introduce −SO3H on the biochar [68]. Li et al. found that the lignin biochar modified by H2SO4 could introduce highly acidic oxygen-containing groups, and the maximum Pb adsorption capacity reached 679.00 mg·g−1 [69]. Meanwhile, the biochar modified by H2SO4 had strong complex ability and high electronegativity, and a large amount of −OH and −COOH were generated on the surface of biochar during the sulfonation process. In addition, the Pb adsorption capacity for biochar was up to 191.07 mg·g−1 within 5 min [70].

3.3.3. Alkali Modification

Alkali modification facilitates the optimization of the pore structure and specific surface area of biochar. The biochar surface has more adsorption sites by improving the pore structure and increasing the specific surface area.
Common alkaline modification reagents for biochar are potassium hydroxide (KOH) and sodium hydroxide (NaOH). KOH modification method could increase the content of functional groups containing O and N on the surface of biochar, and promote ion exchange and surface complexation reaction [71]. For example, Herath et al. found that modified biochar had a higher adsorption capacity than unmodified biochar, because the specific surface area of KOH-activated biochar was doubled compared with the raw biochar [72]. Compared with the pristine biochar, NaOH-modified biochar has a better mesoporous structure and higher adsorption capacity [73]. In addition, Ding et al. demonstrated that the Pb adsorption capacity of NaOH modified biochar was 2.6–5.8 times higher than that of the original biochar, because cation exchange capacity and the surface area of the modified biochar were significantly improved [74].

3.3.4. Metal Oxide Modification

Metal oxides loaded on biochar can change the physicochemical properties of biochar, such as the elemental composition. The Pb adsorption capacity of biochar modified by metal oxide could be improved [75]. MgO-modified biochar or MgO-biochar composites have certain application prospects in the field of heavy metal adsorption [53]. At the same time, Zhang et al. found that the maximum Pb adsorption capacity of the MgO-coated biochar was 558.00 mg·g−1 when the maximum Mg content in the biochar was 10.10% [53]. On the one hand, MgO modification increases the surface area of biochar, and the higher surface area facilitates the adsorption/removal of Pb [76]. Additionally, the presence of the MgO coating significantly enhances the buffering capacity of the solution [77].

3.3.5. Organic Matter Modification

Organic matters were used to optimize the functional groups on biochar. The impacts of organic matter modification on the physicochemical properties of biochar are mainly reflected in two aspects. Firstly, the organic matter had an influence on the content of the functional groups. Chitosan is a surface modifier that can modify biochar to introduce a large number of oxygen-containing functional groups [78]. For example, Deng et al. proved that chitosan-pyromellitic dianhydride-modified biochar had more surface functional groups than the original biochar, so it had better Pb adsorption capacity [79]. The adsorption capacity of modified biochar for Pb improved by 10% compared with the original biochar, and the organic matter-modified biochar had a certain selectivity for heavy metal adsorption in wastewater. Furthermore, the biochar modified by thiourea could introduce C-O, C=O, and C=S, which could adsorb Pb through coordination and ion exchange, thereby thiourea-modified biochar increased the adsorption capacity of Pb(II) by 32% compared with the pristine biochar [80]. The second aspect is that organic matter influenced the elemental composition of biochar. The ammonium polyphosphate-modified biochar had more abundant N and P functional groups than unmodified biochar, so the Pb removal performance of modified biochar (723.60 mg·g−1) was improved compared to unmodified biochar (264.20 mg·g−1) [81].

3.4. Comparison of Different Modification Methods

Among these modification methods, metal oxide and metal salt modification achieved the most significant improvement in the Pb adsorption capacity of biochar, and the rise ranged from 42% to 600% [81,82,83]. The Pb adsorption capacity of Mg-modified coconut shell biochar was 20 times higher than that of the pristine biochar [40], and the N-doped coupling MgO-modified biochar derived from corncob-to-xylose residue exhibited a remarkable Pb adsorption capacity of 1429 mg/g [42]. Besides the enhancement in the surface morphology of biochar, the addition of metal oxide and metal salt could promote the ion exchange capacity of biochar, thus boosting the Pb adsorption capacity. However, the organic modification method has limited effect on improving the Pb adsorption capacity of biochar. The adsorption capacity of chitosan-modified biochar for Pb was only increased by 10% compared with the pristine biochar. Notably, the physical modification (e.g., air activation and ball-milling) has a good potential to optimize the Pb adsorption capacity of biochar by modifying the physicochemical properties. The Pb adsorption capacity of the shredded wood-derived biochar modified by mild air activation was 17 times higher than that of the raw biochar, since the modified biochar had a higher specific surface area and more abundant oxygen-containing functional groups on the surface [61]. The significant improvement in the micropore area and external specific surface area of bone biochar were also observed after ball-milling, and the Pb adsorption capacity of the ball-milled biochar increased by 64.6% compared with the pristine biochar [43]. In contrast, the improvement effect of bio-modified biochar on Pb adsorption is limited, even though the size distribution of pores in biochar provides suitable habitats for many microorganisms.
Hence, the specific chemical modification techniques such as metal oxide and metal salt modification show the greatest potential in terms of the improvement in Pb sorption capacity, while the negative effects of chemical modification including the high cost and the secondary pollution should be concerned during the practical application. Considering the simple operation procedure, good economic feasibility, and environmental friendliness of the physical modification, it could be an alternative technique to promote the Pb sorption capacity of biochar, and the enhancement efficiency could be further improved. In comparison with the aforementioned modification techniques, the biological modification showed a relatively poor potential in the practical application owing to the limited Pb adsorption capacity and strict application conditions.

4. Adsorption Isotherm and Kinetics

4.1. Adsorption Isotherm

The exploration of adsorption isotherm is very important for understanding the interaction between biochar and heavy metals. Freundlich ( q e = K F C e 1 n ), Langmuir ( q e = q m a x C e K L 1 + C e ), and Temkin ( q e = R T b × ln A T × C e ) models are often used to analyze experimental data and describe the adsorption equilibrium of Pb on biochar. According to the existing data of Pb adsorption by biochar in Figure 4, approximately 71.29% of the papers are in line with the Langmuir model [54,57,79,84,85,86,87,88,89,90], implying that the Pb adsorption on the surface of biochar is more like the single layer adsorption. For example, Wongrod et al. found that the adsorption isotherm of Pb by digestate biochar could be well described by the Langmuir isotherm model [91].
About 23.76% of the papers are in line with the Freundlich model [59,66,69,92,93,94,95,96,97]. The heterogeneity of the biochar surface made the adsorption process more complicated, and multi-layer adsorption might occur at sites with uneven energy distribution on the biochar [98]. Gayathri et al. used various equilibrium models to analyze the adsorption equilibrium data, and the results showed that the Freundlich isotherm was more suitable for explaining the adsorption characteristics [93]. Similarly, the adsorption of Pb in aqueous solution by nitrogen-doped magnesium oxide-modified biochar was best fitted to the Freundlich model [42]. A very small proportion of reported data conformed to other related adsorption models, such as Dubinin-Radushkevich model [75,99], Sips model [38,100], Redlich Peterson model [44], and Tempkin model.

4.2. Adsorption Kinetics

Kinetics model was used to study the adsorption rate. Three most popular kinetic models, pseudo-first-order kinetic model ( d q t d t = k 1 q e q t ), pseudo-second-order kinetic model ( d q t d t = k 2 q e q t 2 ) and Elovich model ( q t = 1 B   l n a b   + 1 B l n t ) have been used to study the adsorption of Pb on biochar [33,101].
According to the existing data of Pb adsorption by biochar in Figure 5, approximately 87.5% of the paper conforms to the pseudo-second-order kinetics model [48,54,57,59,66,69,79,84,86,88,90,100,102,103,104,105,106,107,108,109,110,111]. For example, Chen et al. applied the pseudo-first-order model and pseudo-second-order model to study the adsorption kinetics of Pb, and the results showed that the pseudo-second-order model had the best fit, revealing related adsorption processes mainly due to chemical adsorption [112]. Only 6.25% of the paper showed that the adsorption of Pb in water by biochar is in line with the pseudo-first-order kinetic model. Xue et al. found that the Pb adsorption by biochar had the best fit on the pseudo-first-order model [113]. It was assumed that the inorganic components in the biochar had low activity, and the chemical adsorption of Pb on the biochar was not obvious.
In addition, there are two papers showing that the adsorption process of biochar conformed to both the pseudo-first-order kinetics and pseudo-second-order kinetic models. On one hand, Wang et al. discovered that both the first-order and second-order kinetic equations fitted well in describing the adsorption kinetics of Pb, which indicated that the Pb adsorption on biochar might be controlled by multiple mechanisms [114]. On the other hand, the adsorption kinetic was more in line with the pseudo-first-order model when the Pb concentration was low, and it was more in line with the pseudo-second-order model when the Pb concentration was high [115]. Thus, the Pb adsorption might be attributed to the combined effect of physical and chemical adsorption processes [116].

5. Pb Removal Mechanisms by Biochar

The Pb removal mechanisms by biochar are analyzed according to Table 1. The removal mechanisms mainly include precipitation, complexation, ion exchange, electrostatic attraction, chemical bond adsorption, and physical adsorption. The corresponding mechanisms are summarized in Figure 6.

5.1. Precipitation

Chemical precipitation is widely used for heavy metal removal in industry, and it can be used effectively in a large temperature range with low operating costs. A large number of studies have shown that precipitation is one of the main Pb removal mechanisms [92,115,122,131]. For example, the surface of biochar had abundant Pb particles, which proved that precipitation was the removal mechanism [114]. The high removal capacity was mainly due to the formation of carbonate mineral precipitation [146]. The peak of CO32- shifted after the adsorption of Pb, and the production of the mineral PbCO3 indicated that Pb(II) was precipitated together with carbonate [152,153]. Besides, phosphates on the surface of biochar also played an important role in promoting the precipitation of Pb. Precipitation of Pb phosphate was an important mechanism for biochar to adsorb Pb(II) [81]. For instance, Pb5(PO4)3Cl [86], Pb5(PO4)3OH, and Pb3(PO4)2 [66] on the biochar proved that Pb phosphate precipitation was an important mechanism.
Furthermore, pH was an important factor affecting the precipitation of Pb ions. Pb(OH)2 precipitation formed at pH > 6.0, because there were local sites with high alkalinity on the surface of biochar [63,130]. The role of mineral precipitation increased significantly with the increase in pyrolysis temperature [154]. Precipitation of Pb2+ and carbonate minerals has been proved to be the main interaction of Pb removal [155]. Some quantitative analyses had also confirmed the predominant role of the precipitation mechanism. For instance, Wu et al. and Cheng et al. found that the contribution of precipitation accounted for more than 50% of the Pb adsorption capacity, because Pb could react with anions (e.g., CO32−, PO43−, and OH) released from biochar to form mineral precipitates [40,156].

