Next Article in Journal
A Sustainable Approach for Enhancing Stability and Bioactivity of Allium ursinum Extract for Food Additive Applications
Previous Article in Journal
A Chemical Safety Assessment of Lyocell-Based Activated Carbon Fiber with a High Surface Area through the Evaluation of HCl Gas Adsorption and Electrochemical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exploring Humic Acid as an Efficient and Selective Adsorbent for Lead Removal in Multi-Metal Coexistence Systems: A Review

by
Shuwen Xue
1,
Yunhu Hu
1,
Keji Wan
2,* and
Zhenyong Miao
3,*
1
School of Chemistry and Materials Engineering, Huainan Normal University, Huainan 232001, China
2
National Engineering Research Center of Coal Preparation and Purification, China University of Mining and Technology, Xuzhou 221116, China
3
School of Chemical Engineering and Technology, China University of Mining and Technology, Xuzhou 221116, China
*
Authors to whom correspondence should be addressed.
Separations 2024, 11(3), 80; https://doi.org/10.3390/separations11030080
Submission received: 2 February 2024 / Revised: 22 February 2024 / Accepted: 28 February 2024 / Published: 4 March 2024
(This article belongs to the Special Issue Selective Separation of Heavy Metals from Water/Soil)

Abstract

:
Water pollution poses a global threat to human health, particularly in terms of ensuring a safe supply of drinking water. The accumulation of heavy metals from various water sources is increasing, driving the search for effective and environmentally friendly approaches and materials for metal removal. This review investigates the selective adsorption of Pb2+ by humic acid (HA) in a multi-metal coexistence solution. The focus is on discussing approaches to the structural identification of HA, highlighting that separation techniques are an effective method to reduce its heterogeneity. Starting from the key structural units of HA, the study reveals the interaction between HA and heavy metals. Approaches to enhance Pb2+ selective adsorption are explored, proposing that introducing activating groups, Ca ion exchange, and optimizing pore structures are effective approaches for improving lead ion selective adsorption. Ca2+ activation is suggested as a future research direction for lead-selective adsorption. Additionally, attaining lead selective adsorption through pH regulation’s significance is emphasized. This research contributes to a fundamental understanding of HA’s role as a selective lead adsorbent while offering practical implications for developing environmentally friendly adsorbent materials. The results aim to advance knowledge in environmental science and water treatment.

1. Introduction

Water, as the primary constituent of life on Earth, is confronted with persistent challenges to its quality and quantity due to urbanization, industrialization, and the growing global population. The emission of toxic heavy metals, such as lead, has reached substantial levels, with approximately 350,000 tons of Pb released per million tons of emissions [1]. Various industrial processes contribute to the release of Pb, leading to severe water pollution. Wastewater containing Pb2+ poses a significant threat to aquatic ecosystems and human health. Upon entering the human body, lead can adversely affect various organs, resulting in pathological changes. Elevated lead levels in the body can lead to detrimental effects on the brain, kidneys, and reproductive system. The impact on nerve cell myelin can impair muscular functions. Environmental protection standards, such as those set by the Environmental Protection Agency (EPA) and the World Health Organization (WHO), dictate that the Pb2+ content in drinking water should not exceed 0.015 mg L−1 and 0.01 mg L−1, respectively [2,3]. The current challenges in treating lead-contaminated wastewater mainly include high toxicity and low permissible limits, low pH, the complexity of wastewater, cost, as well as sustainability and waste management.
In response to these challenges, ongoing research and development efforts are focused on identifying more efficient, cost-effective, and sustainable methods for the removal of lead from wastewater. These include precipitation [4], electrolysis [5], adsorption [6], photocatalysis [7], membrane separation [8], and fenton-like oxidation [9], which are employed in water treatment. Precipitation methods currently stand as the primary treatment approach for acidic lead-containing wastewater. However, conventional chemical precipitation methods exhibit concentration limitations. Post-precipitation effluent fails to meet national discharge standards, and challenges persist in solid–liquid separation, posing obstacles to achieving comprehensive purification of lead-containing wastewater. Electrochemical methods can efficiently remove lead at low concentrations, but they have high energy consumption. Membrane filtration methods can remove a wide range of pollutants, including Pb2+, to below detectable limits, but membrane fouling reduces efficiency and increases costs. Fenton-like oxidation technology can rapidly and multi functionally treat lead-contaminated wastewater, but it generates sludge, increasing the cost and complexity of the wastewater treatment process. Compared to other heavy metal treatment methods, adsorption offers greater diversity and applicability, reversibility [10], low cost, ease of operation, and efficiency [11]. Therefore, it is widely applied and highly regarded in the field of water treatment. However, the efficiency and cost-effectiveness of adsorption depend on the source of the raw materials, making the selection of good materials for lead-contaminated wastewater crucial [12].
Coal, a traditional high-carbon energy source in China, remains a cornerstone of the current energy supply. With the goals of carbon peak and carbon neutrality, the nation, grounded in its reliance on coal, emphasizes the effective and clean utilization of coal, enhances the capacity for integrating new energy sources, and drives the optimization of coal and new energy synergies [13]. With the reduction in high-quality coal resources and increased difficulty of extraction in recent years, brown coal, characterized by large geological reserves and lower extraction complexity, has gradually become an important component of China’s energy production and supply. The colossal reserves of brown coal in China not only ensure raw material security, but also provide avenues for the materialization and purification of low-quality coal resources. Brown coal, serving as a source of humic acid (HA), is rich in oxygen-containing functional groups, capable of forming coordination bonds with heavy metals. It serves as a subsequent strengthening step in the chemical precipitation treatment of heavy metal wastewater. Constituting 25% to 45% of the elemental composition in coal-derived HAs, the diversified oxygen-containing functional groups exhibit robust electron-donating capabilities and affinity toward metal ions, and are extensively employed in the adsorption and removal of heavy metals [14].
Coal-derived HA, as a potential adsorbent for lead-containing wastewater, and its complexation with heavy metals are typically influenced by the wastewater’s pH. Existing wastewater treatments show enhanced separation efficiency in weakly acidic solutions, while strongly acidic solutions may impact the separation of lead. Lead-containing wastewater is characterized by a complex multi-metal system, with Pb2+/Cu2+/Cd2+ commonly coexisting, especially in wastewater, from lead smelting and battery discharge [2], resulting in reduced lead adsorption. Thus, comprehending the physicochemical properties of HA and heavy metals, along with summarizing effective methods to enhance the selective adsorption of Pb2+, provides a theoretical basis for the preparation of adsorbents, effectively shortening the process flow. Addressing these challenges requires comprehensive reviews of HA structural identification and literature on HA–heavy metal interactions.
This review aims to address the research status of HA, providing an overview based on the molecular structure of HA from two aspects: characterization methods and separation methods, proposing the interaction mechanisms between typical structures of HA and heavy metals. Additionally, this paper explores the selective adsorption mechanism of Pb2+ in a multi-metal coexistence system, considering the types and physical characteristics of heavy metals in aqueous solutions, as well as the application of selective adsorbents in lead-containing wastewater. This contributes to theoretical advancements and methodological innovations for the efficient purification of Pb2+-containing wastewater.

2. Research Status of Molecular Structure of HA

Given the fundamental national conditions in China, where fossil energy dominates, accelerating the research, development, and widespread application of clean and efficient utilization technologies for fossil energy has become a new trend in energy transformation. Coal-based material utilization has emerged as a key aspect of this transition. Lignite, a type of coal rich in organic matter, has historically been primarily used as an energy resource. However, in recent years, lignite has gained extensive attention as a significant chemical raw material. As illustrated in Figure 1, lignite material is commonly employed in the production of carbon nanotubes, carbon dots, porous carbon materials, graphene, fullerenes, carbon fibers, humic acids (HAs) and other applications. HAs exhibit diverse structural features, including functional groups such as carboxyl, phenolic hydroxyl, aldehyde, and ketone [15]. These functional groups confer unique chemical properties to lignite-HA, forming the foundation for its extensive applications. Lignite-HA demonstrates potential value across various fields. Notably, its use as an adsorbent in water treatment is a significant application. The multitude of functional groups in lignite-HA contributes to its outstanding adsorption performance for heavy metal ions, making it a promising green material for addressing water pollution issues. Different humic substances, including humic acid, exhibit significant structural variations due to differences in their formation processes [16]. Humic substances can be categorized into HA, fulvic acid, and hymatomelanic acid based on their solubility in acidic and alkaline conditions [17]. The structural analysis of HAs, as illustrated in Figure 2, faces challenges due to their inherent heterogeneity, resulting in significant variations and incomplete structural identification.
Advancements in understanding the molecular structure of HAs have been pivotal in driving the development and refinement of characterization methods. As researchers have unraveled more complex aspects of HAs, such as their diverse functional groups, molecular size distribution, and heterogeneous composition, the need for more sophisticated and sensitive analytical techniques has become evident. This progression in molecular understanding has directly impacted the development and application of characterization methods in several ways. The first involves characterizing elemental analysis of HAs. This provides information on the carbon, hydrogen, nitrogen, sulfur, and oxygen content of HAs, offering insights into their overall composition and potential reactivity. The second involves characterizing the complex functional groups within HAs using spectroscopic techniques, such as UV-Vis spectroscopy, (Fourier-transform infrared) FIIR spectroscopy, and nuclear magnetic resonance (NMR). These techniques analyze the overall chemical structural features of HAs. These methods have been refined to provide more detailed insights into the functional groups present in HAs, allowing for a deeper understanding of their chemical behavior and interactions in the environment. The third approach utilizes physical separation or chemical degradation methods [18], where samples are fractionated into components of varying molecular sizes or large molecules are fragmented for subsequent identification and structural analysis using spectroscopic techniques. In addition, the recognition of the wide molecular size range and structural diversity within HA populations has led to the development of more advanced chromatographic techniques, such as size-exclusion chromatography (SEC) coupled with multi-angle laser light scattering (MALLS), to accurately determine their molecular size distributions and conformational characteristics. The interaction of HAs with mineral surfaces and their role in soil and water systems have highlighted the need for advanced microscopic and surface analysis techniques, such as scanning electron microscopy (SEM) and X-ray diffraction (XRD). These techniques have been adapted to study the morphology and surface interactions of HAs, further elucidating their environmental roles.
Understanding the chemical composition and structure of HAs is not only crucial for elucidating their origin and formation mechanisms, but also for making inferences about biomass cycling on earth and comprehending the entire geological–chemical evolutionary processes. Moreover, this knowledge holds vital significance in guiding practical applications of HAs.

2.1. Characterization Methods of HA Structure

2.1.1. Characterization of HA Elemental Composition

Analyzing and characterizing the structure and composition of HAs are fundamental to their resource utilization. The characterization of HAs typically involves elemental composition and functional group characterization [19]. Elemental composition, representing the most basic characteristics of HAs, includes common organic elements like carbon, hydrogen, oxygen, nitrogen, and sulfur. The proportions of carbon and oxygen vary with different sources, with carbon ranging from 45% to 66% and oxygen from 26% to 47%. Ratios like H/C, O/C, and N/C serve as indicators of HA’s structure and types, providing insights into structural differences among various HA samples [20]. H/C represents the degree of aromatic condensation, the O/C ratio indicates the quantity of oxygen functional groups, and N/C reflects the number of nitrogen-containing groups, usually within the range of 0.02 to 0.06 [21].

2.1.2. Characterization of HA Functional Groups

In addition to elemental analysis, functional groups of HAs can be obtained through potentiometric titration and spectroscopic analysis. Between 68% and 91% of oxygen in HAs exists in an organic state. The sum of oxygen functional groups constitutes the total acidity of HAs (carboxylic acids, phenolic hydroxyls, quinones, methoxy groups, etc.), determining the primary active sites of HA chemical properties and biological effects [22]. Common spectroscopic analysis methods include UV-Vis spectroscopy, infrared spectroscopy (FTIR), and nuclear magnetic resonance (NMR). UV-Vis spectroscopy, with a significant structural parameter (E4/E6), serves as an indicator of the degree of aromatic condensation and the conjugation degree of other unsaturated bonds within HAs [23]. Datta et al. [24] found that soil HAs exhibit relatively higher E4/E6 values, suggesting a lower degree of aromatic compound condensation. Fatima et al. [25], through UV-Vis spectroscopy, observed changes in the functional groups of HAs in lignite before and after nitric acid treatment, discovering that the E4/E6 ratio of oxidized HAs increased, indicating higher biological activity post-oxidation.
FTIR identifies absorption peaks and quantitatively analyzes the degree of oxidation, aromaticity, and abundance of hydroxyl functional groups. Nuclear magnetic resonance technology provides an effective method for characterizing the structural composition of HA functional groups.
Solid-state NMR, based on the attribution of functional groups containing carbon, allows for the calculation of the esterification degree, aromaticity, and hydrophobicity, providing a comprehensive analysis of HA structure [26]. Verheyen et al. [27] utilized carbon spectroscopy and elemental analysis to identify HA structures, detecting an increased presence of fatty chains and aromatic rings within HAs. Ricca et al. [28] employed a range of spectroscopic methods, including UV-Vis absorption spectroscopy, FIIR, NMR, and thermal analysis, to analyze HA structures, determining an aromaticity of approximately 76% based on NMR data. Balkaş et al. [29] used similar analytical methods for characterization analysis, confirming the presence of numerous fatty chains within HAs. Novák et al. [30], through chemical and spectroscopic methods (UV-Vis absorption spectroscopy, infrared spectroscopy, and NMR), identified abundant sulfonic acid groups in HAs, detected a distinctive methoxy carbon at 55.9 ppm, and found aliphatic carbons, albeit in relatively low quantities, resulting in a calculated high aromaticity of approximately 60.6%, with a fundamental lignin skeleton. Additionally, Peuravuori et al. [31], employing similar spectral analysis, investigated the structural characteristics of South Moravian lignite and its HA component. They discovered that aliphatic compounds are tightly bound within the macromolecular network of lignite HA materials, forming complex bridging chains or complexes with hydroaromatic or polymethylene structural units.
Studies on HA structures can be categorized into destructive and non-destructive methods. The aforementioned methods for studying HA structures generally employ non-destructive techniques. Destructive pyrolysis techniques have proven valuable for obtaining detailed molecular information. Initially, Clemow et al. [32] identified HA structures at different temperatures using NMR spectroscopy and pyrolysis–gas chromatography–mass spectrometry (Py-GC/MS), observing numerous side chains, aromatic hydrocarbons, and phenolic compounds. Building upon this, Qian et al. [33] studied the structure of HAs in eastern Spanish lignite using methylated pyrolysis–mass spectrometry, identifying products of thiophene, methylthiophene, and benzothiophene, as well as benzocarboxylic acid and phenol derivatives derived from lignin. Simultaneously, aliphatic chains on the surface of HAs are readily decomposed at a low-temperature stage, with active destruction occurring in lignin and polyester-like structural domains at an intermediate stage (selective degradation process). In the advanced stage, structural rearrangement occurs in the advanced stage (simultaneous chemical decomposition and endothermic condensation). From the above references, it can be concluded that coal-derived HA structures have lignin-derived compounds as their skeletons, featuring highly condensed aromatic rings substituted with carboxyl, phenolic, and methoxy groups. HAs, being complex, natural, organic compounds, face limitations in physical, chemical, and spectral analysis methods.
Each of these techniques (UV-Vis, FTIR, and NMR) has its advantages, but also comes with limitations and sources of error, particularly from perspectives of sample preparation and handling and sensitivity. For sample preparation and handling, in UV-Vis analysis, the solubility of the sample and the choice of solvent can affect the absorption spectrum. If the sample contains undissolved particles or is not fully homogenized, it may lead to inaccurate measurements. Sample preparation significantly impacts FTIR analysis, as the homogeneity of the sample and the thickness of the film can influence the intensity and shape of absorption peaks. Moisture in the sample can also interfere with measurements due to water’s strong absorption of infrared light. NMR analysis requires the sample to have high purity. Impurities or impurity peaks from solvents may overlap with HA signals, making interpretation difficult. The concentration of the sample also needs to be precisely controlled for reproducible results. From a sensitivity perspective, UV-Vis has high sensitivity to certain functional groups, but may not provide comprehensive information about the complex structure of HAs. While FTIR can provide information about functional groups in HAs, its ability to distinguish between low concentrations and similar functional groups is limited. NMR provides the most detailed information about HA structures, but its sensitivity is limited by the sample concentration and the strength of the instrument’s magnetic field. Obtaining high-quality NMR spectra for lower concentration samples can be challenging. While these characterization techniques provide valuable information regarding the structural analysis of HAs, their limitations and sources of error must be considered when using them. Understanding these limitations helps in selecting the most appropriate analytical method and in the correct interpretation of the results of the analysis. By using a combination of techniques, the limitations of a single method can be overcome, yielding a more comprehensive and accurate understanding of HA structures.

2.2. HA Fractionation

2.2.1. Methods of HA Fractionation

The significant heterogeneity and structural complexity of HAs, attributed to diverse sources and ecological systems, pose challenges to research on HAs [34]. Studies have demonstrated the fundamental importance of isolating and purifying HAs for their characterization and utilization. Current analytical methods are subsequently employed to characterize these purified HAs, enhancing our understanding of the fundamental structure and providing theoretical support for their applications.
HAs, forming macromolecular polymers through hydrophobic or hydrogen bonding interactions, present challenges in terms of heterogeneity [35]. To effectively study HAs, obtaining fractions with reduced heterogeneity through fraction becomes necessary.
Due to the presence of reactive functional groups and highly uneven distribution within HA molecules, significant differences arise in their electron-donating and electron-withdrawing tendencies. These differences result in varied solubilities in solvents with different polarities and pH values [36]. Exploiting these acid–base properties, researchers employ solubility and precipitation grading [37]. Based on the differences in solubility in hydrochloric acid, HA-like substances are divided into HA and fulvic acid fractions [38]. Wu et al. [39] proposed that pH values affect the molecular size, molecular weight, and aggregation of HAs. Fujitake et al. [40], through sequential extraction and isolation of soil HAs using phosphate solutions at different pH values, observed that increasing the pH led to a gradual increase in the H/C ratio, while the O/C and O/H ratios decreased. Additionally, the absorption intensity of aliphatic and amide groups increased, while the carboxyl group intensity decreased. Jovanovic et al. [41] explored the pH (2–11) dependency of HA particle size, finding that HAs exhibit large molecular aggregates or supramolecular structures. Klučáková et al. [42] investigated the linear relationship between functional group intensity and the dissociation ability of HAs at different pH levels. Zhang et al. [43], through continuous alkali extraction procedures, obtained ten fractions of HA, revealing that increased extraction led to higher molecular weight, decreased carboxyl and phenolic group content, and weakened aromaticity and polarity.
Charge-based fractionation primarily involves electrophoresis, segregating solute molecules with different charges based on their differential movement rates in an electric field. HAs are negatively charged polymeric anions, and this process is typically conducted in alkaline buffer solutions. Their migration distance and speed are negatively correlated with the molecular size and positively correlated with charge density [44]. Hermosin et al. [45], utilizing thermal hydrolysis and methylation along with polyacrylamide gel electrophoresis for HAs, identified primary structures that included fatty acids and lignin. Saiz-Jimenez et al. [46], through thermal degradation of HAs, similarly obtained lignin products. Tang et al. [47], employing polyacrylamide gel electrophoresis for the fractionation of HAs and fulvic acids in water, and subsequently analyzing them through Py-GC/MS, discovered that the richest components in the largest fraction were mainly fatty acids (C16 and C18), concurrently detecting lignin-derived compounds. Despite obtaining relatively uniform HA and fulvic acid fractions through electrophoresis, this method cannot entirely prevent urea from being absorbed into the separation solution, limiting its applicability, especially in the presence of nitrogenous components.
Based on adsorption-desorption characteristics, several methods involve the use of porous silica gel, alumina, activated carbon, wood charcoal, etc., for adsorption–desorption grading. XRD-8 resin, due to its relatively weak adsorption and high desorption rate, has been proven to be an effective adsorbent for separating water-soluble organic compounds of the fulvic acid fractionation. Drosos et al. [48], employing XAD-8 resin for the fractionation of lignite HAs, observed that each fraction contained more oxygen functional groups than the original HA, which is consistent with the results of Jones [49].
Fractionation based on molecular size is one of the common methods, primarily relying on different molecular weights as grading criteria. It divides organic polydisperse systems ranging from thousands to hundreds of thousands of molecules into several molecular weight ranges [50], among which ultrafiltration is the most common. Ultrafiltration separates membranes based on pore size and is related to molecular charge and configuration. Li et al. [51], utilizing ultrafiltration technology for HA fractionation and employing elemental analysis and pyrolysis–mass spectrometry, characterized the chemical compositions of HA samples. The results showed that, with increasing molecular weight (Mw), the polarity, aromaticity, and oxygen content decreased. HAs with smaller Mw possibly originate from lignin, while those with larger Mw may originate from lipid-rich biopolymers. Tanaka et al. [52], also employing ultrafiltration technology for HA fractionation, found significant structural and compositional differences in each fraction. As the molecular weight retention decreased, the aromatic carbon content and aromaticity of each fraction gradually decreased, while the total acidity, carboxylic acid content, and fatty carbon content increased.

2.2.2. The Potential Impact of HA Fractionation on the Environment

HA fractionation methods play a crucial role in environmental science and soil chemistry research, yet these methods may also have potential environmental impacts. Analyzing from the perspectives of waste production, energy consumption, and chemical usage, here is a summary of the potential environmental impacts of HA fractionation: (1) Waste production: the fractionation process of HAs can generate solid waste and chemical waste liquids. The use of solvents for the extraction and separation of HAs results in waste liquids containing organic solvents, which, if not properly treated before discharge, can pollute water bodies and soil. Moreover, the unusable fractions produced during the fractionation process may need to be disposed of as solid waste, increasing the cost and complexity of waste management. (2) Energy consumption: the extraction and fractionation processes of HAs often require high-energy-consuming equipment, such as centrifuges, freeze dryers, and high-temperature ovens. While energy consumption may not be a major concern for laboratory-scale studies, it can lead to a significant environmental footprint at an industrial scale, especially when the energy comes from non-renewable sources. (3) Chemical usage: the wide use of chemicals in the HA fractionation process, such as strong acids, strong bases, and organic solvents, has potential negative impacts on the environment. The production, transportation, and use of these chemicals can generate greenhouse gas emissions, and additional environmental protection measures are needed to prevent pollution during the handling and disposal of these chemicals. Selecting more environmentally friendly chemicals and optimizing the amount of chemicals used are key to reducing environmental impacts.
To mitigate the potential environmental impacts of HA fractionation, the following strategies can be adopted: (1) Optimized waste management: Reduce waste generation through waste recycling and reuse strategies, and employ environmentally friendly waste treatment and disposal methods. (2) Increased energy efficiency: Use more energy-efficient equipment and technologies to reduce energy consumption and utilize renewable energy sources as much as possible. (3) Green chemical usage: Choose more environmentally friendly solvents and chemicals; reduce the use of harmful chemicals; and employ safer, more sustainable chemical processing methods. By implementing these strategies, the environmental impact of HA fractionation can be effectively managed and mitigated, contributing to the advancement of the field towards more sustainable research and application directions.