5.2. Surface Complexation

Complexation is a process in which electrons interact with donors and acceptors to form various complexes. Complexation plays an important role in the Pb adsorption by biochar. The peak areas of C-O decreased from 31.29% to 21.95% after the adsorption of Pb, which represented that the C-O group combined with a large amount of Pb in solution through surface complexation during the adsorption process [156]. For example, the contribution of surface complexation in the removal mechanism accounted for 55.1% when Gao et al. studied the sorption mechanisms of Pb by rape straw biochar [157]. The content and types of oxygen-containing functional groups on the surface of biochar have a great relationship with the adsorption capacity. Characterization analyses such as X-ray photoelectron spectroscopy and X-ray diffraction showed that the removal mechanism of Pb by biochar was mainly the chemical complexation between Pb and oxygen-containing functional groups. The single bond O functional group might play a dominant role, because the single bond was observed to have a significant decrease in the strength of the −OH/C-O functional group after the Pb adsorption [23]. Compared with the carboxyl group, the hydroxyl group had smaller binding energy and stronger metal complexing ability [118]. Furthermore, the PbFe12O19 species might be attributed to the complexation between Pb2+ and −OH group on the biochar [158].
Related characterization proved that hydrated manganese oxide nanoparticles could combine with Pb ions through specific inner sphere complexation, and the negatively charged oxygen-containing groups on biochar were beneficial to adsorb Pb [96]. Some studies have shown similar results [126]. Besides, complexation is greatly affected by the preparation temperature of biochar, and the role of surface complexation decreases significantly with the increase in pyrolysis temperature [154].

5.3. Ion Exchange

Ion exchange is the effect or phenomenon of the exchange of ions in a solution with ions on a certain ion exchanger. It uses the exchange of ions in a solid ion exchanger with ions in a dilute solution to remove certain ions in the solution. Ion exchange is also one of the main mechanisms, mainly through ion exchange between the cations on the oxygen-containing functional groups and the Pb ions. The cation exchange capacity is a vital indicator of the Pb removal effect when the ion exchange is dominant [159].
The oxygen-containing functional group on the surface of biochar is an influencing factor of ion exchange. Besides, the ion exchange mechanism is often influenced by the number of acid sites on the biochar surface [160]. Zhou et al. found that Pb-O/Pb-O=C existed on the surface of banana peels biochar, which proved that the dominant mechanism might be ion exchange [54]. According to the isoelectric point of the carboxyl group and the phenolic hydroxyl group, it was speculated that the main removal mechanism was ion exchange through the carboxyl group [69]. In addition, the adsorption capacity of biochar did not change with the change of specific surface area after desorption, which also proved that the main mechanism was ion exchange [161]. Metal ions such as K+ and Mg2+, which were equivalent to 10.00% to 38.50% of the total adsorption capacity, were released into the solution, indicating that ion exchange was an important removal mechanism [154].
At the same time, quantitative analysis of some mechanisms also shows that ion exchange plays a major role. Wu et al. found that the contribution of ion exchange accounted for more than 53.31% [40]. The amount of Pb adsorbed by magnesium-modified biochar through mineral precipitation increased by 214.4 mg/g, and the ion exchange adsorption amount of modified biochar was 49 times that of the pristine biochar [40]. Ion exchange is greatly affected by the preparation temperature of biochar, and the role of ion exchange improved significantly with the increase in pyrolysis temperature [154]. In addition, the change of mineral ion exchange sites with pyrolysis temperature is the main factor affecting the adsorption of heavy metal ions [162].

5.4. Electrostatic Interaction

Electrostatic interaction occurs between positively charged Pb ions and negatively charged groups on biochar, especially oxygen-containing functional groups. The density of charges on the surface of biochar attracts ions with opposite charges and drives away ions with the same charge. For example, banana peel biochar had high electronegativity, and electrostatic attraction was the dominant force for the adsorption of Pb on biochar [50]. Besides, Mohubedu et al. found that the electrostatic interaction was the main removal mechanism [117]. Besides, Pb could combine with −COOH and −OH [52], amide oxygen atom, and amino N [163] on biochar by electrostatic interaction. Zeta potential value is an important indicator to explore the removal mechanism of electrostatic attraction. The high electronegativity of the surface of biochar led to the migration of Pb ions to the biochar [50]. Zeta potential was affected by pH, and the adsorption capacity of biochar was poor when pH < 5.3. The increase in charge neutralization activity improved the electrostatic interaction when 5.5 < pH < 7. The adsorption capacity decreased due to the formation of Pb(OH)2 precipitation when pH > 7.0. Furthermore, electrostatic interaction of deprotonated functional groups enhanced Pb adsorption [100].

5.5. Chemical Bond

The elements on the biochar can form chemical bonds with the Pb in the aqueous solution to achieve the purpose of removing Pb. The main Pb removal mechanism on biochar involves chemical binding adsorption [140]. The main types of chemical bonds are Pb(II)–π interaction [148] and π-π interactions. π-π interaction played a role in the Pb adsorption by biochar. The π-π interaction promoted the combination of Pb ions and C=C [164]. The shift of C=C and the change of the band after the adsorption of Pb ions indicated that the π-π effect played an important role [46]. Therefore, the deepening of the graphitization degree of biochar favored the adsorption of heavy metal Pb [46]. Besides, the adsorption capacity is consistent with the graphitization degree of biochar, because highly graphitized biochar can provide a large number of π donors [165].

5.6. Physical Adsorption

Physical adsorption also contributes to the adsorption of Pb by biochar. Physical adsorption was a major removal mechanism [46]. In addition, physical adsorption is mainly determined by the specific pore structure and surface area, and physical adsorption is a reversible process [166]. Interaction between biochar and Pb can be achieved through weak van der Waals forces [165]. Gayathri et al. prepared crop waste biochar to adsorb Pb in water [93]. Internal and external diffusion were the main separation mechanism for biochar to adsorb Pb. Besides, pore-filling played an important role in removing Pb [167]. In addition, Lee et al. found that boundary layer diffusion played a major role in the Pb adsorption by biochar [115]. SEM-EDX images showed that the surface of biochar contained Pb element crystal particles, so biochar could adsorb Pb through physical adsorption [121].

6. Problems and Future Perspectives

  • The current studies on the reusability of biochar and the interference of other ions in solution should be strengthened. Meanwhile, the effects of other heavy metals, such as copper and cadmium, on Pb adsorption need to be studied more carefully to facilitate the simultaneous removal of multiple heavy metals.
  • The current research on the Pb adsorption by biochar in wastewater has a large gap with the real Pb polluted water. Thus, the experimental research consistent with the actual polluted water should be promoted.
  • The research content of Pb removal mechanism by biochar remains basically unchanged. The main removal mechanisms include ion exchange, surface complexation, electrostatic adsorption, etc. There is a lack of feedback and improvement processes. Besides, the establishment of analytical models between the preparation conditions and the adsorption capacity might help design experiments more rationally.
  • The morphology and stability of Pb on the biochar surface need to be explored after adsorption, and the Pb leaching in the solution also needs to be studied in detail.

7. Conclusions

  • The results show that different raw materials had different physical and chemical properties. The most popular raw material of biochar for wastewater treatment was agricultural waste, accounting for about 50%.
  • The modification methods of biochar were mainly divided into physical and chemical modification. Among them, the chemical modification was the most widely used and had better effect. At the same time, there were many kinds of modifiers, and the most widely used modifier was metal salt.
  • According to the study of removal mechanism, the results show that the Pb adsorption by biochar mostly conforms to the Langmuir model, and most of the studies on adsorption kinetics are in line with the pseudo-second-order model.
  • Although there were many types of Pb removal mechanisms by biochar in wastewater, precipitation, complexation, ion exchange, and electrostatic attraction are the four main removal mechanisms.
Biochar is a suitable adsorbent for Pb removal in water, and it will be industrially applied widely in the future.