2.3. The Interaction between Typical Structures of HA and Heavy Metal Ions

Understanding the adsorption mechanisms of HA on heavy metals is crucial for its practical applications. The interaction between HA and heavy metal ions is complex, owing to their heterogeneous, polyelectrolytic, and polydisperse properties. Various interactions occur, ranging from non-specific electrostatic interactions with the humic particle surface to specific interactions forming complexes and chelates with functional groups. Heavy metal ions can interact with humic substances in several ways [53], ranging from non-specific electrostatic interactions with the surfaces of humic particles to specific interactions forming complexes and chelates with functional groups [54]. The complexity arising from their structural intricacies and heterogeneity restricts the exploration of the interaction between HA and heavy metals. Based on the earlier comprehension of the molecular structure of HA, this section explores the interaction between the characteristic functional groups and the structure of HA concerning heavy metals.
HA exhibits various typical structural features, primarily characterized by functional groups like carboxyl, phenolic hydroxyl, aldehyde, and ketone. These functional groups play crucial roles in the interaction with heavy metal ions. The mechanism of the interaction between HA and heavy metals is summarized in Figure 3.
(1)
Electrostatic interaction: HA exhibits diverse pKa values, functioning as a weak acid and a strong acidic ion-exchange as a polyelectrolyte. This characteristic allows for the formation of diffuse double layers around charged particles, facilitating heavy metal adsorption [55,56]. HISS standards suggest that the content of –COOH in HA significantly exceeds that of hydroxyl groups, and their dissociation occurs at pH > 4.4, resulting in a predominantly negatively charged surface on HA [57]. This facilitates electrostatic reactions with heavy metals. At higher pH values (>9.5), –OH become more prone to metal complexation as they are easily protonated. The increase in pH effectively deprotonates carboxylic groups, enhancing their binding with positively charged metals [58]. Studies have shown that hydroxyl and carboxyl groups in HA chelate with heavy metals, enhancing the adsorption rate [59]. Zhao et al. [44] proposed that phenolic and carboxylic groups are the primary adsorption sites for Pb2+, while, according to Kinnibrugh’s research [60], different functional groups exhibit varying binding affinities with metal ions, with carboxyl groups showing significant binding capacities: Pb2+ > Cu2+ > Cd2+. Beyond active sites, structural features also influence the complexation of heavy metals by HA [61].
(2)
Carboxyl Coordination and Phenolic Hydroxyl Coordination: The carboxyl functional groups (–COOH) in HA are important structures involved in the complexation with heavy metal ions. Heavy metal ions typically form coordination bonds with hydroxy oxygen in carboxyl groups or oxygen in carboxyl groups, resulting in the formation of stable complexes. This coordination interaction forms the basis for the selective adsorption of heavy metal ions by HA. The phenolic hydroxyl functional groups (–OH) in HA are also key structures involved in the complexation with heavy metal ions. The oxygen atoms in these functional groups can form coordination bonds, facilitating the coordination interaction with heavy metal ions and promoting the adsorption and fixation of heavy metals.
(3)
Ion Exchange: Functional groups in HA possess certain ion exchange capabilities, allowing adsorption through ion exchange with heavy metal ions. This interaction is often influenced by factors such as solution pH and ion concentration.
(4)
π electrons: The π electrons of aromatic functional groups play a crucial role in interacting with cations, serving as π electron donors for the adsorption of heavy metals [62]. The cation–π electron interaction is dependent on the aromaticity degree of the HA surface. Higher C/H ratios indicate higher aromaticity, resulting in greater electron donation capacity, reduction capacity, and adsorption capability. Li et al. [63] observed heavy metal adsorption involving cation–π interactions, especially with corn stover biochar.
(5)
Precipitation: The interaction between carbonate (CO32−) and sulfate groups (SO42−) in HA and heavy metals is one of the significant behaviors of HA in the environment. These interactions generally occur through mechanisms such as coordination, ion exchange, and adsorption. (1) Coordination: The carboxyl and phenolic functional groups in HAs can form coordination bonds with heavy metals, resulting in the formation of complexes. The carboxyl groups on carbonate and sulfate can provide coordination sites, forming complexes with heavy metal ions, stabilizing their forms of presence, and reducing their toxicity. (2) Ion exchange: The carbonate and sulfate groups in Hs carry negative charges and can participate in ion exchange reactions. In the presence of heavy metal ions in the environment, they can undergo ion exchange with the carbonate and sulfate groups in HAs, being adsorbed onto the surface of HAs, thus reducing the activity and toxicity of heavy metal ions in the environment. (3) Precipitation: Under the influence of high concentrations of carbonate or sulfate, as well as factors such as the concentration of heavy metal ions and pH values, precipitation reactions of heavy metals with carbonate or sulfate may occur. Some heavy metal ions (such as Ca2+, Pb2+) can form carbonate precipitates with carbonate, especially under alkaline conditions. Similarly, some heavy metal ions (such as Cd2+, Pb2+, Hg2+) can also form sulfate precipitates with sulfate, particularly under acidic conditions.
The adsorption process of HA on heavy metals involves multiple simultaneous mechanisms, and it is intricately linked to both the chemical forms of heavy metals and the physicochemical properties of HA. Current research primarily focuses on the chelation of carboxyl and phenolic hydroxyl groups, with limited exploration of key factors influencing heavy metal adsorption. Additionally, the relationship between the physicochemical properties of HA and its adsorption of heavy metals has not been extensively studied. Therefore, elucidating this relationship holds significance for fundamental research and practical applications of HA. Moreover, research on the interaction between typical structures of HA and heavy metals has been limited to π electrons, with little exploration of other typical structures. Therefore, the identification of typical structures of HA holds significant importance for the study of the interaction between HA and heavy metals, as well as the design of adsorbent materials.
Overall, the interaction of carbonate and sulfate groups in HA with heavy metals through coordination, ion exchange, and adsorption regulates the migration, transformation, and bio-availability of heavy metals in the environment, playing a crucial role in reducing heavy metal pollution.

3. Selective Adsorption of Lead Ion by Coexisting System of Multiple Metals

3.1. Types and Physicochemical Characteristics of Heavy Metals in Aqueous Solution

In the context of the coexistence of multiple metal ions and the treatment of complex wastewater, selectivity becomes a crucial criterion for evaluating overall adsorption performance. Adsorbents demonstrating high selectivity can effectively separate metal ions through a simple adsorption–desorption cycle. Selectivity can be achieved by modifying the surface functional groups of the adsorbent or adjusting the solution pH, thereby altering the surface charge of both the adsorbent and the adsorbate. Selective adsorption of Pb2+ in a multi-element coexistence system has become a focal point in the field of environmental pollution control.
Significant progress has been made in the research on the selective adsorption of Pb2+ in multi-element coexistence systems. Adsorption, recognized as a simple, efficient, economical, safe, and non-secondary pollution method, stands out as one of the most attractive technologies for the separation and recovery of Pb2+. Various adsorption materials, including biomass materials, layered double hydroxides [64], metal–organic frameworks [65], nano-porous adsorbents [66], and carbon-based nanocomposites [67], have been synthesized and applied for the removal of Pb2+ from wastewater. Researchers have explored the impact of multi-element coexistence systems on the selective adsorption of Pb2+ by preparing different adsorption materials and employing diverse adsorption conditions. Additionally, they have investigated the influence of parameters such as solution pH and ionic strength on the adsorption of Pb2+ in the presence of multi-element coexistence [68,69].
Density functional theory (DFT) has been widely employed to elucidate potential adsorption mechanisms. Theoretical methods not only provide insights into the removal efficiency and binding preferences of metal ions, but also offer information on energy and geometric structure changes during the partial interactions between adsorbates and adsorbents. Theoretical studies of the adsorption process provide additional crucial advantages, such as interaction energy at different positions of the adsorbent, preferences for different metal ions, and possible interaction patterns [70]. Research on the structure and coordination characteristics of metal-adsorbent systems plays a vital role in designing and customizing new materials for the efficient removal of heavy metals from wastewater [71]. Combining theoretical research results with experimental findings has proven highly useful for understanding the mechanisms of interaction between adsorbents and metal ions during the adsorption process, as well as for synthesizing functionalized materials suitable for wastewater treatment processes [72]. Molecular dynamics simulation methods have been employed to study the competitive adsorption processes of metal ions such as Zn2+, Cd2+, and Cu2+ for Pb2+, revealing that the interactions between different ions exert a significant impact on selective adsorption.
In actual heavy metal wastewater, multiple metals coexist, each with different heavy metal characteristics. Predicting metal-adsorbent interactions is facilitated by the Hard and Soft Acid and Bases (HSAB) principle. Ionic bonds primarily form between hard metals and hard ligands, while quasi-covalent bonds exist in complexes of soft metal ligands [73]. The covalent index (X2mr), based on electronegativity (Xm) and ionic radius (r), can be used to estimate “hardness” [74,75]. As the covalent index value increases, metal species become “softer”, resulting in stronger covalent characteristics in their bonds with ligands. Neiber et al. [76] investigated the selective adsorption of amino-phosphonates for Pb2+ in a multi-metal system, comparing the characteristics of various metal ions. They proposed that the main reason for the adsorbent’s selective performance is the covalent index, validating the above conclusion. In addition to the mentioned factors, the difference in the electronegativity of metal ions is also a crucial factor affecting selective adsorption. Liu et al. [77] studied the selective removal of Pb2+ and Cu2+ from water using ethylenediamine-functionalized polymer adsorbents, attributing the selective adsorption of Pb2+ to the difference in electronegativity between the two metal ions.

3.2. Application of Selective Adsorbents in Lead-Containing Wastewater

Improving the separation efficiency of specific heavy metal ions during the adsorption process is an effective approach for achieving a balance between environmental safety and economic benefits. The following provides a brief overview of common methods to enhance the selectivity for Pb2+.

3.2.1. Electrostatic Interaction

Electrostatic interaction, common in the environment, plays a key role in attracting heavy metals to adsorbents. The initial step involves the attraction of metal cations to the functional groups on the adsorbent surface through electrostatic forces, followed by further binding through other driving forces [78]. The efficacy of electrostatic interactions is highly influenced by the acidity of the reaction medium, primarily because the charge properties of metal species and the adsorbent surface vary at different pH values [79]. Low pH values promote the adsorption of anions, whereas an increase in solution pH results in the protonation of functional groups on the adsorbent surface due to the dissociation of H+, creating a negatively charged adsorbent surface that is more conducive to binding metal cations. Additionally, the electrostatic adsorption process correlates with the electronegativity of heavy metal ions. Metal cations with higher electronegativity usually exhibit stronger electrostatic attraction to negatively charged sites on the adsorbent [80]. In previous study conducted by our research group [81], HA was modified using hydrogen peroxide as an oxidant to enhance the surface activity sites and charge density of HA, achieving selective adsorption of Pb2+. The experimental results indicated that oxidized HA demonstrated selectivity for heavy metals in the order of Pb2+ (111.10 mg/g) > Cu2+ (76.92 mg/g) > Cd2+ (71.43 mg/g) > Ni2+ (33.33 mg/g). While the strength of the charge interaction allows for the selective separation of some ions, the selectivity is relatively weak.

3.2.2. Specific Chelation Interaction

Chelation interaction is crucial for adsorbing heavy metals, with the adsorbent’s surface directly affecting its ability to capture specific metals [82]. The adsorbent surface typically contains abundant active site structures, and based on the “hard and soft acid-base theory”, oxygen-containing functional groups, amino groups, and thiol groups can effectively and selectively adsorb heavy metals. Zou et al. [83] synthesized a tannic acid-based adsorbent for heavy metal removal through self-assembly, exhibiting outstanding selectivity for Pb2+, with a distribution coefficient for Pb2+ 40–170 times higher than that of other metal ions. XPS and DFT calculations revealed that the phenolic hydroxyl groups on tannic acid served as the primary active sites, engaging in a double proton exchange mechanism after coordination with Pb2+. Apart from phenolic hydroxyl groups, amino groups are also recognized as active sites for selective adsorption. Liu et al. [84] introduced multiple ligands to develop a novel metal–organic framework adsorbent, demonstrating excellent selective adsorption of Pb2+. The maximum adsorption capacity of Pb2+ is 433.15 mg/g at pH 5. The distribution coefficient of lead ions (26.7 L/g) shows that UiO-FHD has excellent selective adsorption for Pb2+. It is revealed that the chelation of the sulfhydryl groups and the electrostatic interaction of the hydroxyl groups are the dominant factors to improve the removal rate of Pb2+ via density functional theory calculations. Tang et al. [85] successfully prepared a sulfur-containing amino urea metal–organic framework adsorbent for Pb2+ removal in wastewater. The absorption results showed that the maximum adsorption capacity of Pb(II) is 200.17 mg/g at 303 K and optimal pH 5. Calculating the separation coefficient of interfering ions from Pb2+ clarified the role of amino groups in selective adsorption. The adsorption mechanism of Pb(II) on the adsorbent occurs mainly through the chelation of Pb(II) with N and S atoms. These results indicate that UiO-66-PTC is an effective material for efficiently and selectively removing Pb(II) from solution, which is of practical significance. Nankawa et al. [86] synthesized an oxalate framework adsorbent for heavy metal adsorption, and adsorption tests revealed high selectivity for Pb2+ compared to other heavy metals. The attributed high selectivity was linked to oxalic acid having a greater affinity for Pb2+. Nuzahat et al. [87], through modeling and simulating the adsorption of Pb2+/Cu2+/Cd2+ on HA, determined that the binding strength of the three metals followed the sequence Pb2+ > Cu2+ > Cd2+. Among these metals, Cd2+ posed the most significant challenging in complex with HA, consistent with the efficiency sequence of heavy metal adsorption obtained by Kerndorff [88]: Pb2+ > Cu2+ > Cd2+ > Zn2+ > Ni2+ > Co2+ > Mn2+. The work explores the interaction process of Pb with HA through integration of synchronous fluorescence and log-transformed UV–vis absorption spectroscopy coupled with spectral slope calculation and two-dimensional correlation analyses. The spectral slope calculation results show that the carboxylic and phenolic groups in HA were the predominant binding sites to Pb, and the interaction process was highly dependent on the solution’s pH. Correlation analyses confirmed a superior binding affinity of carboxylic groups in HA over phenolic groups. This indicates that, as an active site in HA, it is more conducive to the high selective adsorption of Pb2+.

3.2.3. Ion Exchange

Ion exchange enhances metal selectivity by using the charged adsorbent surface to attract and swap metal ions [89]. Under the exchange process, ions undergo strong non-classical polarization near the solid–liquid interface under the influence of the surface electric field. Meanwhile, polarized counterions shield the surface electric field as they approach, and this interaction between ions and the surface is termed the “electric field-quantum fluctuation” coupling effect [90]. Owing to the different polarization effects of ions of the same valence but different types in an electric field, they exhibit selectivity in adsorption on charged particles [91]. The ion exchange capacity of heavy metals and adsorbents is contingent on the pH of the solution, ion composition, electronegativity of heavy metal ions, and surface chemistry of the adsorbent [92,93]. During the adsorption process, the release of other ions and changes in solution pH, usually confirmed by the occurrence of ion exchange mechanisms, are often considered.
Hofmeister effect significantly influences the adsorption behavior of ions, and this effect is commonly observed in ion exchange adsorption as well. In classical theories, the selectivity of ion exchange is correlated with the valence of ions, where higher valence states result in stronger adsorption capabilities. Existing theories attribute the Hofmeister effect to factors such as ion–ion interactions, ion volume, ion hydration effects, and dispersion forces. Table 1 presents the characteristics of common alkali (alkaline earth) metal ions. For ions of the same valence, the larger the hydrated radius, the farther the ion charge center is from the surface, leading to weaker interaction between the ion and the surface. Regarding alkali metals, the ion exchange capacity follows the order K+ > Na+; for alkaline earth metals, the ion exchange capacity is ranked as Ca2+ > Mg2+. The expression for ion exchange reactions can be represented as:
a A + b B c C + d D
In Equation (1), A and B represent the reactants, corresponding to the solid and liquid phase components before the exchange reaction, respectively; C and D represent the products, corresponding to the chemical stoichiometry of solid and liquid components after the exchange reaction.
Ion exchange has been proven to be an effective method for enhancing adsorption performance. The current literature emphasizes ion exchange adsorption of heavy metals, particularly on biochar. Alkali (alkaline earth) metal ions (K+, Ca2+, Na+, and Mg2+) in biochar can exchange heavy metal ions in water [96]. Lu et al. [97] investigated the ion exchange of Pb2+ in sludge-based biochar, where ion exchange accounted for 45.6–60.2% of total adsorption, with a complexation ratio of 38.2–42.4%. Zhang et al. [98] explored the adsorption mechanism of Pb2+ by celery-derived biochar, indicating that ion exchange accounted for 62.8%, precipitation accounted for 34.9%, and only 2.3% was attributed to functional group complexation, achieving a high Pb2+ adsorption capacity of up to 300 mg g−1. Li et al. [99] used Mg-loaded biochar for heavy metal adsorption, indicating that mineral precipitation and cation exchange played a major role in the adsorption process, with theoretical maximum adsorption values reaching 333.33 mg g−1 for Cd2+, 370.37 mg g−1 for Cu2+, and 302.58 mg g−1 for Pb2+. Deng et al. [100], through MgCl2 modification of biochar, found that the modified biochar had an adsorption capacity 2.36–9.34 times higher than that before modification, with ion exchange being the predominant adsorption mechanism. The aforementioned ion exchange applied to biomass modification has also been successfully employed for the modification and preparation of HA materials using alkaline earth metals. Sabaté et al. [94] investigated the ion exchange capacity of peat-extracted HA for 17 heavy metals, and the study found that Pb2+ had the highest affinity for the ion exchange centers of HA, with the adsorption capacity for heavy metals satisfying Pb2+ > Cu2+ > Cd2+. Wang et al. [95] studied the use of Ca-loaded brown coal for the ion exchange removal of heavy metals from wastewater, and the adsorbent exhibited selectivity for Pb2+ > Cu2+ > Cd2+. This suggests that Ca2+ has a stronger displacement ability for Pb2+ and is more suitable for the removal of wastewater containing Pb2+. The regulation of adsorption selectivity can be achieved by changing the chemical properties of the substrate surface and controlling the nature and distribution of adsorption sites (Table 2).
In addition to its ion exchange function, studies have demonstrated its role in activating organic functional groups. The main functional groups of natural organic matter are shown in Table 2. The cation preferences for –COOH are as follows: Ca2+ ≥ Mg2+ ≥ Al3+ ≥ Na+. This indicates that Ca2+ exhibits a higher affinity for functional groups. The bridging role of Ca2+ in the organic network structure has been validated, forming a large molecular network structure of organic matter. This indicates that Ca2+ activates the organic functional groups [106].

3.2.4. Pore Structure and Matching with Heavy Metal Size

The process of heavy metal entry into pore structures through diffusion represents a physical adsorption process, mainly influenced by pore distribution, surface electronegativity, and the characteristics of heavy metals. The capacity for heavy metals in physical adsorption and chemical adsorption increases with the adsorbent’s specific surface area. The size match between adsorbent pores and metal ions is crucial for selective heavy metal adsorption. Alka et al. [107] employed amino and carboxyl-modified cuparenes and cyclodextrin decorated on nanosponges for the removal of lead-containing wastewater. The adsorption experimental results indicated that one of the reasons for the selective adsorption of Pb2+ is the role of the pore structure. Qiao et al. [108], by modifying the pore channels in conjugated microporous polymers with pyridine groups, efficiently removed Pb2+ while exhibiting high selectivity. The pore structure is one of the essential mechanisms for its selective adsorption of Pb2+. Currently, there is less research on the matching effect between pore structure and heavy metal size in the study of the adsorption mechanism of heavy metals, and the interaction between them is relatively weak.

3.2.5. The pH Values

pH levels directly influence how HA interacts with metals by changing the ionization of HA’s functional groups, affecting metal binding and complex formation. The following are the effects of pH on the interaction between HA and metals:
Variation in Ionization State: Functional groups in HA, such as carboxyl and phenolic hydroxyl groups, exhibit acidity. Under different pH conditions, these functional groups may undergo ionization. At lower pH values, carboxyl and phenolic hydroxyl groups may exist in non-ionized forms, while at higher pH values, they may ionize into negatively charged ions. This change in ionization state directly affects the interaction between HA and metals.
Stability of Complexes: The stability of complexes formed between HA and metals is typically influenced by pH. In certain pH ranges, HA may be more conducive to forming stable complexes, while at other pH conditions, dissociation may occur. This change in stability directly influences the degree of binding between metals and HA.
Solubility: The solubility of metal ions under different pH conditions also changes. Under some pH conditions, metal ions may be more prone to forming precipitates or complexes with HA, whereas under other pH conditions, metal ions may be more likely to exist in a dissolved form.
Therefore, understanding the influence of pH on the interaction between HA and metals is crucial for comprehending the behavior of heavy metals in the environment and their applications in environmental remediation and wastewater treatment. pH regulation can be an effective means to control the interaction between HA and metals, enabling selective adsorption or removal under specific environmental conditions.