Author Contributions

Methodology, Formal analysis, Data Curation, Writing-Original draft preparation, C.W.; Resource, Writing-Review & Editing, X.W.; Conceptualization, Writing-Review & Editing, N.L.; Resource, Writing-Review & Editing, J.T.; Supervision, B.Y.; Data Curation, Writing-Review & Editing, X.C.; Writing—Review & Editing, Project administration, G.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology of the People’s Republic of China (2019YFD1100305) Junyu Tao, the National Natural Science Foundation of China (51878557) Junyu Tao, and Shenzhen Key Project of Basic Research (JCYJ20200109150210400) Guanyi Chen.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bogusz, A.; Oleszczuk, P.; Dobrowolski, R. Application of laboratory prepared and commercially available biochars to ad-sorption of cadmium, copper and zinc ions from water. Bioresour. Technol. 2015, 196, 540–549. [Google Scholar] [CrossRef] [PubMed]
  2. Loría, K.C.; Emiliani, J.; Bergara, C.D.; Herrero, M.S.; Salvatierra, L.M.; Pérez, L.M. Effect of daily exposure to Pb-contaminated water on Salvinia biloba physiology and phytoremediation performance. Aquat. Toxicol. 2019, 210, 158–166. [Google Scholar] [CrossRef] [PubMed]
  3. Hashem, A.; Fletcher, A.; Younis, H.; Mauof, H.; Abou-Okeil, A. Adsorption of Pb (II) ions from contaminated water by 1,2,3,4-butanetetracarboxylic acid-modified microcrystalline cellulose: Isotherms, kinetics, and thermodynamic studies. Int. J. Biol. Macromol. 2020, 164, 3193–3203. [Google Scholar] [CrossRef] [PubMed]
  4. Li, Y.H.; Wang, S.; Wei, J.; Zhang, X.; Xu, C.; Luan, Z.; Wu, D.; Wei, B. Lead adsorption on carbon nanotubes. Chem. Phys. Lett. 2002, 357, 263–266. [Google Scholar] [CrossRef]
  5. Xie, Y.; Yuan, X.; Wu, Z.; Zeng, G.; Jiang, L.; Peng, X.; Li, H. Adsorption behavior and mechanism of Mg/Fe layered double hy-droxide with Fe3O4-carbon spheres on the removal of Pb (II) and Cu (II). J. Colloid Interface Sci. 2019, 536, 440–455. [Google Scholar] [CrossRef]
  6. Qu, J.; Meng, Q.; Lin, X.; Han, W.; Jiang, Q.; Wang, L.; Hu, Q.; Zhang, L.; Zhang, Y. Microwave-assisted synthesis of beta-cyclodextrin functionalized celluloses for enhanced removal of Pb (II) from water: Adsorptive performance and mechanism exploration. Sci. Total Environ. 2020, 752, 141854. [Google Scholar] [CrossRef]
  7. Wang, L.; Lei, T.; Ren, Z.; Jiang, X.; Yang, X.; Bai, H.; Wang, S. Fe3O4@PDA@MnO2 core-shell nanocomposites for sensitive elec-trochemical detection of trace Pb (II) in water. J. Electroanal. Chem. 2020, 864, 114065. [Google Scholar] [CrossRef]
  8. Dai, K.; Liu, G.; Xu, W.; Deng, Z.; Wu, Y.; Zhao, C.; Zhang, Z. Judicious fabrication of bifunctionalized graphene oxide/MnFe2O4 magnetic nanohybrids for enhanced removal of Pb (II) from water. J. Colloid Interface Sci. 2020, 579, 815–822. [Google Scholar] [CrossRef]
  9. Awual, M.R.; Hasan, M.M.; Islam, A.; Rahman, M.M.; Asiri, A.M.; Khaleque, M.A.; Sheikh, M.C. Offering an innovative composited material for effective lead (II) monitoring and removal from polluted water. J. Clean. Prod. 2019, 231, 214–223. [Google Scholar] [CrossRef]
  10. Goyal, P.; Tiwary, C.S.; Misra, S.K. Ion exchange based approach for rapid and selective Pb (II) removal using iron oxide dec-orated metal organic framework hybrid. J. Environ. Manag. 2020, 277, 111469. [Google Scholar] [CrossRef]
  11. Huang, L.; Wu, B.; Wu, Y.; Yang, Z.; Yuan, T.; Alhassan, S.I.; Yang, W.; Wang, H.; Zhang, L. Porous and flexible membrane derived from ZIF-8-decorated hyphae for outstanding adsorption of Pb2+ ion. J. Colloid Interface Sci. 2020, 565, 465–473. [Google Scholar] [CrossRef] [PubMed]
  12. Németh, G.; Mlinárik, L.; Török, Á. Adsorption and chemical precipitation of lead and zinc from contaminated solutions in porous rocks: Possible application in environmental protection. J. Afr. Earth Sci. 2016, 122, 98–106. [Google Scholar] [CrossRef]
  13. Zhang, S.; Shi, Q.; Christodoulatos, C.; Korfiatis, G.; Meng, X. Adsorptive filtration of lead by electrospun PVA/PAA nanofiber membranes in a fixed-bed column. Chem. Eng. J. 2019, 370, 1262–1273. [Google Scholar] [CrossRef]
  14. Al-Shannag, M.; Al-Qodah, Z.; Bani-Melhem, K.; Qtaishat, M.R.; Alkasrawi, M. Heavy metal ions removal from metal plating wastewater using electrocoagulation: Kinetic study and process performance. Chem. Eng. J. 2015, 260, 749–756. [Google Scholar] [CrossRef]
  15. Pang, F.M.; Kumar, P.; Teng, T.T.; Omar, A.K.M.; Wasewar, K.L. Removal of lead, zinc and iron by coagulation–flocculation. J. Taiwan Inst. Chem. Eng. 2011, 42, 809–815. [Google Scholar] [CrossRef]
  16. Yogeshwaran, V.; Priya, A. Adsorption of lead ion concentration from the aqueous solution using tobacco leaves. Mater. Today Proc. 2020, 37, 489–496. [Google Scholar] [CrossRef]
  17. Kuganathan, N.; Anurakavan, S.; Abiman, P.; Iyngaran, P.; Gkanas, E.I.; Chroneos, A. Adsorption of lead on the surfaces of pristine and B, Si and N-doped graphene. Phys. B Condens. Matter 2020, 600, 412639. [Google Scholar] [CrossRef]
  18. Menezes, J.M.C.; Bento, A.M.D.S.; da Silva, J.H.; Filho, F.J.D.P.; da Costa, J.G.M.; Coutinho, H.D.M.; Teixeira, R.N.P. Equilibrium, kinetics and thermodynamics of lead (II) adsorption in bioadsorvent composed by Caryocar coriaceum Wittm barks. Chemosphere 2020, 261, 128144. [Google Scholar] [CrossRef]
  19. Shen, M.; Song, B.; Zeng, G.; Zhang, Y.; Teng, F.; Zhou, C. Surfactant changes lead adsorption behaviors and mechanisms on microplastics. Chem. Eng. J. 2020, 405, 126989. [Google Scholar] [CrossRef]
  20. Park, J.-H.; Ok, Y.S.; Kim, S.-H.; Cho, J.-S.; Heo, J.-S.; Delaune, R.D.; Seo, D.-C. Competitive adsorption of heavy metals onto sesame straw biochar in aqueous solutions. Chemosphere 2016, 142, 77–83. [Google Scholar] [CrossRef]
  21. Nadeem, R.; Zafar, M.N.; Afzal, A.; Hanif, M.A.; Saeed, R. Potential of NaOH pretreated Mangifera indica waste biomass for the mitigation of Ni (II) and Co (II) from aqueous solutions. J. Taiwan Inst. Chem. Eng. 2014, 45, 967–972. [Google Scholar] [CrossRef]
  22. Li, X.; Wang, C.; Zhang, J.; Liu, J.; Liu, B.; Chen, G. Preparation and application of magnetic biochar in water treatment: A critical review. Sci. Total Environ. 2019, 711, 134847. [Google Scholar] [CrossRef] [PubMed]
  23. Ahmed, W.; Mehmood, S.; Núñez-Delgado, A.; Ali, S.; Qaswar, M.; Shakoor, A.; Mahmood, M.; Chen, D.-Y. Enhanced adsorption of aqueous Pb (II) by modified biochar produced through pyrolysis of watermelon seeds. Sci. Total Environ. 2021, 784, 147136. [Google Scholar] [CrossRef] [PubMed]
  24. Li, Y.; Yu, H.; Liu, L.; Yu, H. Application of co-pyrolysis biochar for the adsorption and immobilization of heavy metals in contaminated environmental substrates. J. Hazard. Mater. 2021, 420, 126655. [Google Scholar] [CrossRef] [PubMed]
  25. Lee, E.-S.; Lee, J.; Kim, H.-J.; Kim, Y.-P. Fluorogenic Aptasensors with Small Molecules. Chemosensors 2021, 9, 54. [Google Scholar] [CrossRef]
  26. Oliveira, G.A.; Gevaerd, A.; Mangrich, A.S.; Marcolino-Junior, L.H.; Bergamini, M.F. Biochar obtained from spent coffee grounds: Evaluation of adsorption properties and its application in a voltammetric sensor for lead (II) ions. Microchem. J. 2021, 165, 106114. [Google Scholar] [CrossRef]
  27. Diao, Z.-H.; Dong, F.-X.; Yan, L.; Chen, Z.-L.; Guo, P.-R.; Xia, X.-J.; Chu, W. A new insight on enhanced Pb (II) removal by sludge biochar catalyst coupling with ultrasound irradiation and its synergism with phenol removal. Chemosphere 2020, 263, 128287. [Google Scholar] [CrossRef]
  28. Iamsaard, K.; Weng, C.-H.; Yen, L.-T.; Tzeng, J.-H.; Poonpakdee, C.; Lin, Y.-T. Adsorption of metal on pineapple leaf biochar: Key affecting factors, mechanism identification, and regeneration evaluation. Bioresour. Technol. 2021, 344, 126131. [Google Scholar] [CrossRef]
  29. Uday, V.; Harikrishnan, P.; Deoli, K.; Zitouni, F.; Mahlknecht, J.; Kumar, M. Current trends in production, morphology, and real-world environmental applications of biochar for the promotion of sustainability. Bioresour. Technol. 2022, 359, 127467. [Google Scholar] [CrossRef]
  30. Inyang, M.; Gao, B.; Yao, Y.; Xue, Y.; Zimmerman, A.R.; Pullammanappallil, P.; Cao, X. Removal of heavy metals from aqueous solution by biochars derived from anaerobically digested biomass. Bioresour. Technol. 2012, 110, 50–56. [Google Scholar] [CrossRef]
  31. Luo, M.; Lin, H.; Li, B.; Dong, Y.; He, Y.; Wang, L. A novel modification of lignin on corncob-based biochar to enhance removal of cadmium from water. Bioresour. Technol. 2018, 259, 312–318. [Google Scholar] [CrossRef]
  32. Li, B.; Yang, L.; Wang, C.-Q.; Zhang, Q.-P.; Liu, Q.-C.; Li, Y.-D.; Xiao, R. Adsorption of Cd (II) from aqueous solutions by rape straw biochar derived from different modification processes. Chemosphere 2017, 175, 332–340. [Google Scholar] [CrossRef] [PubMed]