3.2.6. Chemical Precipitation

The relationship between chemical precipitation and selective adsorption of Pb2+ is closely intertwined, particularly in the fields of environmental engineering and pollution control for treating lead-contaminated wastewater. Chemical precipitation removes metals by adjusting the pH and metal concentration to form insoluble complexes with the target metal [109]. Chen et al. [110] observed that phosphate precipitation, to some extent, elucidates the mechanism of selective adsorption of Pb2+.
In the treatment of complex wastewater, chemical precipitation and selective adsorption can be part of a multi-stage treatment process, where chemical precipitation removes the bulk of Pb2+, followed by selective adsorption for further refinement to meet lower lead ion concentration standards. Research and development of composite materials capable of both chemical precipitation and selective adsorption can achieve effective removal of lead ions in one step, improving treatment efficiency and reducing costs.

3.3. The Limitations, Challenges, and Future Prospects of Selective Adsorption Mechanisms for Pb2+

Discussing the limitations, challenges, and future prospects of selective adsorption mechanisms for Pb2+ involves analyzing aspects such as electrostatic interactions, chelation, ion exchange, pore structure and its match with heavy metal sizes, pH values, and chemical precipitation.

3.3.1. Limitations and Challenges of Electrostatic Interactions

Limitations: Electrostatic attraction highly depends on the pH and ionic strength of the solution, which may limit its application across diverse water qualities. Challenges: In high-salinity wastewater, background ions may shield the electrostatic attraction between the adsorbent and Pb2+, reducing adsorption efficiency. Prospects: Developing new adsorbent materials to enhance the specificity of electrostatic attraction for Pb2+ and to minimize interference from background ions.

3.3.2. Limitations and Challenges of Chelation

Limitations: Specific chelators may be effective only for certain metal ions, limiting their widespread application in complex wastewaters. Challenges: The stability and regenerability of chelating agents need further investigation for sustainable use. Prospects: Creating multifunctional chelators that work across a broader pH range and improve regeneration efficiency.

3.3.3. Limitations and Challenges of Ion Exchange

Limitations: Ion exchange capacity is limited by the adsorbent’s saturation, requiring replacement or regeneration after long-term use. Challenges: Competitive adsorption in multi-metal wastewaters may reduce selectivity for Pb2+. Prospects: Synthesizing new ion exchange materials with higher selectivity and larger capacity to address these issues.

3.3.4. Limitations and Challenges of Pore Structure and Heavy Metal Size Matching

Limitations: Designing and synthesizing adsorbents with precise pore structures is complex and challenging. Challenges: Ensuring adsorbent pore structures effectively differentiate between metal ions of similar sizes. Prospects: Using nanotechnology and molecular simulations to design adsorbents with highly selective pore structures.

3.3.5. Limitations and Challenges of pH Values

Limitations: Fluctuations in pH affect the adsorbent surface charge and the form of Pb2+, potentially limiting adsorption effectiveness. Challenges: Developing adsorbents that maintain high adsorption performance across a wide pH range. Prospects: Exploring automatic pH adjustment systems to optimize adsorption process pH conditions.

3.3.6. Limitations and Challenges of Chemical Precipitation

Limitations: Chemical precipitation may lead to the production of secondary pollutants, necessitating additional treatment steps. Challenges: Balancing the efficiency and cost of precipitation reactions, especially for large-scale applications. Prospects: Developing more environmentally friendly and cost-effective precipitants, as well as improving precipitation processes to reduce secondary pollution.
While there are several limitations and challenges associated with selective adsorption mechanisms for Pb2+, ongoing technological innovation and material development hold the promise of overcoming these hurdles in the future, achieving more efficient and environmentally friendly Pb2+ removal methods.
Table 3 above synthesizes findings from various studies on the use of HA for the removal of Pb2+ from different media. HA has been shown to be an effective agent for Pb2+ removal, leveraging mechanisms like ion exchange, complexation, and adsorption. This is a testament to the versatility of HA in treating Pb2+ contamination. The modification method of combining HA with nanometal materials enhances its adsorption capacity for heavy metal ions, and the modification with Ca2+ has verified ion exchange as an effective way to improve selectivity. The doping of Ca2+ has activated the adsorption active sites of HA. Concurrently, mechanistic studies have elucidated the fundamental interactions, especially the role of HA’s functional groups in binding with Pb2+. This is crucial for optimizing HA-based material in wastewater remediation strategies. The comparative study highlights HA’s superior performance relative to other adsorbents, suggesting that HA could replace or complement existing materials in water treatment.
In conclusion, while existing selective adsorption mechanisms for Pb2+ face several limitations and challenges, the unique characteristics of HA offer a promising outlook for its application as an adsorption material. Its potential for enhanced specificity, environmental sustainability, and economic viability in Pb2+ removal positions HA as a valuable resource for addressing current and future water treatment challenges.

4. Conclusions

The comprehensive review and analysis of studies on HA’s role in lead removal has illuminated the physical and chemical properties of HA, structural identification methods, and microscopic interactions with heavy metals. While many forms of HA demonstrate significant efficacy in binding and removing Pb from aqueous solutions, the chemical structure of HA, environmental conditions, and the physical properties of heavy metal ions affect the selective adsorption of Pb2+.
Enhancements of the selective adsorption capacity for Pb2+ can be achieved through modifications, such as increasing activating groups, employing ion exchange, and optimizing pore structures to match the size of heavy metals. Notably, the use of Ca2+ has been proven to be an effective method for activating HA groups. The modifications not only increase the number of active sites available for complexation with Pb2+, but also improve the stability of the resulting complexes, leading to higher removal efficiencies. Additionally, the modified HAs exhibit better performance under a wider range of environmental conditions, such as varying pH levels and the presence of competing ions, making them more versatile for real-world applications.
Future research should focus on further optimizing these HA modifications, exploring cost-effective and scalable modification processes, and evaluating the long-term sustainability and reusability of these materials in environmental remediation projects. By advancing our understanding and application of these optimized HAs, we can improve our strategies for mitigating lead pollution and protecting environmental and human health.