  33. Tan, Z.; Wang, Y.; Kasiulienė, A.; Huang, C.; Ai, P. Cadmium removal potential by rice straw-derived magnetic biochar. Clean Technol. Environ. Policy 2016, 19, 761–774. [Google Scholar] [CrossRef]
  34. Xu, X.; Cao, X.; Zhao, L.; Wang, H.; Yu, H.; Gao, B. Removal of Cu, Zn, and Cd from aqueous solutions by the dairy ma-nure-derived biochar. Environ. Sci. Pollut. Res. 2013, 20, 358–368. [Google Scholar] [CrossRef] [PubMed]
  35. Zuo, W.-Q.; Chen, C.; Cui, H.-J.; Fu, M.-L. Enhanced removal of Cd (II) from aqueous solution using CaCO3 nanoparticle mod-ified sewage sludge biochar. RSC Adv. 2017, 7, 16238–16243. [Google Scholar] [CrossRef] [Green Version]
  36. Chen, G.; Wang, C.; Tian, J.; Liu, J.; Ma, Q.; Liu, B.; Li, X. Investigation on cadmium ions removal from water by different raw materials-derived biochars. J. Water Process Eng. 2020, 35, 101223. [Google Scholar] [CrossRef]
  37. Tan, X.; Liu, Y.; Zeng, G.; Wang, X.; Hu, X.; Gu, Y.; Yang, Z. Application of biochar for the removal of pollutants from aqueous solutions. Chemosphere 2015, 125, 70–85. [Google Scholar] [CrossRef] [PubMed]
  38. Mahdi, Z.; El Hanandeh, A.; Yu, Q.J. Preparation, characterization and application of surface modified biochar from date seed for improved lead, copper, and nickel removal from aqueous solutions. J. Environ. Chem. Eng. 2019, 7, 103379. [Google Scholar] [CrossRef]
  39. Nzediegwu, C.; Naeth, M.A.; Chang, S.X. Lead (II) adsorption on microwave-pyrolyzed biochars and hydrochars depends on feedstock type and production temperature. J. Hazard. Mater. 2021, 412, 125255. [Google Scholar] [CrossRef]
  40. 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. 2020, 754, 142150. [Google Scholar] [CrossRef]
  41. Tao, Q.; Chen, Y.; Zhao, J.; Li, B.; Li, Y.; Tao, S.; Li, M.; Li, Q.; Xu, Q.; Li, Y.; et al. Enhanced Cd removal from aqueous solution by biologically modified biochar derived from digestion residue of corn straw silage. Sci. Total Environ. 2019, 674, 213–222. [Google Scholar] [CrossRef] [PubMed]
  42. Shang, H.; Li, Y.; Liu, J.; Wan, Y.; Feng, Y.; Yu, Y. Preparation of nitrogen doped magnesium oxide modified biochar and its sorption efficiency of lead ions in aqueous solution. Bioresour. Technol. 2020, 314, 123708. [Google Scholar] [CrossRef]
  43. Xiao, J.; Hu, R.; Chen, G. Micro-nano-engineered nitrogenous bone biochar developed with a ball-milling technique for high-efficiency removal of aquatic Cd (II), Cu (II) and Pb (II). J. Hazard. Mater. 2019, 387, 121980. [Google Scholar] [CrossRef]
  44. Wang, H.; Gao, B.; Wang, S.; Fang, J.; Xue, Y.; Yang, K. Removal of Pb (II), Cu (II), and Cd (II) from aqueous solutions by biochar derived from KMnO4 treated hickory wood. Bioresour. Technol. 2015, 197, 356–362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Sellaoui, L.; Wang, H.; Badawi, M.; Bonilla-Petriciolet, A.; Chen, Z. Synergistic adsorption of Pb2+ and CrO42− on an engineered biochar highlighted by statistical physical modeling. J. Mol. Liq. 2020, 312, 113483. [Google Scholar] [CrossRef]
  46. Zhang, L.; Li, W.; Cao, H.; Hu, D.; Chen, X.; Guan, Y.; Tang, J.; Gao, H. Ultra-efficient sorption of Cu2+ and Pb2+ ions by light biochar derived from Medulla tetrapanacis. Bioresour. Technol. 2019, 291, 121818. [Google Scholar] [CrossRef] [PubMed]
  47. Zhou, Z.; Xu, Z.; Feng, Q.; Yao, D.; Yu, J.; Wang, D.; Lv, S.; Liu, Y.; Zhou, N.; Zhong, M.-E. Effect of pyrolysis condition on the ad-sorption mechanism of lead, cadmium and copper on tobacco stem biochar. J. Clean. Prod. 2018, 187, 996–1005. [Google Scholar] [CrossRef]
  48. Luo, X.; Shen, M.; Huang, Z.; Chen, Z.; Chen, Z.; Lin, B.; Cui, L. Efficient removal of Pb (II) through recycled biochar-mineral composite from the coagulation sludge of swine wastewater. Environ. Res. 2020, 190, 110014. [Google Scholar] [CrossRef]
  49. Abdelhafez, A.A.; Li, J. Removal of Pb (II) from aqueous solution by using biochars derived from sugar cane bagasse and orange peel. J. Taiwan Inst. Chem. Eng. 2016, 61, 367–375. [Google Scholar] [CrossRef]
  50. Ahmad, Z.; Gao, B.; Mosa, A.; Yu, H.; Yin, X.; Bashir, A.; Ghoveisi, H.; Wang, S. Removal of Cu (II), Cd (II) and Pb (II) ions from aqueous solutions by biochars derived from potassium-rich biomass. J. Clean. Prod. 2018, 180, 437–449. [Google Scholar] [CrossRef]
  51. Jin, Q.; Wang, Z.; Feng, Y.; Kim, Y.-T.; Stewart, A.C.; O’Keefe, S.F.; Neilson, A.P.; He, Z.; Huang, H. Grape pomace and its secondary waste management: Biochar production for a broad range of lead (Pb) removal from water. Environ. Res. 2020, 186, 109442. [Google Scholar] [CrossRef] [PubMed]
  52. Zhao, T.; Yao, Y.; Li, D.; Wu, F.; Zhang, C.; Gao, B. Facile low-temperature one-step synthesis of pomelo peel biochar under air atmosphere and its adsorption behaviors for Ag (I) and Pb (II). Sci. Total Environ. 2018, 640–641, 73–79. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, J.; Hou, D.; Shen, Z.; Jin, F.; O’Connor, D.; Pan, S.; Ok, Y.S.; Tsang, D.C.W.; Bolan, N.S.; Alessi, D.S. Effects of excessive im-pregnation, magnesium content, and pyrolysis temperature on MgO-coated watermelon rind biochar and its lead removal capacity. Environ. Res. 2020, 183, 109152. [Google Scholar] [CrossRef]
  54. Zhou, N.; Chen, H.; Xi, J.; Yao, D.; Zhou, Z.; Tian, Y.; Lu, X. Biochars with excellent Pb (II) adsorption property produced from fresh and dehydrated banana peels via hydrothermal carbonization. Bioresour. Technol. 2017, 232, 204–210. [Google Scholar] [CrossRef]
  55. Qiu, B.; Tao, X.; Wang, H.; Li, W.; Ding, X.; Chu, H. Biochar as a low-cost adsorbent for aqueous heavy metal removal: A review. J. Anal. Appl. Pyrolysis 2021, 155, 105081. [Google Scholar] [CrossRef]
  56. Wang, Y.; Selvamani, V.; Yoo, I.-K.; Kim, T.W.; Hong, S.H. A Novel Strategy for the Microbial Removal of Heavy Metals: Cell-surface Display of Peptides. Biotechnol. Bioprocess Eng. 2021, 26, 1–9. [Google Scholar] [CrossRef]
  57. Wang, T.; Sun, H.; Ren, X.; Li, B.; Mao, H. Adsorption of heavy metals from aqueous solution by UV-mutant Bacillus subtilis loaded on biochars derived from different stock materials. Ecotoxicol. Environ. Saf. 2018, 148, 285–292. [Google Scholar] [CrossRef]
  58. Liu, Z.; Xu, Z.; Xu, L.; Buyong, F.; Chay, T.C.; Li, Z.; Cai, Y.; Hu, B.; Zhu, Y.; Wang, X. Modified biochar: Synthesis and mechanism for removal of environmental heavy metals. Carbon Res. 2022, 1, 1–21. [Google Scholar] [CrossRef]
  59. Kwak, J.-H.; Islam, S.; Wang, S.; Messele, S.A.; Naeth, M.A.; El-Din, M.G.; Chang, S.X. Biochar properties and lead (II) adsorption capacity depend on feedstock type, pyrolysis temperature, and steam activation. Chemosphere 2019, 231, 393–404. [Google Scholar] [CrossRef]
  60. Bardestani, R.; Roy, C.; Kaliaguine, S. The effect of biochar mild air oxidation on the optimization of lead (II) adsorption from wastewater. J. Environ. Manag. 2019, 240, 404–420. [Google Scholar] [CrossRef]
  61. Lyu, H.; Gao, B.; He, F.; Ding, C.; Tang, J.; Crittenden, J.C. Ball-Milled Carbon Nanomaterials for Energy and Environmental Applications. ACS Sustain. Chem. Eng. 2017, 5, 9568–9585. [Google Scholar] [CrossRef]
  62. Li, A.; Ge, W.; Liu, L.; Qiu, G. Preparation, adsorption performance and mechanism of MgO-loaded biochar in wastewater treatment: A review. Environ. Res. 2022, 212, 113341. [Google Scholar] [CrossRef] [PubMed]
  63. Mohan, D.; Singh, P.; Sarswat, A.; Steele, P.H.; Pittman, C.U., Jr. Lead sorptive removal using magnetic and nonmagnetic fast pyrolysis energy cane biochars. J. Colloid Interface Sci. 2015, 448, 238–250. [Google Scholar] [CrossRef] [PubMed]
  64. Xu, J.; Zhang, Y.; Li, B.; Fan, S.; Xu, H.; Guan, D.-X. Improved adsorption properties of tetracycline on KOH/KMnO4 modified biochar derived from wheat straw. Chemosphere 2022, 296, 133981. [Google Scholar] [CrossRef]
  65. Dai, X.; Nhung, N.; Hamza, M.F.; Guo, Y.; Chen, L.; He, C.; Ningc, S.; Weic, Y.; Dodbibad, G.; Fujita, T. Selective adsorption and recovery of scandium from red mud leachate by using phosphoric acid pre-treated pitaya peel biochar. Sep. Purif. Technol. 2022, 292, 121043. [Google Scholar] [CrossRef]
  66. Chen, H.; Li, W.; Wang, J.; Xu, H.; Liu, Y.; Zhang, Z.; Li, Y.; Zhang, Y. Adsorption of cadmium and lead ions by phosphoric ac-id-modified biochar generated from chicken feather: Selective adsorption and influence of dissolved organic matter. Bioresour. Technol. 2019, 292, 121948. [Google Scholar] [CrossRef]
  67. He, X.; Hong, Z.-N.; Jiang, J.; Dong, G.; Liu, H.; Xu, R.-K. Enhancement of Cd (II) adsorption by rice straw biochar through oxidant and acid modifications. Environ. Sci. Pollut. Res. 2021, 28, 42787–42797. [Google Scholar] [CrossRef]
  68. Xiong, X.; Yu, I.K.; Chen, S.S.; Tsang, D.C.; Cao, L.; Song, H.; Kwon, E.E.; Ok, Y.S.; Zhang, S.; Poon, C.S. Sulfonated biochar as acid catalyst for sugar hydrolysis and dehydration. Catal. Today 2018, 314, 52–61. [Google Scholar] [CrossRef]
  69. Li, Y.; Wang, F.; Miao, Y.; Mai, Y.; Li, H.; Chen, X.; Chen, J. A lignin-biochar with high oxygen-containing groups for adsorbing lead ion prepared by simultaneous oxidization and carbonization. Bioresour. Technol. 2020, 307, 123165. [Google Scholar] [CrossRef]
  70. 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. 2020, 750, 141545. [Google Scholar] [CrossRef]
  71. Chen, H.; Yang, X.; Liu, Y.; Lin, X.; Wang, J.; Zhang, Z.; Li, N.; Li, Y.; Zhang, Y. KOH modification effectively enhances the Cd and Pb adsorption performance of N-enriched biochar derived from waste chicken feathers. Waste Manag. 2021, 130, 82–92. [Google Scholar] [CrossRef] [PubMed]
  72. Herath, A.; Layne, C.A.; Perez, F.; Hassan, E.B.; Pittman, C.U., Jr.; 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]
  73. Mu, Y.; Ma, H. NaOH-modified mesoporous biochar derived from tea residue for methylene Blue and Orange II removal. Chem. Eng. Res. Des. 2021, 167, 129–140. [Google Scholar] [CrossRef]
  74. Ding, Z.; Hu, X.; Wan, Y.; Wang, S.; Gao, B. Removal of lead, copper, cadmium, zinc, and nickel from aqueous solutions by alkali-modified biochar: Batch and column tests. J. Ind. Eng. Chem. 2016, 33, 239–245. [Google Scholar] [CrossRef] [Green Version]
  75. Cruz, G.J.F.; Mondal, D.; Rimaycuna, J.; Soukup, K.; Gómez, M.M.; Solis, J.L.; Lang, J. Agrowaste derived biochars impregnated with ZnO for removal of arsenic and lead in water. J. Environ. Chem. Eng. 2020, 8, 103800. [Google Scholar] [CrossRef]
  76. Shen, Z.; Jin, F.; Wang, F.; McMillan, O.; Al-Tabbaa, A. Sorption of lead by Salisbury biochar produced from British broadleaf hardwood. Bioresour. Technol. 2015, 193, 553–556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Shen, Z.; Hou, D.; Zhao, B.; Xu, W.; Ok, Y.S.; Bolan, N.S.; Alessi, D.S. Stability of heavy metals in soil washing residue with and without biochar addition under accelerated ageing. Sci. Total Environ. 2018, 619–620, 185–193. [Google Scholar] [CrossRef]
  78. Vasconcelos, H.; Guibal, E.; Laus, R.; Vitali, L.; Fávere, V.T. Competitive adsorption of Cu (II) and Cd (II) ions on spray-dried chitosan loaded with Reactive Orange 16. Mater. Sci. Eng. C 2009, 29, 613–618. [Google Scholar] [CrossRef]
  79. Deng, J.; Liu, Y.; Liu, S.; Zeng, G.; Tan, X.; Huang, B.; Tang, X.; Wang, S.; Hua, Q.; Yan, Z. Competitive adsorption of Pb (II), Cd (II) and Cu (II) onto chitosan-pyromellitic dianhydride modified biochar. J. Colloid Interface Sci. 2017, 506, 355–364. [Google Scholar] [CrossRef]
  80. Liu, J.; Huang, Z.; Chen, Z.; Sun, J.; Gao, Y.; Wu, E. Resource utilization of swine sludge to prepare modified biochar adsorbent for the efficient removal of Pb (II) from water. J. Clean. Prod. 2020, 257, 120322. [Google Scholar] [CrossRef]
  81. Fan, Y.; Wang, H.; Deng, L.; Wang, Y.; Kang, D.; Li, C.; Chen, H. Enhanced adsorption of Pb (II) by nitrogen and phosphorus co-doped biochar derived from Camellia oleifera shells. Environ. Res. 2020, 191, 110030. [Google Scholar] [CrossRef] [PubMed]
  82. Zhao, Y.; Zhang, R.; Liu, H.; Li, M.; Chen, T.; Chen, D.; Zou, X.; Frost, R.L. Green preparation of magnetic biochar for the effective accumulation of Pb (II): Performance and mechanism. Chem. Eng. J. 2019, 375, 122011. [Google Scholar] [CrossRef]
  83. Jia, Y.; Zhang, Y.; Fu, J.; Yuan, L.; Li, Z.; Liu, C.; Zhao, D.; Wang, X. A novel magnetic biochar/MgFe-layered double hydroxides composite removing Pb2+ from aqueous solution: Isotherms, kinetics and thermodynamics. Colloids Surf. A Physicochem. Eng. Asp. 2019, 567, 278–287. [Google Scholar] [CrossRef]
  84. Yang, W.; Wang, Z.; Song, S.; Han, J.; Chen, H.; Wang, X.; Sun, R.; Cheng, J. Adsorption of copper (II) and lead (II) from seawater using hydrothermal biochar derived from Enteromorpha. Mar. Pollut. Bull. 2019, 149, 110586. [Google Scholar] [CrossRef] [PubMed]
  85. Nnadozie, E.C.; Ajibade, P.A. Adsorption, kinetic and mechanistic studies of Pb (II) and Cr (VI) ions using APTES functionalized magnetic biochar. Microporous Mesoporous Mater. 2020, 309, 110573. [Google Scholar] [CrossRef]
  86. 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] [PubMed]
  87. Zhang, W.; Du, W.; Wang, F.; Xu, H.; Zhao, T.; Zhang, H.; Ding, Y.; Zhu, W. Comparative study on Pb2+ removal from aqueous solutions using biochars derived from cow manure and its vermicompost. Sci. Total Environ. 2020, 716, 137108. [Google Scholar] [CrossRef]
  88. Lee, M.-E.; Park, J.H.; Chung, J.W. Comparison of the lead and copper adsorption capacities of plant source materials and their biochars. J. Environ. Manag. 2019, 236, 118–124. [Google Scholar] [CrossRef]
  89. Ni, B.-J.; Huang, Q.-S.; Wang, C.; Ni, T.-Y.; Sun, J.; Wei, W. Competitive adsorption of heavy metals in aqueous solution onto biochar derived from anaerobically digested sludge. Chemosphere 2018, 219, 351–357. [Google Scholar] [CrossRef]
  90. Wang, Y.-Y.; Liu, Y.-X.; Lu, H.-H.; Yang, R.-Q.; Yang, S.-M. Competitive adsorption of Pb (II), Cu (II), and Zn (II) ions onto hy-droxyapatite-biochar nanocomposite in aqueous solutions. J. Solid State Chem. 2018, 261, 53–61. [Google Scholar] [CrossRef]
  91. Wongrod, S.; Simon, S.; Guibaud, G.; Lens, P.N.; Pechaud, Y.; Huguenot, D.; van Hullebusch, E.D. Lead sorption by biochar produced from digestates: Consequences of chemical modification and washing. J. Environ. Manag. 2018, 219, 277–284. [Google Scholar] [CrossRef] [PubMed]
  92. Wu, Z.; Chen, X.; Yuan, B.; Fu, M.-L. A facile foaming-polymerization strategy to prepare 3D MnO2 modified biochar-based porous hydrogels for efficient removal of Cd (II) and Pb (II). Chemosphere 2019, 239, 124745. [Google Scholar] [CrossRef] [PubMed]
  93. Gayathri, R.; Gopinath, K.; Kumar, P.S. Adsorptive separation of toxic metals from aquatic environment using agro waste biochar: Application in electroplating industrial wastewater. Chemosphere 2020, 262, 128031. [Google Scholar] [CrossRef]
  94. Zhang, C.; Shan, B.; Tang, W.; Zhu, Y. Comparison of cadmium and lead sorption by Phyllostachys pubescens biochar produced under a low-oxygen pyrolysis atmosphere. Bioresour. Technol. 2017, 238, 352–360. [Google Scholar] [CrossRef] [PubMed]
  95. He, J.; Song, Y.; Chen, J.P. Development of a novel biochar/PSF mixed matrix membrane and study of key parameters in treatment of copper and lead contaminated water. Chemosphere 2017, 186, 1033–1045. [Google Scholar] [CrossRef] [PubMed]
  96. Wan, S.; Wu, J.; Zhou, S.; Wang, R.; Gao, B.; He, F. Enhanced lead and cadmium removal using biochar-supported hydrated manganese oxide (HMO) nanoparticles: Behavior and mechanism. Sci. Total Environ. 2017, 616–617, 1298–1306. [Google Scholar] [CrossRef] [PubMed]
  97. Zhang, S.; Du, Q.; Sun, Y.; Song, J.; Yang, F.; Tsang, D.C.W. Fabrication of L-cysteine stabilized alpha-FeOOH nanocomposite on porous hydrophilic biochar as an effective adsorbent for Pb2+ removal. Sci. Total Environ. 2020, 720, 137415. [Google Scholar] [CrossRef] [PubMed]
  98. Fan, R.; Chen, C.-L.; Lin, J.-Y.; Tzeng, J.-H.; Huang, C.-P.; Dong, C. Adsorption characteristics of ammonium ion onto hydrous biochars in dilute aqueous solutions. Bioresour. Technol. 2018, 272, 465–472. [Google Scholar] [CrossRef]
  99. Ding, Z.; Wan, Y.; Hu, X.; Wang, S.; Zimmerman, A.R.; Gao, B. Sorption of lead and methylene blue onto hickory biochars from different pyrolysis temperatures: Importance of physicochemical properties. J. Ind. Eng. Chem. 2016, 37, 261–267. [Google Scholar] [CrossRef] [Green Version]
  100. Zhou, X.; Liu, Y.; Zhou, J.; Guo, J.; Ren, J.; Zhou, F. Efficient removal of lead from aqueous solution by urea-functionalized magnetic biochar: Preparation, characterization and mechanism study. J. Taiwan Inst. Chem. Eng. 2018, 91, 457–467. [Google Scholar] [CrossRef]
  101. Nandi, B.; Goswami, A.; Purkait, M. Adsorption characteristics of brilliant green dye on kaolin. J. Hazard. Mater. 2009, 161, 387–395. [Google Scholar] [CrossRef] [PubMed]
  102. Liu, L.; Huang, Y.; Zhang, S.; Gong, Y.; Su, Y.; Cao, J.; Hu, H. Adsorption characteristics and mechanism of Pb (II) by agricultural waste-derived biochars produced from a pilot-scale pyrolysis system. Waste Manag. 2019, 100, 287–295. [Google Scholar] [CrossRef] [PubMed]
  103. Deng, J.; Li, X.; Liu, Y.; Zeng, G.; Liang, J.; Song, B.; Wei, X. Alginate-modified biochar derived from Ca (II)-impregnated biomass: Excellent anti-interference ability for Pb (II) removal. Ecotoxicol. Environ. Saf. 2018, 165, 211–218. [Google Scholar] [CrossRef] [PubMed]
  104. Amin, M.; Alazba, A.; Shafiq, M. Application of biochar derived from date palm biomass for removal of lead and copper ions in a batch reactor: Kinetics and isotherm scrutiny. Chem. Phys. Lett. 2019, 722, 64–73. [Google Scholar] [CrossRef]
  105. Yang, Y.; Wei, Z.; Zhang, X.; Chen, X.; Yue, D.; Yin, Q.; Xiao, L.; Yang, L. Biochar from Alternanthera philoxeroides could remove Pb (II) efficiently. Bioresour. Technol. 2014, 171, 227–232. [Google Scholar] [CrossRef]