Author Contributions

S.X.: data curation, writing—original draft. Y.H.: writing—review and editing. K.W.: conceptualization, funding acquisition, writing—review and editing, supervision. Z.M.: funding acquisition, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The current study was funded by the National Natural Science Foundation of China (Grant No. 52004280), the Natural Science Foundation of Jiangsu Province of China (Grant No. BK20190629), and the National Key Research and Development Program of China (No. 2021YFC2902602).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chooto, P.; Wararatananurak, P.; Innuphat, C. Determination of trace levels of Pb(II) in tap water by anodic stripping voltammetry with boron-doped diamond electrode. Sci. Asia 2010, 36, 150–156. [Google Scholar] [CrossRef]
  2. Mansoor, L.; Erkan, Y.; Basak, S. Preparation and characterization of magnetic allylamine modified graphene oxide-poly(vinyl acetate-co-divinylbenzene) nanocomposite for vortex assisted magnetic solid phase extraction of some metal ions. Talanta Int. J. Pure Appl. Anal. Chem. 2016, 146, 130–137. [Google Scholar]
  3. Özdemir, C.; Saçmacı, Ş.; Kartal, Ş. A coprecipitation procedure for the determination of some metals in food and environmental samples by flame atomic absorption spectroscopy. Anal. Methods 2013, 5, 3977–3983. [Google Scholar] [CrossRef]
  4. Igawa, M.; Kanamori, H.; Nanzai, B. Gel-phase Extraction for the Removal of Heavy-metal Ions. Chem. Lett. 2010, 39, 996–997. [Google Scholar] [CrossRef]
  5. Zhang, W.; Liang, Z.; Feng, Q.; Wei, N.; Zhang, Y. Reed hemicellulose-based hydrogel prepaped by glow discharge electrolysis plasma and its adsorption propertion for heavy metal ions. Fresenius Environ. Bull. 2016, 25, 1791–1798. [Google Scholar]
  6. Jin, K.; Lin, X.; Li, A. Porosint-Supported PVDF Derivatives with Good Adsorption Performance and Recyclability for the Removal of Aqueous Heavy Metal Ions. Chem. Sel. 2023, 8, 2–9. [Google Scholar] [CrossRef]
  7. Mazumder, A.; Bhattacharya, S.; Bhattacharjee, C. Role of nano-photocatalysis in heavy metal detoxification. In Nanophotocatalysis and Environmental Applications: Detoxification and Disinfection; Springer: Cham, Switzerland, 2019; pp. 1–33. [Google Scholar]
  8. Chen, J.W. Research Progress of Membrane Technology Application in Heavy Metal Wastewater Treatment. Guangdong Chem. Ind. 2009, 36, 132–136. [Google Scholar]
  9. Yu, H.; Qiu, C.; Wang, C.; Jie, J.; Sun, L.; Luo, S.; Liu, F.; Chen, J. Influence of Fenton pretreatment on heavy metal speciation and bioleaching efficiency in municipal sludge. Chin. J. Environ. Eng. 2019, 13, 725–731. [Google Scholar]
  10. Mo, Z.; Zhang, H.; Shahab, A.; Khan, F.A.; Chen, J.; Huang, C. Functionalized metal-organic framework UIO-66 nanocomposites with ultra-high stability for efficient adsorption of heavy metals: Kinetics, thermodynamics, and isothermal adsorption. J. Taiwan Inst. Chem. Eng. 2023, 146, 104778. [Google Scholar] [CrossRef]
  11. Petranovska, A.L.; Abramov, N.V.; Turanska, S.P.; Gorbyk, P.P.; Kaminskiy, A.N.; Kusyak, N.V. Adsorption of cis -dichlorodiammineplatinum by nanostructures based on single-domain magnetite. J. Nanostruct. Chem. 2015, 5, 275–285. [Google Scholar] [CrossRef]
  12. Sudhakar, A. An Efficient, Uncatalyzed, and Rapid Synthesis of Thiazoles and Aminothiazoles Under Microwave Irradiation and Investigation of Their Biological Activity. Phosphorus 2010, 185, 103–109. [Google Scholar]
  13. Zhan, K.; Wen, X.; Wang, X.; Kong, C. A method for characterization of stress concentration degree of coal mine roadway surrounding rock. J. Geophys. Eng. 2023, 4, 699–711. [Google Scholar] [CrossRef]
  14. Lai, Y.; Hao, L.; Dong, L.; Yu, S.; Liu, J. Coating zirconium oxide-nanocomposite with humic acid for recovery of mercury and chromium in hazardous waste of chemical oxygen demand test. J. Environ. Sci. 2023, 126, 40–47. [Google Scholar] [CrossRef]
  15. Zhou, L.P.; Yuan, L.; Zhang, S.Q. Advances in humic acid structures and their regulatory role in maize roots. J. Plant Nutr. Fertil. 2022, 28, 334–343. [Google Scholar]
  16. Plechanov, N. Studies of molecular weight distributions of fulvic and humic acids by gel permeation chromatography. Examination of the solute molecular composition using RI, UV, fluorescence and weight measurement as detection techniques. Org. Geochem. 1983, 5, 143–149. [Google Scholar] [CrossRef]
  17. Parsons, J.W. Humus Chemistry—Genesis, Composition, Reactions. Soil Sci. 1983, 135, 129–130. [Google Scholar] [CrossRef]
  18. Zhao, X.; Xi, B.; Tan, W.; Li, X.; Dang, Q.; Li, R. On the applicability of the ‘humic acids’ nomenclature from natural ecosystems to engineering sciences. Environ. Sci. Ecotechnol. 2021, 6, 100082. [Google Scholar] [CrossRef]
  19. Wang, T.; Liu, W.; Xiong, L.; Xu, N.; Ni, J. Influence of pH, ionic strength and humic acid on competitive adsorption of Pb(II), Cd(II) and Cr(III) onto titanate nanotubes. Chem. Eng. J. 2013, 215–216, 366–374. [Google Scholar] [CrossRef]
  20. Kort, J.; Kyasaram, R.; Cao, S.; Fu, P.; Driscoll, J.; Malek, E. Prognostic Value of Dynamic Monoclonal Protein Pattern on Probability of Myeloma Progression from the Precursor State. Blood 2021, 138, 3781. [Google Scholar] [CrossRef]
  21. Tan, K.H.; Giddens, J.E. Molecular weights and spectral characteristics of humic and fulvic acids. Geoderma 1972, 8, 221–229. [Google Scholar] [CrossRef]
  22. Zheng, E.; Zhu, Y.; Hu, J.; Zhang, Z.; Xu, T. Effects of humic acid on japonica rice production under different irrigation practices and a TOPSIS-based assessment on the Songnen Plain, China. Irrig. Sci. 2022, 40, 87–101. [Google Scholar] [CrossRef]
  23. Fan, X.J.; Yu, X.F.; Cao, T. Light Absorption and Fluorescence Characteristics of Atmospheric Water-soluble Organic Compounds and Humic-like Substances During the Winter Season in Guangzhou. Environ. Sci. 2019, 40, 532–539. [Google Scholar]
  24. Datta, A.; Choudhury, M.; Sharma, P.C.; Jat, H.S.; Jat, M.L. Stability of humic acid carbon under conservation agriculture practices. Soil Tillage Res. 2022, 216, 105240. [Google Scholar] [CrossRef]
  25. Fatima, N.; Jamal, A.; Huang, Z.; Liaquat, R.; Ahmad, B.; Haider, R.; Ali, M.I.; Shoukat, T.; ALOthman, Z.A.; Ouladsmane, M.; et al. Extraction and Chemical Characterization of Humic Acid from Nitric Acid Treated Lignite and Bituminous Coal Samples. Sustainability 2021, 13, 8969. [Google Scholar] [CrossRef]
  26. Peuravuori, J.; Simpson, A.J.; Lam, B. Structural features of lignite humic acid in light of NMR and thermal degradation experiments. J. Mol. Struct. 2007, 826, 131–142. [Google Scholar] [CrossRef]
  27. Verheyen, T.V.; Johns, R.B. Structural investigations of Australian coals—III. A 13C-NMR study on the effects of variation in rank on coal humic acids. Geochim. Cosmochim. Acta 1982, 46, 2061–2067. [Google Scholar] [CrossRef]
  28. Ricca, G.; Federico, L.; Astori, C.; Gallo, R. Structural investigations of humic acid from leonardite by spectroscopic methods and thermal analysis. Geoderma 1993, 57, 263–274. [Google Scholar] [CrossRef]
  29. Balkas, T.I.; Bastürk, O.; Gaines, A.F.; Salihoǧlu, I.; Yilmaz, A. Comparison of five humic acids. Fuel 1983, 62, 373–379. [Google Scholar] [CrossRef]
  30. Novk, F.; Estauberov, M.; Hrabal, R.; Zhang, S.Q.; Yuan, L.; Zhao, B.Q. Structural Features of Lignohumic Acids. Humic Acid 2016, 01, 24–32. [Google Scholar]
  31. Peuravuori, J.; Zbankova, P.; Pihlaja, K. Aspects of structural features in lignite and lignite humic acids. Fuel Process. Technol. 2006, 87, 829–839. [Google Scholar] [CrossRef]
  32. Clemow, L.M.; Favas, G.; Jackson, W.R.; Marshall, M.; Redlich, P.J. Humic acids and methylated humic acids as models for reactions of brown coal with CO/H2O and with H2. Fuel 1999, 78, 567–572. [Google Scholar] [CrossRef]
  33. Qian, S.; Ding, W.; Li, Y.; Liu, G.; Sun, J.; Ding, Q. Characterization of humic acids derived from Leonardite using a solid-state NMR spectroscopy and effects of humic acids on growth and nutrient uptake of snap bean. Chem. Speciat. Bioavailab. 2015, 27, 156–161. [Google Scholar] [CrossRef]
  34. Chi, X.Y.; Zeng, L.X.; Du, Y.J.; Huang, J.Q.; Kang, Y.; Luo, J.W.; Zhang, Q.Y. Adsorption of levofloxacin on natural zeolite: Effects of ammonia nitrogen and humic acid. Water Sci. Technol. 2022, 85, 2928–2944. [Google Scholar] [CrossRef]
  35. Xing, Z.B. Sorption of phenanthrene by nanosized alumina coated with sequentially extracted humic acids. Environ. Sci. Pollut. Res. 2010, 2, 17–27. [Google Scholar]
  36. Dong, W.X.; Tian, H.T.; Ping, Z.Y. Complexation characteristics of humic acids with Fe2+ and their fractions from soil with different fertilization II—The complexation of humic acid fractions with Fe2+. Plant Nutr. Fertil. Sci. 2001, 7, 139–144. [Google Scholar]
  37. Sievers, D.A.; Kuhn, E.M.; Tucker, M.P.; Mcmillan, J.D. Effects of Dilute-Acid Pretreatment Conditions on Filtration Performance of Corn Stover Hydrolyzate. Bioresour. Technol. 2017, 243, 474–480. [Google Scholar] [CrossRef] [PubMed]
  38. Liu, H.; Qi, Z.; Liu, C. Inhibition mechanisms of humic acid and protein on the degradation of sulfamethazine by horseradish peroxidase. Colloids Surf. A Physicochem. Eng. Asp. 2021, 629, 127473. [Google Scholar] [CrossRef]
  39. Wang, J.; Lu, S.; Guo, L.; Wang, P.; He, C.; Liu, D.; Bian, H.; Sheng, L. Effects of polystyrene nanoplastics with different functional groups on rice (Oryza sativa L.) seedlings: Combined transcriptome, enzymology, and physiology. Sci. Total Environ. 2022, 834, 155092. [Google Scholar] [CrossRef]
  40. Fujitake, N.; Kusumoto, A.; Tsukamoto, M.; Noda, Y.; Suzuki, T.; Otsuka, H. Properties of soil humic substances in fractions obtained by sequential extraction with pyrophosphate solutions at different pHs. Soil Sci. Plant Nutr. 1999, 45, 349–358. [Google Scholar] [CrossRef]
  41. Hardy, E.; Rodriguez, C.; Trujillo, L.E. Lipopolysaccharide (LPS) and Protein-LPS complexes: Detection and Characterization by Gel Electrophoresis, Mass Spectrometry and Bioassays. Biol. Med. 2016, 8, 2–7. [Google Scholar] [CrossRef]
  42. Klučáková, M. Characterization of pH-fractionated humic acids with respect to their dissociation behaviour. Environ. Sci. Pollut. Res. 2016, 23, 7722–7731. [Google Scholar] [CrossRef]
  43. Zhang, S.; Yuan, L.; Li, W.; Lin, Z.; Li, Y.; Hu, S.; Zhao, B. Characterization of pH-fractionated humic acids derived from Chinese weathered coal. Chemosphere 2017, 166, 334–342. [Google Scholar] [CrossRef] [PubMed]
  44. Zhao, J.; Li, Y.; Li, Y.; Zhang, K.; Zhang, H.; Li, Y. Effects of humic acid on sludge performance, antibiotics resistance genes propagation and functional genes expression during Cu(II)-containing wastewater treatment via metagenomics analysis. Bioresour. Technol. 2021, 323, 124575. [Google Scholar] [CrossRef]
  45. Hermosin, B.; Trubetskoj, O.A.; Trubetskaya, O.E.; Saiz-Jimenez, C. Thermally assisted hydrolysis and methylation of humic fractions obtained by polyacrylamide gel electrophoresis. J. Anal. Appl. Pyrolysis 2001, 58, 341–347. [Google Scholar] [CrossRef]
  46. Saiz-Jimenez, C.; Hermosin, B.; Trubetskaya, O.E.; Reznikova, O.I.; Trubetskoj, O.A. Thermochemolysis of genetically different soil humic acids and their fractions obtained by tandem size exclusion chromatography–polyacrylamide gel electrophoresis. Geoderma 2006, 131, 22–32. [Google Scholar] [CrossRef]
  47. Ouatmane, A.; D’Orazio, V.; Hafidi, M.; Senesi, N. Chemical and Physicochemical Characterization of Humic Acid-like Materials from Composts. Compost Sci. Util. 2002, 1, 39–46. [Google Scholar] [CrossRef]
  48. Drosos, M.; Leenheer, J.A.; Avgeropoulos, A.; Deligiannakis, Y. H-binding of size-and polarity-fractionated soil and lignite humic acids after removal of metal and ash components. Environ. Sci. Pollut. Res. 2014, 21, 3963–3971. [Google Scholar] [CrossRef] [PubMed]
  49. Jones, M.N.; Bryan, N.D. Colloidal properties of humic substances. Adv. Colloid Interface Sci. 1998, 78, 1–48. [Google Scholar] [CrossRef]
  50. And, S.K.; Xing, B. Humic Acid Fractionation upon Sequential Adsorption onto Goethite. Langmuir 2008, 24, 2525–2531. [Google Scholar]
  51. Monsallier, J.M. Influence of Humic Acid Size on Actinide Complexation; The Florida State University: Tallahassee, FL, USA, 1998. [Google Scholar]
  52. Tanaka, T. Functional groups and reactivity of size-fractionated Aldrich humic acid. Thermochim. Acta 2012, 532, 60–64. [Google Scholar] [CrossRef]
  53. Masoodi, H.R.; Bagheri, S.; Mohammadi, M.; Zakarianezhad, M.; Makiabadi, B. The influence of cation-π and anion-π interactions on some NMR data of s-triazine… HF hydrogen bonding: A theoretical study. Chem. Phys. Lett. 2013, 588, 4–10. [Google Scholar] [CrossRef]
  54. Zhang, J.; Qiao, Z.A.; Mahurin, S.M.; Jiang, X.; Chai, S.H.; Lu, H.; Nelson, K.; Dai, S. Hypercrosslinked Phenolic Polymers with Well-Developed Mesoporous Frameworks. Angew. Chem. Int. Ed. 2015, 54, 4582–4586. [Google Scholar] [CrossRef]
  55. Vorov, O.K.; Livesay, D.R.; Jacobs, D.J. Nonadditivity in Conformational Entropy upon Molecular Rigidification Reveals a Universal Mechanism Affecting Folding Cooperativity. Biophys. J. 2011, 100, 1129–1138. [Google Scholar] [CrossRef]
  56. Chappaz, A.; Curtis, P.J. Integrating empirically dissolved organic matter quality for WHAM VI using the DOM optical properties: A case study of Cu-Al-DOM interactions. Environ. Sci. Technol. 2013, 47, 2001–2007. [Google Scholar] [CrossRef]
  57. El-Wakeel, S.T.; Radwan, E.K.; Ghafar, H.H.A.; Moursy, A.S. Humic acid-carbon hybrid material as lead(II) ions adsorbent. Desalination Water Treat. 2017, 74, 216–223. [Google Scholar] [CrossRef]
  58. Christl, I.; Kretzschmar, R. Relating Ion Binding by Fulvic and Humic Acids to Chemical Composition and Molecular Size. Environ. Sci. Technol. 2001, 35, 2505–2511. [Google Scholar] [CrossRef]
  59. Wang, Y.; Envelope, Z.H.P.; Sheng, L.; Zhao, M.; Feng, J. Adsorption of lead ions from aqueous solution using NH4H2PO4 modified humic acid residue. Environ. Technol. Innov. 2022, 28, 102920. [Google Scholar] [CrossRef]
  60. Kinniburgh, D.G.; Milne, C.J.; Benedetti, M.F.; Pinheiro, J.P.; Filius, J.; Koopal, L.K.; Van Riemsdijk, W.H. Metal ion binding by humic acid: Application of the NICA-Donnan model. Environ. Sci. Technol. 1996, 30, 1687–1698. [Google Scholar] [CrossRef]
  61. Qiong, Z. Study on the Preparation of Activated Carbon from Coal and Its Adsorption Properties; Shanxi Chemical Industry: Taiyuan, China, 2018. [Google Scholar]
  62. Wang, R.Z.; Huang, D.L.; Liu, Y.G. Investigating the adsorption behavior and the relative distribution of Cd2+ sorption mechanisms on biochars by different feedstock. Bioresour. Technol. 2018, 261, 265–271. [Google Scholar] [CrossRef] [PubMed]
  63. Huang, L.; Mao, N.; Shuai, Q. Efficient removal of tetracycline from aqueous solution by covalent organic frameworks derived porous carbon. J. Environ. Chem. Eng. 2021, 9, 104842. [Google Scholar] [CrossRef]
  64. Jana, A.; Roy, O.; Ravuru, S.S.; De, S. Tuning of graphene oxide intercalation in magnesium aluminium layered double hydroxide and their immobilization in polyacrylonitrile beads by single step mussel inspired phase inversion: A super adsorbent for lead. Chem. Eng. J. 2019, 391, 123587. [Google Scholar] [CrossRef]
  65. Chen, D.; Shen, W.; Wu, S.; Chen, C.; Luo, X.; Guo, L. Ion exchange induced removal of Pb(ii) by MOF-derived magnetic inorganic sorbents. Nanoscale 2016, 8, 7172–7179. [Google Scholar] [CrossRef] [PubMed]
  66. Liu, Y.L.; Xu, J.; Cao, Z.; Fu, R.; Zhou, C.; Wang, Z.; Xu, X. Adsorption behavior and mechanism of Pb(II) and complex Cu(II) species by biowaste-derived char with amino functionalization. J. Colloid Interface Sci. 2020, 559, 215–225. [Google Scholar] [CrossRef] [PubMed]
  67. Xu, J.; Cao, Z.; Zhang, Y.; Yuan, Z.; Lou, Z.; Xu, X.; Wang, X. A review of functionalized carbon nanotubes and graphene for heavy metal adsorption from water: Preparation, application, and mechanism. Chemosphere 2018, 195, 351–364. [Google Scholar] [CrossRef]
  68. Wang, L.-M.; Li, Y.-H. Preparation of a Novel Bi-imprinted Adsorbent for Pb(II) and Its Application to FAAS Determination of Trace Amount of Lead in Water Sample. Ptca (Part B Chem. Anal.) 2014, 50, 589–594. [Google Scholar]
  69. Cimino, R.; Cychosz, K.A.; Thommes, M.; Neimark, A.V. Experimental and theoretical studies of scanning adsorption–desorption isotherms. Colloids Surf. A Physicochem. Eng. Asp. 2013, 437, 76–89. [Google Scholar] [CrossRef]
  70. Song, L.; Zhao, R.; Yun, D.; Lu, P.; He, J.; Wang, X. Influence of Co(II), Ni(II), tartrate, and ethylenediaminetetraacetic acid on Cu(II) adsorption onto a polyvinylidene fluoride-based chelating membrane. Toxicol. Environ. Chem. 2014, 96, 362–378. [Google Scholar] [CrossRef]
  71. Fu, L.; Liu, F.; Ma, Y.; Tao, X.; Ling, C.; Li, A.; Shuang, C.; Li, Y. High-efficient technique to simultaneous removal of Cu(II), Ni(II) and tannic acid with magnetic resins: Complex mechanism behind integrative application. Chem. Eng. J. 2015, 63, 83–91. [Google Scholar] [CrossRef]
  72. Lv, J.; Chen, J. Synthesis of Core-Shell Nanoscale Zero Valent Iron under Air Atmosphere and Its Enlightenment for Groundwater In-Situ Remediation. Geol. Sci. Technol. Inf. 2017, 3, 242–248. [Google Scholar]
  73. Wang, N.; Xu, X.; Li, H.; Yuan, L.; Yu, H. Enhanced Selective Adsorption of Pb(II) from Aqueous Solutions by One-Pot Synthesis of Xanthate-Modified Chitosan Sponge: Behaviors and Mechanisms. Ind. Eng. Chem. Res. 2016, 55, 1222–1231. [Google Scholar] [CrossRef]
  74. Dragan, E.S.; Loghin, D.F.A. Fabrication and characterization of composite cryobeads based on chitosan and starches-g-PAN as efficient and reusable biosorbents for removal of Cu2+, Ni2+, and Co2+ ions. Int. J. Biol. Macromol. 2018, 120, 1872–1883. [Google Scholar] [CrossRef]
  75. Sharma, P.; Singh, A.K.; Shahi, V.K. Selective Adsorption of Pb(II) from Aqueous Medium by Cross-Linked Chitosan-Functionalized Graphene Oxide Adsorbent. ACS Sustain. Chem. Eng. 2019, 7, 1427–1436. [Google Scholar] [CrossRef]
  76. Neiber, R.R.; Galhoum, A.A.; Sayed, E.T.E.; Guibal, E.; Xin, J.; Lu, X. Selective lead (II) sorption using aminophosphonate-based sorbents: Effect of amine linker, characterization and sorption performance. Chem. Eng. J. 2022, 442, 136300–136317. [Google Scholar] [CrossRef]
  77. Liu, C.; Bai, R.; Ly, Q.S. Selective removal of copper and lead ions by diethylenetriamine-functionalized adsorbent: Behaviors and mechanisms. Water Res. 2008, 42, 1511–1522. [Google Scholar] [CrossRef]
  78. Bassam, R.; Alouani, M.E.; Jarmouni, N.; Maissara, J.; Chbihi, M.E.M.; Belaaouad, S. Investigation of competitive adsorption and desorption of heavy metals from aqueous solution using raw rock: Characterization kinetic, isotherm, and thermodynamic. Mater. Today Proc. 2021, 52, 158–165. [Google Scholar] [CrossRef]
  79. Gao, H.; Sun, Y.; Zhou, J.; Xu, R.; Duan, H. Mussel-Inspired Synthesis of Polydopamine-Functionalized Graphene Hydrogel as Reusable Adsorbents for Water Purification. ACS Appl. Mater. Interfaces 2013, 5, 425–432. [Google Scholar] [CrossRef] [PubMed]
  80. Zhan, W.; Gao, L.; Fu, X.; Siyar, S.H.; Sui, G.; Yang, X. Green synthesis of amino-functionalized carbon nanotube-graphene hybrid aerogels for high performance heavy metal ions removal. Appl. Surf. Sci. 2019, 467–468, 1122–1133. [Google Scholar] [CrossRef]
  81. Xue, S.; Xiao, Y.; Wang, G.; Fan, J.; Wan, K.; He, Q.; Gao, M.; Miao, Z. Adsorption of heavy metals in water by modifying Fe3O4 nanoparticles with oxidized humic acid. Colloids Surf. A Physicochem. Eng. Asp. 2021, 616, 126333–126342. [Google Scholar] [CrossRef]
  82. Parajuli, D.; Adhikari, C.R.; Kuriyama, M.; Kawakita, H.; Ohto, K.; Inoue, K.; Funaoka, M. Selective recovery of gold by novel lignin-based adsorption gels. Ind. Eng. Chem. 2006, 45, 8–14. [Google Scholar] [CrossRef]
  83. Zou, L.; Shao, P.; Zhang, K.; Yang, L.; You, D.; Shi, H.; Pavlostathis, S.G.; Lai, W.; Liang, D.; Luo, X. Tannic acid-based adsorbent with superior selectivity for lead(II) capture: Adsorption site and selective mechanism. Chem. Eng. J. 2019, 364, 1–7. [Google Scholar] [CrossRef]
  84. Liu, X.; Fu, L.; Liu, H.; Zhang, D.; Xiong, C.; Wang, S.; Zhang, L. Design of Zr-MOFs by Introducing Multiple Ligands for Efficient and Selective Capturing of Pb(II) from Aqueous Solutions. ACS Appl. Mater. Interfaces 2023, 15, 5974–5989. [Google Scholar] [CrossRef]
  85. Tang, J.; Chen, Y.; Zhao, M.; Wang, S.; Zhang, L. Phenylthiosemicarbazide-functionalized UiO-66-NH2 as highly efficient adsorbent for the selective removal of lead from aqueous solutions. J. Hazard. Mater. 2021, 413, 125289. [Google Scholar] [CrossRef]