  106. Kalinke, C.; Mangrich, A.S.; Marcolino-Junior, L.H.; Bergamini, M.F. Biochar prepared from castor oil cake at different tem-peratures: A voltammetric study applied for Pb2+, Cd2+ and Cu2+ ions preconcentration. J. Hazard. Mater. 2016, 318, 526–532. [Google Scholar] [CrossRef]
  107. Zhang, Y.; Cao, B.; Zhao, L.; Sun, L.; Gao, Y.; Li, J.; Yang, F. Biochar-supported reduced graphene oxide composite for adsorption and coadsorption of atrazine and lead ions. Appl. Surf. Sci. 2018, 427, 147–155. [Google Scholar] [CrossRef]
  108. Yan, Y.; Zhang, L.; Wang, Y.; Wang, X.; Wang, S.; Li, Q.; Liu, X.; Xu, Y.; Yang, J.; Bolan, N. Clanis bilineata larvae skin-derived biochars for immobilization of lead: Sorption isotherm and molecular mechanism. Sci. Total Environ. 2020, 704, 135251. [Google Scholar] [CrossRef]
  109. Kołodyńska, D.; Krukowska, J.; Thomas, P. Comparison of sorption and desorption studies of heavy metal ions from biochar and commercial active carbon. Chem. Eng. J. 2017, 307, 353–363. [Google Scholar] [CrossRef]
  110. Li, M.; Wei, D.; Liu, T.; Liu, Y.; Yan, L.; Wei, Q.; Du, B.; Xu, W. EDTA functionalized magnetic biochar for Pb (II) removal: Adsorption performance, mechanism and SVM model prediction. Sep. Purif. Technol. 2019, 227, 115696. [Google Scholar] [CrossRef]
  111. Shen, Z.; Hou, D.; Jin, F.; Shi, J.; Fan, X.; Tsang, D.C.; Alessi, D.S. Effect of production temperature on lead removal mechanisms by rice straw biochars. Sci. Total Environ. 2018, 655, 751–758. [Google Scholar] [CrossRef] [PubMed]
  112. Chen, Y.-D.; Ho, S.-H.; Wang, D.; Wei, Z.-S.; Chang, J.-S.; Ren, N.-Q. Lead removal by a magnetic biochar derived from persulfate-ZVI treated sludge together with one-pot pyrolysis. Bioresour. Technol. 2018, 247, 463–470. [Google Scholar] [CrossRef] [PubMed]
  113. Xue, Y.; Wang, C.; Hu, Z.; Zhou, Y.; Xiao, Y.; Wang, T. Pyrolysis of sewage sludge by electromagnetic induction: Biochar properties and application in adsorption removal of Pb (II), Cd (II) from aqueous solution. Waste Manag. 2019, 89, 48–56. [Google Scholar] [CrossRef]
  114. Wang, S.; Gao, B.; Li, Y.; Mosa, A.; Zimmerman, A.R.; Ma, L.Q.; Harris, W.G.; Migliaccio, K.W. Manganese oxide-modified biochars: Preparation, characterization, and sorption of arsenate and lead. Bioresour. Technol. 2015, 181, 13–17. [Google Scholar] [CrossRef]
  115. Lee, X.J.; Lee, L.Y.; Hiew, B.Y.Z.; Gan, S.; Thangalazhy-Gopakumar, S.; Ng, H.K. Multistage optimizations of slow pyrolysis synthesis of biochar from palm oil sludge for adsorption of lead. Bioresour. Technol. 2017, 245, 944–953. [Google Scholar] [CrossRef] [PubMed]
  116. Ho, Y.S.; Porter, J.F.; McKay, G. Equilibrium Isotherm Studies for the Sorption of Divalent Metal Ions onto Peat: Copper, Nickel and Lead Single Component Systems. Water Air Soil Pollut. 2002, 141, 1–33. [Google Scholar] [CrossRef]
  117. Mohubedu, R.P.; Diagboya, P.N.; Abasi, C.Y.; Dikio, E.D.; Mtunzi, F. Magnetic valorization of biomass and biochar of a typical plant nuisance for toxic metals contaminated water treatment. J. Clean. Prod. 2018, 209, 1016–1024. [Google Scholar] [CrossRef]
  118. Zhao, M.; Dai, Y.; Zhang, M.; Feng, C.; Qin, B.; Zhang, W.; Zhao, N.; Li, Y.; Ni, Z.; Xu, Z.; et al. Mechanisms of Pb and/or Zn adsorption by different biochars: Biochar characteristics, stability, and binding energies. Sci. Total Environ. 2020, 717, 136894. [Google Scholar] [CrossRef]
  119. Qu, J.; Dong, M.; Wei, S.; Meng, Q.; Hu, L.; Hu, Q.; Wang, L.; Han, W.; Zhang, Y. Microwave-assisted one pot synthesis of β-cyclodextrin modified biochar for concurrent removal of Pb (II) and bisphenol a in water. Carbohydr. Polym. 2020, 250, 117003. [Google Scholar] [CrossRef]
  120. Wang, C.; Wang, H. Pb (II) sorption from aqueous solution by novel biochar loaded with nano-particles. Chemosphere 2018, 192, 1–4. [Google Scholar] [CrossRef]
  121. Zhou, R.; Zhang, M.; Li, J.; Zhao, W. Optimization of preparation conditions for biochar derived from water hyacinth by using response surface methodology (RSM) and its application in Pb2+ removal. J. Environ. Chem. Eng. 2020, 8, 104198. [Google Scholar] [CrossRef]
  122. Wang, C.; Wang, H.; Cao, Y. Pb (II) sorption by biochar derived from Cinnamomum camphora and its improvement with ultrasound-assisted alkali activation. Colloids Surfaces A Physicochem. Eng. Asp. 2018, 556, 177–184. [Google Scholar] [CrossRef]
  123. Yan, Y.; Sarkar, B.; Zhou, L.; Zhang, L.; Li, Q.; Yang, J.; Bolan, N. Phosphorus-rich biochar produced through bean-worm skin waste pyrolysis enhances the adsorption of aqueous lead. Environ. Pollut. 2020, 266, 115177. [Google Scholar] [CrossRef] [PubMed]
  124. 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] [PubMed]
  125. Faheem; Yu, H.; Liu, J.; Shen, J.; Sun, X.; Li, J.; Wang, L. Preparation of MnOx-loaded biochar for Pb2+ removal: Adsorption performance and possible mechanism. J. Taiwan Inst. Chem. Eng. 2016, 66, 313–320. [Google Scholar] [CrossRef]
  126. Lin, S.; Huang, W.; Yang, H.; Sun, S.; Yu, J. Recycling application of waste long-root Eichhornia crassipes in the heavy metal removal using oxidized biochar derived as adsorbents. Bioresour. Technol. 2020, 314, 123749. [Google Scholar] [CrossRef]
  127. Wang, S.-Y.; Tang, Y.-K.; Chen, C.; Wu, J.-T.; Huang, Z.; Mo, Y.-Y.; Zhang, K.-X.; Chen, J.-B. Regeneration of magnetic biochar derived from eucalyptus leaf residue for lead (II) removal. Bioresour. Technol. 2015, 186, 360–364. [Google Scholar] [CrossRef]
  128. Lu, H.; Zhang, W.; Yang, Y.; Huang, X.; Wang, S.; Qiu, R. Relative distribution of Pb2+ sorption mechanisms by sludge-derived biochar. Water Res. 2012, 46, 854–862. [Google Scholar] [CrossRef]
  129. Neeli, S.T.; Ramsurn, H.; Ng, C.Y.; Wang, Y.; Lu, J. Removal of Cr (VI), As (V), Cu (II), and Pb (II) using cellulose biochar supported iron nanoparticles: A kinetic and mechanistic study. J. Environ. Chem. Eng. 2020, 8, 103886. [Google Scholar] [CrossRef]
  130. Shi, J.; Fan, X.; Tsang, D.C.; Wang, F.; Shen, Z.; Hou, D.; Alessi, D.S. Removal of lead by rice husk biochars produced at different temperatures and implications for their environmental utilizations. Chemosphere 2019, 235, 825–831. [Google Scholar] [CrossRef]
  131. Tan, G.; Wu, Y.; Liu, Y.; Xiao, D. Removal of Pb (II) ions from aqueous solution by manganese oxide coated rice straw biochar A low-cost and highly effective sorbent. J. Taiwan Inst. Chem. Eng. 2018, 84, 85–92. [Google Scholar] [CrossRef]
  132. Wang, H.; Gao, B.; Fang, J.; Ok, Y.S.; Xue, Y.; Yang, K.; Cao, X. Engineered biochar derived from eggshell-treated biomass for removal of aqueous lead. Ecol. Eng. 2017, 121, 124–129. [Google Scholar] [CrossRef]
  133. Zhang, J.; Shao, J.; Jin, Q.; Zhang, X.; Yang, H.; Chen, Y.; Zhang, S.; Chen, H. Effect of deashing on activation process and lead adsorption capacities of sludge-based biochar. Sci. Total Environ. 2020, 716, 137016. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, T.; Zhu, X.; Shi, L.; Li, J.; Li, S.; Lü, J.; Li, Y. Efficient removal of lead from solution by celery-derived biochars rich in alkaline minerals. Bioresour. Technol. 2017, 235, 185–192. [Google Scholar] [CrossRef]
  135. Wang, H.; Wang, S.; Chen, Z.; Zhou, X.; Wang, J.; Chen, Z. Engineered biochar with anisotropic layered double hydroxide nanosheets to simultaneously and efficiently capture Pb2+ and CrO42− from electroplating wastewater. Bioresour. Technol. 2020, 306, 123118. [Google Scholar] [CrossRef]
  136. Zhao, H.T.; Ma, S.; Zheng, S.Y.; Han, S.W.; Yao, F.X.; Wang, X.Z.; Wang, S.S.; Feng, K. Beta-cyclodextrin functionalized biochars as novel sorbents for high-performance of Pb2+ removal. J. Hazard. Mater. 2019, 362, 206–213. [Google Scholar] [CrossRef]
  137. Yin, W.; Dai, D.; Hou, J.; Wang, S.; Wu, X.; Wang, X. Hierarchical porous biochar-based functional materials derived from bio-waste for Pb (II) removal. Appl. Surf. Sci. 2019, 465, 297–302. [Google Scholar] [CrossRef]
  138. Li, C.; Zhang, L.; Gao, Y.; Li, A. Facile synthesis of nano ZnO/ZnS modified biochar by directly pyrolyzing of zinc contaminated corn stover for Pb (II), Cu (II) and Cr (VI) removals. Waste Manag. 2018, 79, 625–637. [Google Scholar] [CrossRef]
  139. Chandraiah, M.R. Facile synthesis of zero valent iron magnetic biochar composites for Pb (II) removal from the aqueous medium. Alex. Eng. J. 2016, 55, 619–625. [Google Scholar] [CrossRef] [Green Version]
  140. Niu, Z.; Feng, W.; Huang, H.; Wang, B.; Chen, L.; Miao, Y.; Su, S. Green synthesis of a novel Mn–Zn ferrite/biochar composite from waste batteries and pine sawdust for Pb2+ removal. Chemosphere 2020, 252, 126529. [Google Scholar] [CrossRef]
  141. Jellali, S.; Diamantopoulos, E.; Haddad, K.; Anane, M.; Durner, W.; Mlayah, A. Lead removal from aqueous solutions by raw sawdust and magnesium pretreated biochar: Experimental investigations and numerical modelling. J. Environ. Manag. 2016, 180, 439–449. [Google Scholar] [CrossRef] [PubMed]
  142. Lian, W.; Yang, L.; Joseph, S.; Shi, W.; Bian, R.; Zheng, J.; Li, L.; Shan, S.; Pan, G. Utilization of biochar produced from invasive plant species to efficiently adsorb Cd (II) and Pb (II). Bioresour. Technol. 2020, 317, 124011. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, Q.; Wang, B.; Lee, X.; Lehmann, J.; Gao, B. Sorption and desorption of Pb (II) to biochar as affected by oxidation and pH. Sci. Total Environ. 2018, 634, 188–194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Li, J.; Zheng, L.; Wang, S.L.; Wu, Z.; Wu, W.; Niazi, N.K.; Shaheen, S.M.; Rinklebe, J.; Bolan, N.; Ok, Y.S.; et al. Sorption mech-anisms of lead on silicon-rich biochar in aqueous solution: Spectroscopic investigation. Sci. Total Environ. 2019, 672, 572–582. [Google Scholar] [CrossRef]
  145. El-Banna, M.F.; Mosa, A.; Gao, B.; Yin, X.; Ahmad, Z.; Wang, H. Sorption of lead ions onto oxidized bagasse-biochar mitigates Pb-induced oxidative stress on hydroponically grown chicory: Experimental observations and mechanisms. Chemosphere 2018, 208, 887–898. [Google Scholar] [CrossRef]
  146. Chang, J.; Zhang, H.; Cheng, H.; Yan, Y.; Chang, M.; Cao, Y.; Huang, F.; Zhang, G.; Yan, M. Spent Ganoderma lucidum substrate derived biochar as a new bio-adsorbent for Pb2+/Cd2+ removal in water. Chemosphere 2020, 241, 125121. [Google Scholar] [CrossRef]
  147. 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]
  148. Wu, J.; Wang, T.; Zhang, Y.; Pan, W.P. The distribution of Pb (II)/Cd (II) adsorption mechanisms on biochars from aqueous solution: Considering the increased oxygen functional groups by HCl treatment. Bioresour. Technol. 2019, 291, 121859. [Google Scholar] [CrossRef]
  149. Zama, E.F.; Zhu, Y.-G.; Reid, B.J.; Sun, G.-X. The role of biochar properties in influencing the sorption and desorption of Pb (II), Cd (II) and As (III) in aqueous solution. J. Clean. Prod. 2017, 148, 127–136. [Google Scholar] [CrossRef] [Green Version]
  150. Pan, J.; Gao, B.; Wang, S.; Guo, K.; Xu, X.; Yue, Q. Waste-to-resources: Green preparation of magnetic biogas residues-based biochar for effective heavy metal removals. Sci. Total Environ. 2020, 737, 140283. [Google Scholar] [CrossRef]
  151. Xu, X.; Hu, X.; Ding, Z.; Chen, Y.; Gao, B. Waste-art-paper biochar as an effective sorbent for recovery of aqueous Pb (II) into value-added PbO nanoparticles. Chem. Eng. J. 2017, 308, 863–871. [Google Scholar] [CrossRef]
  152. Trakal, L.; Bingöl, D.; Pohořelý, M.; Hruška, M.; Komárek, M. Geochemical and spectroscopic investigations of Cd and Pb sorption mechanisms on contrasting biochars: Engineering implications. Bioresour. Technol. 2014, 171, 442–451. [Google Scholar] [CrossRef] [PubMed]
  153. Xu, Y.; Bai, T.; Li, Q.; Yang, H.; Yan, Y.; Sarkar, B.; Lam, S.S.; Bolan, N. Influence of pyrolysis temperature on the characteristics and lead (II) adsorption capacity of phosphorus-engineered poplar sawdust biochar. J. Anal. Appl. Pyrolysis 2020, 154, 105010. [Google Scholar] [CrossRef]
  154. Yang, Z.; Hou, J.; Wu, J.; Miao, L. The effect of carbonization temperature on the capacity and mechanisms of Pb (II) adsorption by microalgae residue-derived biochar. Ecotoxicol. Environ. Saf. 2021, 225, 112750. [Google Scholar] [CrossRef] [PubMed]
  155. Wang, Z.; Liu, G.; Zheng, H.; Li, F.; Ngo, H.H.; Guo, W.; Liu, C.; Chen, L.; Xing, B. Investigating the mechanisms of biochar’s removal of lead from solution. Bioresour. Technol. 2014, 177, 308–317. [Google Scholar] [CrossRef] [PubMed]
  156. Cheng, S.; Liu, Y.; Xing, B.; Qin, X.; Zhang, C.; Xia, H. Lead and cadmium clean removal from wastewater by sustainable biochar derived from poplar saw dust. J. Clean. Prod. 2021, 314, 128074. [Google Scholar] [CrossRef]
  157. Gao, R.; Xiang, L.; Hu, H.; Fu, Q.; Zhu, J.; Liu, Y.; Huang, G. High-efficiency removal capacities and quantitative sorption mechanisms of Pb by oxidized rape straw biochars. Sci. Total Environ. 2019, 699, 134262. [Google Scholar] [CrossRef]
  158. Gong, H.; Chi, J.; Ding, Z.; Zhang, F.; Huang, J. Removal of lead from two polluted soils by magnetic wheat straw biochars. Ecotoxicol. Environ. Saf. 2020, 205, 111132. [Google Scholar] [CrossRef]
  159. Yang, X.; Wan, Y.; Zheng, Y.; He, F.; Yu, Z.; Huang, J.; Wang, H.; Ok, Y.S.; Jiang, Y.; Gao, B. Surface functional groups of carbon-based adsorbents and their roles in the removal of heavy metals from aqueous solutions: A critical review. Chem. Eng. J. 2019, 366, 608–621. [Google Scholar] [CrossRef]
  160. Dhouibi, N.; Binous, H.; Dhaouadi, H.; Dridi-Dhaouadi, S. Hydrodistillation residues of Centaurea nicaeensis plant for copper and zinc ions removal: Novel concept for waste re-use. J. Clean. Prod. 2020, 261, 121106. [Google Scholar] [CrossRef]
  161. Zhang, J.; Hu, X.; Zhang, K.; Xue, Y. Desorption of calcium-rich crayfish shell biochar for the removal of lead from aqueous solutions. J. Colloid Interface Sci. 2019, 554, 417–423. [Google Scholar] [CrossRef] [PubMed]
  162. Gong, H.; Chi, J. Contribution of different mechanisms to Pb2+ and Cd2+ sorption on magnetic wheat straw biochars: Impact of pyrolysis temperature and DOM in biochar. J. Environ. Chem. Eng. 2022, 10, 107851. [Google Scholar] [CrossRef]
  163. Fu, C.; Zhang, H.; Xia, M.; Lei, W.; Wang, F. The single/co-adsorption characteristics and microscopic adsorption mechanism of biochar-montmorillonite composite adsorbent for pharmaceutical emerging organic contaminant atenolol and lead ions. Ecotoxicol. Environ. Saf. 2019, 187, 109763. [Google Scholar] [CrossRef]
  164. Zhou, B.; Wang, Z.; Shen, D.; Shen, F.; Wu, C.; Xiao, R. Low cost earthworm manure-derived carbon material for the adsorption of Cu2+ from aqueous solution: Impact of pyrolysis temperature. Ecol. Eng. 2016, 98, 189–195. [Google Scholar] [CrossRef]
  165. Cui, X.; Fang, S.; Yao, Y.; Li, T.; Ni, Q.; Yang, X.; He, Z. Potential mechanisms of cadmium removal from aqueous solution by Canna indica derived biochar. Sci. Total Environ. 2016, 562, 517–525. [Google Scholar] [CrossRef] [PubMed]
  166. Liu, T.; Lawluvy, Y.; Shi, Y.; Ighalo, J.O.; He, Y.; Zhang, Y.; Yap, P.-S. Adsorption of cadmium and lead from aqueous solution using modified biochar: A review. J. Environ. Chem. Eng. 2021, 10, 106502. [Google Scholar] [CrossRef]
  167. Ramola, S.; Belwal, T.; Li, C.J.; Wang, Y.Y.; Lu, H.H.; Yang, S.M.; Zhou, C.H. Improved lead removal from aqueous solution using novel porous bentonite—And calcite-biochar composite. Sci. Total Environ. 2019, 709, 136171. [Google Scholar] [CrossRef]
Figure 1. The proportion of raw materials for biochar to adsorb Pb in water.
Figure 1. The proportion of raw materials for biochar to adsorb Pb in water.
Cleantechnol 04 00039 g001
Figure 2. The Pb adsorption capacity of biochars derived from different raw materials.
Figure 2. The Pb adsorption capacity of biochars derived from different raw materials.
Cleantechnol 04 00039 g002
Figure 3. Biochar modification methods for adsorbing Pb.
Figure 3. Biochar modification methods for adsorbing Pb.
Cleantechnol 04 00039 g003
Figure 4. The proportion of different isotherm models of Pb adsorption by biochar in water.
Figure 4. The proportion of different isotherm models of Pb adsorption by biochar in water.
Cleantechnol 04 00039 g004
Figure 5. Proportion of different kinetic models of Pb adsorption by biochar in water.
Figure 5. Proportion of different kinetic models of Pb adsorption by biochar in water.
Cleantechnol 04 00039 g005
Figure 6. Mechanisms of Pb sorption by biochar.
Figure 6. Mechanisms of Pb sorption by biochar.
Cleantechnol 04 00039 g006
Table 1. Information about the adsorption of heavy metal Pb in water by biochar.
Table 1. Information about the adsorption of heavy metal Pb in water by biochar.
FeedstockModifying Agent Temperature (°C)/Residence Time (h)Concentration Range (mg/L)Contact Time (h)Isotherm ModelKinetic ModelAdsorption Capacity (mg/g)Adsorption MechanismRef.
Douglas firKOH900–1000, 10 s25–10001LangmuirSecond140.00---[72]
SludgePersulfate-zvi600/1.50–1008LangmuirSecond180.00Electrostatic attraction, electrostatic outer sphere complexation, ion exchange, reduction[112]
Raw sewage sludgePotassium hydroxide350/15 min10–100080 min---Second106.00Precipitation[91]
Quercus roburFeCl3 and FeCl2250/425–1504LangmuirSecond63.60Electrostatic interactions[117]
Pine woodMnCl2·4H2O600/11–30024Langmuir and FreundlichFirst and second47.05Precipitation[114]
Rice straw---550/20–1 mmol·L−172Langmuir---0.85
176.12
Chemical complexation[118]
Cow bone---600/20–12024LangmuirSecond558.88Surface complexation, cation exchange, chemical precipitation, electrostatic interaction, and cation-π bonding[43]
Rice huskβ-cyclodextrin300/210–3002LangmuirSecond240.13Electrostatic attraction and complexation[119]
Palm oil sludge---500/0.50–1504FreundlichFirst, second-21.76Boundary layer diffusion[115]
Sewage sludge---500/0.50–3206LangmuirFirst41.20---[113]
Fallen leafFe2+/Fe3+450/10–10002LangmuirSecond146.84---[120]
Water hyacinth---433/2.650–10006FreundlichSecond251.39Precipitation, electrostatic adsorption, surface physical adsorption, ion exchange, and complexation of functional groups.[121]
Cinnamomum camphorUltrasound-assisted alkali450/150–10006LangmuirSecond98.33Electrostatic attraction and surface complexation[122]
Bean-worm skin waste---500/420–300200LangmuirSecond62.00Chemisorption and precipitation[123]
Corn stalksNanoscale zero-valent iron, KOH800/210–2008LangmuirSecond480.90Reduction reaction, complexation, and co-precipitation[124]
Rice huskManganese oxide800/315–2501.5LangmuirSecond86.50Electron density reduction in π-bond aromatic moieties due to the addition of -COOH[125]
Corncob-to-xylose residueNitrogen doped magnesium oxide400/20–50024FreundlichSecond1429.00Ion exchange, precipitation and complexation[42]
Date seedHCl550/3------Sips model---188.55---[38]
Long-root Eichhornia crassipes---350/20 min1.0 mmol·L−150 minLangmuirSecond118.10Complexation[126]
Eucalypts leaf residueZnCl2, FeCl3 and FeSO4700/20–10048------52.40---[127]
Sludge---550/2200–10004FreundlichSecond30.88Co-precipitation, complexation[128]
Microcrystalline celluloseIron nanoparticles10000–150024TempkinSecond17.30Reduction[129]
Banana peels---600/20–60048FreundlichSecond247.10Electrostatic attraction[50]
Rice husk---700/10–60048FreundlichSecond26.70Precipitation[130]
Hickory woodNaOH600/22–10024Langmuir---53.60---[74]
Orange peel---5006–2236LangmuirSecond86.96Ion exchange and surface precipitation[49]
Rice strawKMnO4420/41.0 mmol·L−18LangmuirSecond305.25Complexation[131]
Hickory woodKMnO4600/10–10024Redlich Peterson Richie n-th-order153.10Surface adsorption mechanisms[44]
Swine sludgeThiourea300/0.50–10048LangmuirSecond145.00Ion exchange[80]
Hickory---350/55–25024Dubinin---16.30Cation exchange[99]
ShellFeCl3·6H2O, EDTA200/850–50012LangmuirSecond129.31Electrostatic interaction and chemical complexation[110]
Bamboo---450/32–50048LangmuirFirst261.10---[132]
Tobacco stem---700/20–100012Langmuir---2047.00Precipitation[47]
Bagasse---300/2---24FreundlichIntraparticle diffusion---Ion exchange, precipitation[95]
SludgePotassium acetate700/15–30024LangmuirSecond49.47Complexation, surface precipitation[133]
Watermelon rindMgO600/150 mmol·L−124------742.00---[53]
Palm fiberFeSO4·7H2O and FeCl3·6H2O400/225–30024Sips modelSecond188.18Electrostatic interaction, ion exchange, and complexation[100]
Celery---500/360–40024------304.00Precipitation, cation exchange, and surface complexation[134]
Coagulation sludgeMgCl2⋅6H2O, MgFe3⋅6H2O500/40–14024LangmuirSecond488.78Ion exchange, electrostatic interaction[48]
Pinewood
sawdust
Al(NO3)3·9H2O, MgSO4·7H2O350/110–50024LangmuirSecond591.20Complexation and electrostatic interaction[135]
C. Oleifera shellsPolyammonium phosphate550/1100–20006LangmuirSecond723.60Surface complexation[81]
Rice strawsβ–cyclodextrin500/450–50004LangmuirSecond131.24Ion exchange and complexation [136]
Soybean cake---700/2602.5 Second133.60Cation exchange, precipitation[137]
Shredded woodMild air 475/0.250–500---------44.00 Precipitation[60]
Pomelo peelH3PO4250/210–200024LangmuirSecond88.70Surface complexation and electrostatic interactions[52]
Date seedHCl550/3------Sips---188.55---[38]
Peanut shellHydrated manganese oxide400/10–3024FreundlichSecond330.00Complexation[96]
Corn stalks FeSO4·7H2O, L-cysteine120/1210–15010FreundlichSecond103.04 Electrostatic attraction [97]
Axonopus compressusSulfuric acid180/0.50–20015 minLangmuirSecond191.07Complexation and ion exchange[70]
Corn stoverZnO/ZnS600/15–10048FreundlichSecond135.80Ion exchange, inner sphere complexation[138]
Eucalyptus globules barkZero valent iron7500–2002LangmuirSecond60.80---[139]
Grape pomace---700/250–30024LangmuirSecond134.00Electrostatic attraction, cation exchange, complexation[51]
Wheat strawNatural hematite800/25–150024FreundlichSecond196.91Precipitation and the surface complexation[83]
Pine sawdustMagnetic ferrite200/85–1006LangmuirSecond99.50Chemical binding adsorption, electrostatic attraction, and ion exchange[140]
Medulla tetrapanacis---700/150–40024LangmuirSecond1031.23Complexation, precipitation, π-π interactions, ion exchange[46]
Raw sawdustMagnesium600/120–6001.5------202.20Ion exchange[141]
Douglas firKOH900–1000/
10 s
25–10001LangmuirSecond140.00---[72]
SludgePersulfate-zvi600/1.50–1008LangmuirSecond180.00Electrostatic attraction, electrostatic outer sphere complexation, ion exchange, reduction[112]
Raw sewage sludgePotassium hydroxide350/15
min
10–100080 min---Second106.00Precipitation[91]
Quercus roburFeCl3 and FeCl2250/425–1504LangmuirSecond63.60Electrostatic interactions[117]
Pine woodMnCl2·4H2O600/11–30024Langmuir and FreundlichFirst and second47.05Precipitation[101]
Rice straw---550/20–1 mmol·L−172Langmuir---0.85
176.12
Chemical complexation[118]
Rice huskβ-cyclodextrin300/210–3002LangmuirSecond240.13Electrostatic attraction and complexation[119]
Water hyacinth---433/2.650–10006FreundlichSecond251.39Precipitation, electrostatic adsorption, surface physical adsorption, ion exchange, and complexation of functional groups.[121]
Palm oil sludge---500/0.50–1504FreundlichFirst, second-21.76Boundary layer diffusion[115]
Cinnamomum camphorUltrasound-assisted alkali450/150–10006LangmuirSecond98.33Electrostatic attraction and surface complexation[122]
Fallen leafFe2+/Fe3+450/10–10002LangmuirSecond146.84---[120]
Bean-worm skin waste---500/420–300200LangmuirSecond62.00Chemisorption and precipitation[123]
Corn stalksNanoscale zero-valent iron, KOH800/210–2008LangmuirSecond480.90Reduction reaction, complexation, and co-precipitation[124]
Rice huskManganese oxide800/315–2501.5LangmuirSecond86.50Electron density reduction in π-bond aromatic moieties due to the addition of -COOH[125]
Corncob-to-xylose residueNitrogen doped magnesium oxide400/20–50024FreundlichSecond1429.00Ion exchange, precipitation and complexation[42]
Ragweed---450/21000 ppm24LangmuirSecond358.70Precipitation, ion exchange, complexation[142]
Sewage sludge---500/0.50–3206LangmuirFirst41.20---[113]
Sludge---550/2200–10004FreundlichSecond30.88Co-precipitation, complexation[128]
Long-root Eichhornia crassipes---350/20 min1.0 mmol·L−150minLangmuirSecond118.10Complexation[126]
Eucalypts leaf residueZnCl2, FeCl3 and FeSO4700/20–10048------52.40---[127]
Microcrystalline celluloseIron nanoparticles10000–150024TempkinSecond17.30Reduction[129]
Banana peels---600/20–60048FreundlichSecond247.10Electrostatic attraction[50]
Rice husk---700/10–60048FreundlichSecond26.70Precipitation[130]
Hickory woodNaOH600/22–10024Langmuir---53.60---[74]
Orange peel---5006–2236LangmuirSecond86.96Ion exchange and surface precipitation[49]
Rice strawKMnO4420/41.0 mmol·L−18LangmuirSecond305.25Complexation[131]
Hickory woodKMnO4600/10–10024Redlich Peterson Richie n-th-order153.10Surface adsorption mechanisms[44]
Swine sludgeThiourea300/0.50–10048LangmuirSecond145.00Ion exchange[80]
Hickory---350/55–25024Dubinin–Radushkevich---16.30Cation exchange[99]
Maple woodH2O2550/5–55024Langmuir---43.30 ---[143]
Coconut fiber---500/450–50024LangmuirSecond175.40Cation exchange, complexation with functional group, precipitation[144]
British broadleaf hardwood---600---24LangmuirSecond47.66Cation exchange[76]
Raw bagasseKMnO4600/85–20048LangmuirSecond37.45Precipitation, ion exchange[145]
Ganoderma lucidum substrate---650/20–30024FreundlichSecond262.76Precipitation[146]
Aerobic granular sludgeFeCl3·6H2O, FeSO4·7H2O200/85–15012LangmuirSecond127.00Surface complexation, electrostatic attraction, and precipitation[147]
Camellia seed husk---700/10–30048LangmuirSecond109.67Ion exchange, complexation, Pb–π interaction, and precipitation[148]
Mulberry wood---650/4–––---Freundlich---250.00Ion exchange and chemical precipitation, Pb2+-p-electrons interaction[149]
Biogas residueFeCl3, FeSO4·7H2O700/225–3004LangmuirSecond181.82Surface complexation and precipitation, electrostatic attraction[150]
Printing leaflets---600/220–40024LangmuirSecond1555.00Electrostatic interactions, and pi-pi interactions[151]
Note: --- means no relevant data.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, C.; Wang, X.; Li, N.; Tao, J.; Yan, B.; Cui, X.; Chen, G. Adsorption of Lead from Aqueous Solution by Biochar: A Review. Clean Technol. 2022, 4, 629-652. https://doi.org/10.3390/cleantechnol4030039

AMA Style

Wang C, Wang X, Li N, Tao J, Yan B, Cui X, Chen G. Adsorption of Lead from Aqueous Solution by Biochar: A Review. Clean Technologies. 2022; 4(3):629-652. https://doi.org/10.3390/cleantechnol4030039

Chicago/Turabian Style

Wang, Chuanbin, Xutong Wang, Ning Li, Junyu Tao, Beibei Yan, Xiaoqiang Cui, and Guanyi Chen. 2022. "Adsorption of Lead from Aqueous Solution by Biochar: A Review" Clean Technologies 4, no. 3: 629-652. https://doi.org/10.3390/cleantechnol4030039

Article Metrics

Back to TopTop