  86. Huang, Y.; Hu, C.; An, Y.; Xiong, Z.; Hu, X.; Zhang, G.; Zheng, H. Magnetic phosphorylated chitosan composite as a novel adsorbent for highly effective and selective capture of lead from aqueous solution. J. Hazard. Mater. 2020, 405, 124195. [Google Scholar] [CrossRef]
  87. Nuzahat, H.; Wei, C. Structural response of humic acid upon binding with lead: A spectroscopic insight. Sci. Total Environ. 2018, 43, 479–485. [Google Scholar]
  88. Kadlubanski, P.; Calderón-Mojica, K.; Rodriguez, W.A.; Majumdar, D.; Roszak, S.; Leszczynski, J. Role of the Multipolar Electrostatic Interaction Energy Components in Strong and Weak Cation-pi Interactions. J. Phys. Chem. A 2013, 117, 7989–8000. [Google Scholar] [CrossRef]
  89. Appel, C.; Ma, L.Q.; Rhue, R.D.; Reve, W. Selectivities of Potassium-Calcium and Potassium-Lead Exchange in Two Tropical Soils. Soil Sci. Soc. Am. J. 2003, 67, 1707–1714. [Google Scholar] [CrossRef]
  90. Teppen, B.J.; Miller, D.M. Hydration Energy Determines Isovalent Cation Exchange Selectivity by Clay Minerals. Soil Sci. Soc. Am. J. 2006, 70, 31–40. [Google Scholar] [CrossRef]
  91. Charles, S.; Teppen, B.J.; Li, H.; Laird, D.A.; Boyd, S.A. Exchangeable Cation Hydration Properties Strongly Influence Soil Sorption of Nitroaromatic Compounds. Soil Sci. Soc. Am. J. 2006, 70, 1470–1479. [Google Scholar] [CrossRef]
  92. Zhuang, Y.; Yu, F. Alginate/graphene double-network nanocomposite hydrogel beads with low-swelling, enhanced mechanical properties, and enhanced adsorption capacity. J. Mater. Chem. A 2016, 4, 10885–10892. [Google Scholar] [CrossRef]
  93. Liu, J.T.; Xiao, G. 3D graphene/delta-MnO2 aerogels for highly efficient and reversible removal of heavy metal ions. J. Mater. Chem. A 2016, 4, 1970–1979. [Google Scholar] [CrossRef]
  94. Sabaté, J.; Labanda, J.; Llorens, J. Influence of coion and counterion size on multi-ionic solution nanofiltration. J. Membr. Sci. 2009, 345, 298–304. [Google Scholar] [CrossRef]
  95. Wang, D.X.; Su, M.; Yu, Z.Y.; Wang, X.L.; Ando, M.; Shintani, T. Separation performance of a nanofiltration membrane influenced by species and concentration of ions. Desalination 2005, 175, 219–225. [Google Scholar] [CrossRef]
  96. Gao, L.Y.; Deng, J.H.; Huang, G.F.; Li, K.; Cai, K.Z.; Liu, Y.; Huang, F. Relative distribution of Cd2+ adsorption mechanisms on biochars derived from rice straw and sewage sludge. Bioresour. Technol. 2019, 272, 114–122. [Google Scholar] [CrossRef] [PubMed]
  97. 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] [PubMed]
  98. Zhang, T.; Zhu, X.X.; Shi, L.N.; Li, J.F.; Li, S.J.; Lü, J.H.; Li, Y.M. Efficient removal of lead from solution by celery-derived biochars rich in alkaline minerals. Bioresour. Technol. 2017, 235, 185–192. [Google Scholar] [CrossRef]
  99. Li, A.Y.; Deng, H.; Jiang, Y.H.; Ye, C.H.; Yu, B.G.; Zhou, X.L.; Ma, A.Y. Superefficient Removal of Heavy Metals from Wastewater by Mg-Loaded Biochars: Adsorption Characteristics and Removal Mechanisms. Langmuir 2020, 36, 9160–9174. [Google Scholar] [CrossRef]
  100. Deng, Y.; Li, X.; Ni, F.; Liu, Q.; Yang, Y.; Wang, M.; Ao, T.; Chen, W. Synthesis of Magnesium Modified Biochar for Removing Copper, Lead and Cadmium in Single and Binary Systems from Aqueous Solutions: Adsorption Mechanism. Water 2021, 13, 599. [Google Scholar] [CrossRef]
  101. Kalinichev, A.G. Complexation with metal ions and colloidal aggregation of natural organic matter in aqueous solutions: A computational molecular modeling perspective (invited talk). In Proceedings of the ANDRA Workshop on Organic Matter in Clay Rock, Nantes, France, 20–24 September 2010. [Google Scholar]
  102. Kinniburgh, D.G.; van Riemsdijk, W.H.; Koopal, L.K.; Borkovec, M.; Benedetti, M.F.; Avena, M.J. Ion binding to natural organic matter: Competition, heterogeneity, stoichiometry and thermodynamic consistency. Colloids Surf. A 1999, 151, 147–166. [Google Scholar] [CrossRef]
  103. Yan, M.; Lu, Y.; Gao, Y.; Benedetti, M.F.; Korshin, G.V. In-Situ Investigation of Interactions between Magnesium Ion and Natural Organic Matter. Environ. Sci. Technol. 2015, 9, 8323–8329. [Google Scholar] [CrossRef]
  104. Pinheiro, J.P.; Mota, A.M.; Benedetti, M.F. Effect of aluminum competition on lead and cadmium binding to humic acids at variable ionic strength. Environ. Sci. Technol. 2000, 34, 5137–5143. [Google Scholar] [CrossRef]
  105. Marsac, R.; Davranche, M.; Gruau, G.; Dia, A.; Coz, B.L. Aluminium competitive effect on rare earth elements binding to humic acid. Geochim. Cosmochim. Acta 2012, 89, 1–9. [Google Scholar] [CrossRef]
  106. Xue, S.; Wang, Y.; Bo, W.; Wan, K.; Miao, Z. Calcium-doped magnetic humic acid nano particles for the efficient removal of heavy metals from wastewater: The role of Ca. Environ. Technol. 2023, 233, 1–16. [Google Scholar] [CrossRef] [PubMed]
  107. Alka, S.; Dipal, K. Response of plant species grown on central verge area of nh-8 near Navsari, India. J. Environ. Res. Dev. Environ. Res. Dev. 2012, 7, 1016–1020. [Google Scholar]
  108. Qiao, X.X.; Liu, G.F.; Wang, J.T.; Zhang, Y.Q.; Lü, J. Highly Efficient and Selective Removal of Lead Ions from Aqueous Solutions by Conjugated Microporous Polymers with Functionalized Heterogeneous Pores. Cryst. Growth Des. 2019, 20, 337–344. [Google Scholar] [CrossRef]
  109. Wang, L.; Hung, Y.T.; Shammas, N. Environmental Engnieering. In Advanced Physicochemical Treatment Processes; Humana: Totowa, NJ, USA, 2006; pp. 1–44. [Google Scholar]
  110. Chen, H.; Li, W.; Wang, J.; Xu, H.; Liu, Y.; Zhang, Z.; Li, Y.; Zhang, Y. Adsorption of cadmium and lead ions by phosphoric acid-modified biochar generated from chicken feather: Selective adsorption and influence of dissolved organic matter. Bioresour. Technol. 2019, 292, 121948. [Google Scholar] [CrossRef] [PubMed]
  111. Masoud, M.S.; Zidan, A.A.; El Zokm, G.M.; Okbah, M.A. Humic acid and nano-zeolite NaX as low cost and eco-friendly adsorbents for removal of Pb (II) and Cd (II) from water: Characterization, kinetics, isotherms and thermodynamic studies. Biomass Convers. Biorefin. 2024, 14, 3615–3632. [Google Scholar] [CrossRef]
  112. Liu, J.F.; Zhao, Z.S.; Jiang, G.B. Coating Fe3O4 magnetic nanoparticles with humic acid for high efficient removal of heavy metals in water. Environ. Sci. Technol. 2008, 42, 6949–6954. [Google Scholar] [CrossRef]
  113. Xue, S.; Fan, J.; Wan, K.; Wang, G.; Xiao, Y.; Bo, W.; Gao, M.; Miao, Z. Calcium-Modified Fe3O4 Nanoparticles Encapsulated in Humic Acid for the Efficient Removal of Heavy Metals from Wastewater. Langmuir 2021, 37, 10994–11007. [Google Scholar] [CrossRef] [PubMed]
  114. Xue, S.W.; Miao, Z.Y.; Gao, M.Q.; Wan, K.J. Structural analysis of lignite-derived humic acid and its microscopic interactions with heavy metal ions in aqueous solution. Sci. Total Environ. 2023, 897, 165385. [Google Scholar] [CrossRef]
  115. Nassar, N.N. Kinetics, equilibrium and thermodynamic studies on the adsorptive removal of nickel, cadmium and cobalt from wastewater by superparamagnetic iron oxide nanoadsorbents. Can. J. Chem. Eng. 2012, 90, 1231–1238. [Google Scholar] [CrossRef]
  116. Nassar, N.N. Rapid removal and recovery of Pb(II) from wastewater by magnetic nanoadsorbents. J. Hazard. Mater. 2010, 184, 538–546. [Google Scholar] [CrossRef] [PubMed]
  117. Tahir, M.U.; Su, X.M.; Zhao, Y.; Liao, R. Interfaces, Preparation of hydroxypropyl-cyclodextrin-graphene/Fe3O4 and its adsorption properties for heavy metals. Surf. Interfaces 2019, 16, 43–49. [Google Scholar] [CrossRef]
Figure 1. Application of lignite materialization.
Figure 1. Application of lignite materialization.
Separations 11 00080 g001
Figure 2. Research status of HA molecular structure.
Figure 2. Research status of HA molecular structure.
Separations 11 00080 g002
Figure 3. Interaction between HA and metal ions.
Figure 3. Interaction between HA and metal ions.
Separations 11 00080 g003
Table 1. Characteristics of common alkali and alkaline earth metal ions [94,95].
Table 1. Characteristics of common alkali and alkaline earth metal ions [94,95].
IonsRadius
(Å)
Hydration
Radius (Å)
Hydration Energy
(kJ mol−1)
Diffusion Coefficient
(10−9 m2 s−1)
Stokes Radius
(nm)
K+1.333.31−2951.9570.124
Na+0.953.58−3651.3330.183
Ca2+0.994.12−15050.7180.307
Mg2+0.654.28−18300.7060.345
Table 2. Cation preference for the main functional groups of natural organic matter.
Table 2. Cation preference for the main functional groups of natural organic matter.
Functional Groups/Ions–COOH–OH–NH2–SHReferences
Na+N/AN/AN/A[101]
Ca++ (low pH)+ (high pH)+N/A[102]
Mg+++N/AN/A[103]
Al+++N/AN/A[104,105]
Note: the more symbols (+), the stronger the affinity of functional groups and structures for Na+/Ca2+/Mg2+/Al3+.
Table 3. Studies on HAs for Pb2+ removal.
Table 3. Studies on HAs for Pb2+ removal.
ReferencesYearsModification MethodsAdsorption Capacity (mg/g)Conclusions
[111]2022-Pb2+: 23.11
Cd2+: 12.00
HA has an affinity for Pb2+
[112]2021Magnetic HA nanoparticlesPb2+: 105.60
Cu2+: 67.43
Cd2+: 65.23
HA exhibits selectivity for Pb2+
[113]2021Calcium-modified Fe3O4 nanoparticles encapsulated in HAPb2+: 208.33
Cu2+: 98.33
Cd2+: 99.01
Ca-ion exchange is the main mechanism for the selective adsorption of Pb2+
[106]2023Calcium-doped magnetic HA nano particlesPb2+: 278.65
Cu2+: 154.31
Cd2+: 145.55
Ca activates the functional groups of HA
[114]2023-Pb2+: 67.67Carboxylic acid is the key functional group affecting the selective adsorption of HA
[115]2012Fe3O4Pb2+: 29.00
Cd2+: 18.60
-
[116]2010Resin microspheresPb2+: 99.19
Cu2+: 45.80
Cd2+: 13.75
-
[117]2019Hydroxypropyl-cyclodextrin-graphenePb2+: 99.19
Cu2+: 45.80
-
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xue, S.; Hu, Y.; Wan, K.; Miao, Z. Exploring Humic Acid as an Efficient and Selective Adsorbent for Lead Removal in Multi-Metal Coexistence Systems: A Review. Separations 2024, 11, 80. https://doi.org/10.3390/separations11030080

AMA Style

Xue S, Hu Y, Wan K, Miao Z. Exploring Humic Acid as an Efficient and Selective Adsorbent for Lead Removal in Multi-Metal Coexistence Systems: A Review. Separations. 2024; 11(3):80. https://doi.org/10.3390/separations11030080

Chicago/Turabian Style

Xue, Shuwen, Yunhu Hu, Keji Wan, and Zhenyong Miao. 2024. "Exploring Humic Acid as an Efficient and Selective Adsorbent for Lead Removal in Multi-Metal Coexistence Systems: A Review" Separations 11, no. 3: 80. https://doi.org/10.3390/separations11030080

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop