Next Article in Journal
How Do Bank Customers’ Perceptions of CSR Influence Marketing Outcomes: Their Trust, Identification, and Commitment?
Previous Article in Journal
Anaerobic Digestion of Poultry Droppings in Semi-Continuous Mode and Effect of Their Co-Digestion with Physico-Chemical Sludge on Methane Yield
Previous Article in Special Issue
Knowledge: A Factor for Acceptance of Insects as Food
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Agricultural Byproducts Used as Low-Cost Adsorbents for Removal of Potentially Toxic Elements from Wastewater: A Comprehensive Review

1
National Research & Development Institute for Food Bioresources, 021102 Bucharest, Romania
2
Agrovet SA, 014354 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(7), 5999; https://doi.org/10.3390/su15075999
Submission received: 1 February 2023 / Revised: 25 March 2023 / Accepted: 27 March 2023 / Published: 30 March 2023
(This article belongs to the Special Issue Food Choice and Environmental Concerns)

Abstract

:
Potentially toxic elements (PTEs) are ubiquitous chemical compounds in the environment due to contamination of air, water, or soil. They are primarily sourced from fossil fuel combustion, mining and smelting, electroplating, dyes and pigments, agricultural treatments, and plastic and metallic industries. These chemical contaminants can produce various adverse effects when they enter the human body and can also affect crops and aquatic ecosystems. To address these issues, researchers are developing various techniques, including ion exchange, membrane filtration, photocatalysis, electrochemical methods, bioadsorption, and combinations of these processes, to reduce the levels of these contaminants, especially from wastewater. Among these methods, bioadsorption has gained much attention due to its high efficiency, low cost, and abundance of adsorbent materials. Agricultural byproducts used as biosorbents include rice husk and bran, citrus peel, banana peel, coconut husk, sugarcane bagasse, soybean hulls, walnut and almond shells, coconut fiber, barley straws, and many others. Biosorption capacity can be described using adsorption kinetic models such as Elovich, Ritchie’s, and pseudo-second-order models, as well as different adsorption isotherm models such as Freundlich, Langmuir, Temkin isotherm, and BET models. Both conventional processes and adsorption models are influenced by parameters such as pH, agitation speed, contact time, particle size, concentration of the adsorbent material, initial concentration of the contaminant, and the type of modifying agent used. This review paper aims to examine the low-cost adsorbents and their removal efficiency and bioadsorption capacity for different PTEs present in wastewater, and their potential as decontamination methods.

1. Introduction

Potentially toxic elements encompass both nonmetals, such as arsenic (As) and selenium (Se), and metals, including cadmium (Cd), lead (Pb), mercury (Hg), aluminum (Al), chromium (Cr), cobalt (Co), and antimony (Sb), as well as copper (Cu), nickel (Ni), and zinc (Zn), which are micronutrients capable of inducing adverse effects in living organisms when present in excessive concentrations [1]. Despite being present in the Earth’s crust since its formation, pollution with potentially toxic elements, originating from both natural and anthropogenic sources, poses a significant concern. Anthropogenic sources of contamination, such as metal industries (mining, smelting, foundries), automobile exhaust, industrial waste, agricultural treatments (inorganic fertilizers, pesticides, insecticides), wastewater irrigation, and others, represent the primary contributors to pollution. Natural sources of pollution include volcanic activity, geological weathering, soil erosion, metal corrosion, metal evaporation from contaminated water, and sediment resuspension [2].
Wastewaters are the primary source of pollution for both soil and surface waters. They are often used for irrigation without adequate treatment, leading to contamination. Even treated wastewaters can still contain substances that have passed through biological treatment plants or were discharged into water bodies, contributing to pollution [1].
Approximately 20 million hectares of arable land is being irrigated with wastewater, which has the potential to cause the accumulation of toxic metals in soil, resulting in soil contamination [1]. Additionally, this practice can lead to pollution of crops, which may compromise food quality and safety [3]. Large quantities of metals can reach the human body through the cultivation of plants and crops on contaminated soil, making this pathway a major route of exposure to potentially toxic elements [3]. In addition, toxic metals can enter the body in smaller amounts through ingestion of contaminated water, dermal contact, or inhalation [4].
Toxic metals may accumulate in vital organs (such as the liver, brain, kidney, and heart) of the human body, leading to disturbances in their normal functioning and blocking vital activity. Continuous exposure to potentially toxic elements can cause an internal imbalance, as they may be used as substitutes for the body’s essential elements [4]. Table 1 presents the health effects of the most important potentially toxic metals.
The contamination of air, water, and food with potentially toxic elements is a global issue, and researchers are working on developing new and effective methods for removing them from wastewater sources. The elimination of these elements from water is crucial to protecting human health due to their negative impact [11].
Agricultural byproducts are environmentally friendly and exhibit potential biosorption capacity for potentially toxic metals, but also for other chemical contaminants (nitrates, pesticides, colorants, bisphenol A) harmful to living organisms. The major advantages of agricultural waste include (i) high regeneration and reusability of the spent adsorbents, (ii) reduced sludge generation, (iii) high efficiency, (iv) low cost, (v) availability of many adsorbent materials (husks, shells, peels, vegetable waste, leaves, stalks), (vi) strong affinity, and (vii) high selectivity for many metal ions (due to the availability of functional groups from the adsorbent surface). Agricultural adsorbents have demonstrated high removal performances and adsorption capacities due to their porous structure and high amount of various functional groups (hydroxyl, carboxyl) capable of binding ionic compounds. Also, agricultural adsorbents can be recovered and reused without posing severe impact on environment, being consistent with concepts of effective, innovative and suitable waste management. All these aspects make them superior to conventional removal processes [12].
The primary objectives of this review study are to (1) characterize the mitigation techniques used for the removal of potentially toxic elements from wastewaters; (2) present the effects of various parameters, such as pH, adsorbent dosage, contact time, initial concentration of the contaminant, ionic strength, temperature, and modifying agents on adsorption capacity and removal efficiency; and (3) investigate the adsorption capacities and removal percentages of different agricultural byproducts for various potentially toxic elements.

2. Estimation of Removal Percentage and Adsorption Capacity

“Removal percent means a percentage expression of the removal efficiency across a treatment facility for a given pollutant or contaminant parameter as determined from a comparison of influent and effluent concentrations for a given time period” [13].
Removal efficiency can be estimated using Equation (1).
R e m o v a l   e f f i c i e n c y % = C i C e C i × 100
where Ci is the initial concentration of the adsorbate (mg L−1) and Ce is the final concentration of the adsorbate (mg L−1) [14].
“Adsorption capacity is the amount of adsorbate taken up by the adsorbent per unit mass (or volume) of the adsorbent” [15] and can be calculated using Equation (2).
A d s o r p t i o n   c a p a c i t y , q e m g g 1 = C i C e M s × V
where Ci is the initial concentration of the adsorbate (mg L−1) and Ce is the final concentration of the adsorbate (mg L−1), V is the volume of the adsorbate solution (L), Ms is the adsorbent dose (g) [14].
The adsorption capacity of an adsorbent material can be assessed using various adsorption isotherms or kinetic models, which estimate the quantity of target metal adsorbed onto the surface of the material per unit weight, under equilibrium conditions and at a constant temperature [16]. The most commonly used isotherms are Langmuir and Freundlich, but there are several other isotherms and models available, such as pseudo-first-order (PFO), pseudo-second-order (PSO), Dubinin–Radushkevich, Elovich, Redlich–Peterson, Flory–Huggins, Henry, Temkin, Radke and Prausnitz, Halsey, Sips, Toth, Harkins–Jura, Koble and Corrigan, and many others [16,17,18]. Table 2 presents the equations for the most commonly used isotherms and models for bioadsorption studies.

3. Removal Techniques of Potentially Toxic Metals

Wastewater from different industrial sectors contains various chemical contaminants, such as dyes, pigments, potentially toxic elements, and pesticides, among many other hazardous chemicals. The elimination of these compounds is a necessary step for wastewater treatment [20]. Techniques available for mitigating potentially toxic elements from wastewater include conventional methods such as membrane filtration, chemical precipitation, ion exchange, chemical coagulation, bioremediation, flocculation, and adsorption [21], as well as modern methods based on nanotechnology [11]. Adsorption-based techniques are more effective and economical than other conventional methods, due to the large surface area of solid adsorbents, the large number of active sites, small size, porosity, easy preparation, amorphous nature, and low cost [11].
Conventional methods can be grouped into membrane-based filtration methods, chemical-based separation methods, electricity-based separation methods, photocatalytic-based separation methods, and adsorption-based separation methods.
An overview of the most important techniques for heavy metals removal is presented in Figure 1.
The membrane-based filtration process is a separation technique that uses a porous membrane and a driving force applied across the membrane to separate contaminants from solution, based on the substrate size, membrane pore size, solution concentration, and applied pressure [23]. It is a pressure-based removal process that can be enhanced by treating membranes with various chemical compounds [24]. These treatments, based on use of membranes in the form of a complex structure that contains elements at a nanometric scale, showed excellent results for mitigation of potentially toxic elements from wastewater [21]. Depending on the size of the pores, membrane filtration processes can be grouped as ultrafiltration, nanofiltration, or microfiltration. This category also includes reverse osmosis, forward osmosis, and electrodialysis [24,25].
Chemical-based separation techniques involve the transformation of toxic metals into solid precipitates and separation from the liquid phase [23]. These chemical methods for mitigation of heavy metals include chemical precipitation, coagulation–flocculation, and flotation [24]. In chemical precipitation, metal ions are combined with precipitants, Ca(OH)2, NaOH, Na2CO3, Na(HCO3)2, Na2S, and NaHS, to obtain less soluble compounds (hydroxides, carbonates, sulfides), which can be removed from the solution by using physical processes such as sedimentation, flotation, or filtration [26].
Electricity-based separation methods are extremely effective treatments for toxic metal removal based on their recovery in the elemental state by using anodic and cathodic reactions in the electrochemical cell [21]. They include electrochemical methods (electrodeposition, electrocoagulation, electroflotation, electrooxidation, electrodestabilization) and ion exchange [24,25]. Electrochemical separation is an advanced treatment involving the application of a constant voltage to a pair of electrodes immersed in a solution containing metal ions for their partial or complete decomposition [27].
In photocatalytic separation, a photocatalyst accelerates the process by increasing the necessary activation energy and reaction rate without participating in the reaction [28]. Photocatalytic-based separation has gained considerable attention for toxic metal removal and involves the separation of potentially toxic metals by using light and semiconductors (e.g., UV-irradiated TiO2) to remove chemical contaminants [24,29]. The strategies used for photochemical degradation of toxic metals include direct photolysis, oxidation induced through photolysis, oxidation mediated by photosensitive molecules, and photocatalytic oxidation [30].
Interfacial solar steam generation (ISSG) is a promising and sustainable solution that directly utilizes the solar energy to obtain fresh water by locally heating water at the evaporator–air interface [31]. For this purpose, various interfacial solar evaporators, obtained from different photothermal materials, such as metal nanoparticles, carbonaceous materials (supercapacitor, lithium-ion battery), hydrogels, polymers, or semiconductors, were assessed to obtain high performance solar steam generation [31,32]. Using this technique, light-to-heat conversion takes place on the surface of solar evaporators, the energy loss being converted into electricity using a photovoltaic, thermoelectric (TE), piezoelectric, or triboelectric module [32].
Adsorption-based separation is a solid–liquid mass transfer based on the migration of the adsorbate (toxic metal) to the surface of the solid adsorbent and their bonding over the adsorbent surface due to the chemical and physical interactions [21]. This process involves physisorption; when the adsorbate is linked to the adsorbent using van de Waals forces and chemisorption, chemical reactions take place between them. The physisorption process is reversible, weak, and endothermic compared with chemisorption, which is irreversible, selective, and exothermic [16].
The advantages and drawbacks of the processes mentioned above are presented in Table 3.
Adsorption process can be achieved by using different adsorbents, such as carbon-based adsorbents, mineral adsorbents, magnetic adsorbents, metal–organic frameworks adsorbents (MOFs), agricultural waste adsorbents, and others. Each type of adsorbent has its advantages and disadvantages, and their efficacy for heavy metal removal is influenced by various factors such as surface area, pore size and volume, pH, temperature, and initial concentration of the contaminant. Different modification methods can enhance the adsorption capacity and removal performance of the adsorbents. However, some adsorbents have drawbacks such as low stability, poor chemical stability, low mechanical strength, and demand for pretreatments to enhance their adsorption capacity [24,25,29,37,38].
In the adsorption process, metal ions from solvent are attached to the high area surface of solid adsorbent (sorbent) and bond with the functional groups (-COOH, -NH2, -OH, -CONH2, -SH, -OCH3) presented on the adsorbent surface through different processes, such as chemisorption, chelation, complexation, adsorption by physical forces, adsorption on surface and pores, and diffusion through the cell wall and membrane [39].
The adsorption process includes three steps: (i) physical adsorption of ions on adsorbent surface; (ii) precipitation, complexation, and deposition of adsorbate on the adsorbent surface; (iii) pore filling—the adsorbate condensed into the pore of the adsorbent [40].
The adsorbed molecules create a molecular of atomic film on the surface of adsorbent, due to the imbalance and residual forces that continue to attract and retain the adsorbate on the adsorbent. Van der Waals forces and covalent bonds take place between adsorbent and adsorbent material [41].
Adsorption process can be divided into physical adsorption, when the metal ions are adsorbed on the surface of adsorbent using van de Waals forces, hydrogen bonding, dipole–dipole interactions, and chemical adsorption based on the irreversible attachment of adsorbate to adsorbent through chemical bonding or electron transfer [41].

4. Key Factors That Affect the Adsorption Process

The factors that affect the removal of heavy metals through adsorption include the characteristics of the toxic metals, the adsorbent material, the water quality conditions, and the adsorption process. The pH value of the solution affects the speciation and ionization of the metal, as well as the surface charge of the adsorbent. The initial concentration of the metal ions, the adsorbent dose, and the temperature also affect the removal efficiency. Increasing the adsorbent dose and temperature generally increases removal efficiency, while increasing the initial concentration of the metal ions can initially increase adsorption capacity [20,38].
The optimum parameters used to obtain higher removal efficiency and/or adsorption capacity for different types of agricultural waste are presented in Table 4.

5. Surface Modification of Agricultural Wastes

Modification of adsorbent surface represents adjusting of its physical, chemical, or biological characteristics to obtain different forms on the surface, making it suitable for a desired purpose. This process involves the removal of surface impurities to produce physical and chemical changes, such as surface energy, reactivity, surface charge, hydrophobicity, surface area using mechanical, thermal, and chemical methods, or using a combination of methods, such as mechanochemical and thermochemical processes. Mechanical and thermal processes are used to create pores on the surface, while chemical methods are used to improve the characteristics of the surface. In the case of chemical modification, the skeleton of the adsorbent is affected, leading to improvement of the functional groups and properties of the adsorbent’s surface due to chemical interactions between the modifying agent and the surface of the adsorbent material. Chemical modification can be performed by using acid, alkaline, or neutral solutions [55]. Chemical modification involves processes such as acidification, alkalization, esterification, etherification, carbonization, magnetization, and grafting, which are more effective concerning the surface area, functional groups, and adsorption performances compared with physical modification processes [56].
Acid modification is a wet oxidation process that involves the use of organic acids (acetic, carboxylic, formic, and oxalic acid), mineral acids, or other oxidants such as HNO3, HClO, HCl, SO4, and PO4. Acidification of the surface improves the adsorption of positively charged metal ions due to the positive charge of the surface [55].
During this process, acids can dissolve the constituents from agricultural byproducts with a simultaneous increase in O2 content and decrease in the tortuosity of the pores. Also, this treatment can enhance the hydrolysis of cellulose, resulting a more reactive adsorbent material, compared with untreated waste [56].
Huang et al., 2019 [57], observed an increase in the content of polar functional groups (such as hydroxyl, carboxyl, and carbonyl) and the appearance of nitro groups on the surface of the modified adsorbent. This treatment with nitric acid improved the adsorbent’s polarity and negatively-charged properties.
According to Lesoana et al., 2019 [58], acidification of macadamia-activated carbon with sulfuric, nitric, and phosphoric acids increases the surface area of modified adsorbent from 545 to 824 m2 g−1. Acidification of the material surface reduces the mineral content and increases the adsorption capacity of toxic metals from solutions because the competition between cations and active sites decreases. In the case of dilute acids, the amount of C–H, O–H, and C–O functional groups increase compared with concentrated acids, where the hydroxyl and aldehyde functional groups are transformed in carboxyl groups, which are more oxidized [56].
Boeykens et al., 2019 [59], used H3PO4 as modifying agent for avocado seeds and demonstrated an increase in the adsorption capacity of modified adsorbent for Pb2+ and Cr6+. Kim et al., 2020 [60] and Thabede et al., 2020 [61] also used H3PO4 to improve banana peel biochar and black cumin seeds, respectively. They observed an increase in the adsorption capacity for Mn2+ and Fe2+ [60], and Cd2+ [61]. When H2SO4 was used as modifying agent, Wanja et al., 2016 [62]; and Abid et al., 2016 [63]; reported an improvement in the adsorption capacity of avocado seeds and orange peel for Pb2+, Cd2+, Cu2+ [62], and As6+ [63]. Tejada-Tovar et al., 2015 [64], demonstrated a higher adsorption capacity by using citric acid as modifying agent for raw palm bagasse (from 162 to 451 mg g−1). Ding et al., 2014 [65], and Shooto et al., 2019 [66], reported a higher adsorption capacity of peanut hull and black cumin seeds by using mercaptoacetic acid and HCl as modifying agents for the adsorption of Hg2+ [65] and Co2+ [66].
Alkaline modification improves the adsorption capacity of negatively charged species due to the increase in alkali functional groups that adsorb a positive charge on the surface. This process can be performed by treating adsorbents with NaOH, KOH, LiOH, Na2SiO3, Na2CO3, and other oxides [55].
Alkalization reduces the content of lignin and hemicellulose in the adsorbent material and improves its surface area, mechanical properties, and thermal stability through H-bonding, ion exchange, and complexation reactions between the alkaline-treated agricultural material and pollutants [56].
Zheng et al., 2013 [67], obtained an increase in specific area and pore volume with the increase in NaOH concentration, which involves a higher adsorption capacity of contaminants. In the case of Gao et al., 2013 [68], after treating of Enteromorpha prolifera-activated carbon with KOH, the surface area increased up to 3500 m2 g−1 and the total pore volume up to 2872 cm3 g−1, while the adsorption capacity increased from 454 mg g−1 to 2500 mg g−1.
Research studies have demonstrated that KOH alkalization unlocks structural pores from adsorbent materials, obtaining more visible pores compared with unmodified material, which improves the retention capacity for chemical contaminants. Also, alkaline treatments can enhance the release of cations, such as K+ and Ca2+ from the treated adsorbent materials and improve ion exchange capacity and the retention of cationic pollutants [56].
Ali and Saeed, 2015 [65], who investigated the effect of NaOH as a modifying agent for banana peel, demonstrated that this treatment improved the adsorption capacity of the biosorbent. Higher adsorption capacity (897.54 mg g−1) was reported by Shooto et al., 2019 [66], by treating black cumin seeds with NaOH. Most probably, these differences are also due to the concentration in which the modifying agent is added.
Other chemical methods (oxidants, organic and metal ions): Another chemical process used for modification of the surface is the use of oxidizing agents (KMnO4, H2O2), neutral agents (ZnCl2 and NaCl), metal impregnation (ferric chloride, cerium, zirconium, hydroxides, carbonates, chromates, or nitrates), and organic agents (ethanol) [55].
According to Enaime et al., 2020 [69], oxidizing agents are responsible for increasing the number of oxygen functional groups on adsorbent materials. Wang et al., 2015 [70], also used KMnO4 to modify the surface of hickory wood, and observed an increase in the number of oxygen-containing functional groups as well as an increase in surface area.
Ahmed et al., 2021 [71], and Thabede et al., 2020 [61], used oxidizing agents, such as H2O2 and KMnO4, to modify the surface of watermelon seeds biochar and black cumin seeds. The authors demonstrated that the maximum adsorption capacity of 60.87 mg g−1 for Pb2+ [71] and 23.87 mg g−1 for Cd2+ [61] are higher than of the unmodified adsorbent. Contrarily, Tejada-Tovar et al., 2018 [72], who used a neutral agent (CaCl2) to improve the sorption capacity of orange peel, demonstrated that this pretreatment negatively influences the absorbability of the adsorbent, and the adsorption capacity of Cr6+ decreased from 12.29 mg g−1 for unmodified orange peel to 4.950 mg g−1 in treated orange peel.
Esterification: Esters result from the reaction of hydroxyl groups presented in cellulose with carboxyl groups (anhydrides), which improve the hydrophobicity and mechanical strength of agricultural adsorbents, in this way enhancing the adsorption performances of the waste material. The most used compounds for the esterification of an adsorbent surface are represented by succinic anhydride, EDTA (ethylenediaminetetraacetic acid) dianhydride, citric acid anhydride, and maleic anhydride. The -OH bound with cellulose reacts with citric acid-resulting carboxyl groups on the straw surface, which creates complexation reactions with metal ions. In the case of EDTA, the esterification reaction generates amino and carboxyl groups, which bind with metal ions. Carboxyl groups resulting after esterification are involved in the roughness of the absorbent surface, increasing the surface area or enhancing the porosity of the adsorbent. Also, the type of the esterification reagent strongly influences the adsorption capacity of the adsorbent material due to differences in molecular structure. For example, citric acid, which contains more carboxyl groups compared with tartaric acid, can generate more active sites after treatment of adsorbent and implicitly has a higher adsorption capacity. The esterification reaction between citric acid and hydroxyl groups can be enhanced by the addition of a mild catalyst such as NaH2PO2 · H2O [56].
Etherification: Etherification of agricultural waste involves the substitution of -OH groups with other functional groups from modifying agent, resulting in ethers that can bind potentially toxic metal ions. The most used modifying agents are triethylenetetramine, diethylenetriamine, and ethylenediamine, which can generate amino groups for adsorbent material, able to bind various ions from polluted water.
In addition to amino groups, after the etherification process, the adsorbent material also contains carboxyl and thiol functional groups, which can adsorb chemical pollutants. Carboxyl groups enhance the charge transfer to the oxygen to attract metal ions and promote pollutant retention, while thiol groups can bind to the cation pollutants [56].
Magnetization: Modification of agricultural waste using magnetization involves the introduction of transition metals or their oxides into the adsorbent material to obtain a magnetic material, which are easily removed with an external magnet, making them very effective for the retention of chemical pollutants from aqueous solutions. FeCl3, Fe3O4, Fe2O3, and zero-valent iron nanoparticles are the most used compounds for magnetic modification of adsorbent materials. In the case of nanoscale zero-valent iron nanoparticles, Shao et al., 2020 [73], demonstrated that modified adsorbent revealed enhanced removal performance compared with raw adsorbent. Also, they found that a portion of the metal ions was directly adsorbed by the waste material, while the other part was first reduced to zero-valent Cu and Cu2O, and then adsorbed on the material surface due to the Fe-oxide structure. The addition of Fe3O4 to the surface of agricultural waste is the most used magnetization technique, which can be applied for ferrofluids, microwave-assisted, and mechanochemical techniques for toxic metal removal. Fe3O4 from the adsorbent surface can be obtained using coprecipitation, the functional groups OH and Fe–O from modified adsorbent resulting from the interaction between O and Fe from Fe3O4. This magnetization with Fe3O4 can increase surface area, enhance porosity, activate functional groups, and stabilize surface properties. Magnetization can enhance adsorption properties through specific actions, such as modification of lignocellulose using magnetic materials, easy recovery, and reuse of the adsorbent during the removal process [56].
Surfactant modification: Surfactants are compounds that contain various functional groups, such as phosphates, sulphonates, and quaternary ammonium salts, involved in selective adsorption of the pollutants from the solution. Surfactant modification improves the adsorption capacity of modified adsorbent through the improvement of its surface hydrophobic/hydrophilic properties and enrichment of the variety and quantity of functional groups in the structure of the modified material. The most used compounds for surfactant modification are linear alkyl benzene sulphonates, secondary alkane sulphonates, alkyl trimethyl ammonium halides, and quaternary ammonium compounds. Agricultural byproducts treated with cationic surfactants have revealed good applicability for anionic contaminants removal from contaminated wastewater. These surfactants create numerous positive charges on the material surface and improve the retention of negatively charged pollutants [56]. Soldatkina and Zavrichko, 2019 [74], observed that cationic surfactants can reduce the surface area of the adsorbent material due to the blockage of pores and reduction in access to the internal surface area without influencing adsorption performance.
Carbonization: Carbonization is a thermal decomposition process that involves the heating of agricultural waste in a low-oxygen or anaerobic environment and production of carbonaceous residues, such as biochar, biooil, and non-condensable gas products [56,75]. The main purpose of carbonization is the conversion of agricultural byproducts into solid biochar, which is a heterogeneous charcoal material [56,75]. Synthesis of biochar can be performed using various processes, such as traditional pyrolysis carbonization, hydrothermal carbonization, and microwave carbonization, traditional pyrolysis being the most used process for biochar products due to its high efficiency, simple operation, and large output [75]. Compared with raw agricultural byproducts, obtained biochar products are characterized by a higher specific area, improved porosity and an increased number of functional groups capable of retaining a higher number of pollutants [56].
The yield, performance, structural characteristics, and chemical and physical properties of the biochar obtained using pyrolysis carbonization depend on reaction time and, especially, carbonization temperature [56].
Acidic or alkaline compounds can be added to agricultural products, resulting in an effect on the characteristics, structure, and properties of the obtained biochar. Also, post-optimization of biochar is useful for improving its potential value after pyrolysis [56]. For this purpose, Goswami and Phukan, 2017 [76], added H2SO4 during the carbonization process, resulting in -SO3 groups on the biochar surface that increased the adsorption of adsorbent for chemical contaminants through covalent bonding. Also, Yu et al., 2018 [77], treated biochar obtained with HNO3, generating N− and ≡N+ groups, which can interact with negative metal ions.
Adsorption capacities of different agricultural biosorbents, modified by various processes or unmodified are presented in Table 5.

6. Adsorption Capacity and Removal Performances of Different Types of Agricultural Waste

According to Karic et al., 2022 [95], the global levels of various agricultural waste are represented by 80 million tons of husk, 70–140 thousand tons of potato peel, 4.24 million tons of banana peel (40% of global produce of 10.6 million tons), 55 million tons of corn cob, 300 million tons of rice straw, 10 million tons of medicinal plant waste, 23–28 million tons of citrus peel (50–60% of global production of 47 million tons), 25 million tons of coconut waste, and 1436.1 thousand tons of shell waste (20% of global production of 7180.5 thousand tons).
Agricultural waste can be grouped into four groups, namely (i) cereal waste (straw, husks), (ii) fruit waste (pits, peels, shells), (iii) plant waste (cobs, beet, stems), and (4) biosorbents (represented by living and nonliving biomass: algae, microorganisms) [95].
For the mitigation of potentially toxic elements, various types of plants or different parts of plants, such as stems, barks, straws, husks, stalks, roots, leaves, and others, can be utilized, as these byproducts are readily and abundantly available [96].
Husks are a byproduct widely used for removing toxic metals from wastewater sources. Rice husks are the most commonly used agricultural waste for metal mitigation processes [96], but research studies have revealed the use of other husks such as peanut husk [97], sunflower husk [98], corn husk [99], coffee husk [100], or walnut husk [101]. As can be seen from Table 6, husks can be used to remove various toxic metals such as Cr6+ [102], Pb2+, Cd2+, Cu2+, and Cr3+ [100], Ni2+ [103], or Zn2+ [104]. Good absorbability was demonstrated in studies conducted by Sugashini et al., 2015 [103], for Cr6+, with a retention capacity of 52.1 mg g−1; by Olguin et al., 2013 [102], for Cr6+, with a retention capacity of 33.1 mg g−1; and by Sobhanardakani et al., 2013 [105], for Cr3+ and Cu2+, with retention capacities of 22.5 mg g−1 and 30.0 mg g−1, respectively. However, a lower retention capacity of 1.470 mg g−1, 0.174 mg g−1, 0.188 mg g−1, and 0.259 mg g−1 was demonstrated by Guevara-Bernal et al., 2022 [100], for Pb2+, Cd2+, Cr3+, and Cu2+, using coffee husk as the adsorbent material.
Concerning the removal efficiency, the results obtained by Priya et al., 2022 [103]; Radenkovic et al., 2022 [98]; Ismail et al., 2022 [99]; and Dalali and Hagghi, 2015 [101]; revealed removal performances higher than 80% for Pb (98.7%; 81.76%), Cu (90.3%), Cd (80.0%; 90.3%), Cr (90.84%), and Ni (94.0 %).
Straws are lignocellulosic waste materials that result from the cereal industry. The most commonly used adsorbent materials are those obtained from wheat straws and barley straws [96], but research studies have shown improved absorbability for other types of straw, such as rice straw [112], corn straw [113], or rape straw [115]. Isotherm and kinetic models reveal that these byproducts are suitable for various metals, such as Cu2+, Ni2+, Co2+, Cd2+ [111], Pb2+, Zn2+ [112], and Cr6+, Ni2+ [108]. Higher retention capacity was obtained for Cr6+ and Ni2+ at 47.16 mg g−1 and 41.84 mg g−1, respectively [108], and for Cd2+ at 39.22 mg g−1 [109] and 32.737 mg g−1 [115]. Conversely, Pehlivan et al., 2009 [110], and Arshadi et al., 2014 [111], obtained reduced absorbability for Cu2+ (4.64 mg g−1) [110], Cd2+ (1.42 mg g−1), Co2+ (6.58 mg g−1), and Ni2+ (8.25 mg g−1) [111]. As shown in Table 6, there are significant differences in the adsorption capacity of Ni2+ in the studies of Dhir and Kumar, 2010 [108], and Arshadi et al., 2014 [111], but these differences may be caused by the adsorbent used or the modification of the adsorption capacity using pretreatment. With regard to the removal efficiency, Chi et al., 2017 [114], and Wu and Wang, 2016 [116], reported values higher than 98% for Cd2+ and Pb2+ [114] and Ni2+ [116].
Fruit peel is a waste material from fruits, which is abundant and has high potential for mitigating toxic metals due to its high content of carbon-rich organic compounds such as cellulose, hemicellulose, pectin, chlorophyll, and other low molecular weight compounds. Orange peel is the most commonly used adsorbent material from this category, and several researchers have reported the removal of toxic compounds after chemically modifying the peel surface using acid or alkaline treatments [213]. In addition to orange peel, pomegranate peel [117], grapefruit peel [118], banana peel [119], potato peel [123], tangerine peel [124], lemon peel [125], and melon peel [126] have been used as adsorbent materials. Concerning the adsorption capacity, sweet lime peel revealed a higher absorbability for Cr6+ (250 mg g−1) [120], but also melon peel for Pb2+ (191.93 mg g−1) [126] and potato peel for Cu2+ (84.74 mg g−1) [123]. Lower retention capacity was obtained by Yirga et al., 2022 [122], for Cu2+ and Cd2+ adsorption using orange peel, of 2.78 mg g−1, respectively 2.57 mg g−1, but also by Sabanovic et al., 2020 [125], for retention of Cd2+, Co2+, Cr6+, Cu2+, Mn2+, Ni2+, and Pb2+ using lemon peel, and by Mahmood-ul-Hassan et al., 2015 [119], for adsorption of Cd2+ and Cr6+ using banana peel. In the case of removal efficiency, Abdic et al., 2017 [124], demonstrated performances higher than 88% for Cr3+, Cu2+, Mn2+, Co2+, Ni2+, Pb2+, Cd2+, and Zn2+ by using tangerine peel as adsorbent material.
Shells are highly porous materials with a large surface area and a high affinity for water and cell wall components [96]. Research studies have demonstrated the capacity of various materials, such as cashew nutshell, walnut shell [127], almond shell, hazelnut shell [128], Bael fruit shell [131], coconut shell [18], and cocoa shell [133], for the removal of toxic metals. As can be seen from Table 6, higher adsorption capacities were obtained for cashew nutshells of 406.6 mg g−1, 436.7 mg g−1, 455.7 mg g−1, 456.3 mg g−1 for Cu2+, Cd2+, Zn2+, Ni2+, respectively, and for walnut shell, for Cr6+, of 200 mg g−1 [127]. Contrarily, low absorbability was demonstrated by Pehlivan et al., 2009 [128], for almond shell, of 8.08 mg g−1, for Pb2+.
The use of stalks as an adsorbent material is owed to their high content in cellulose, hemicellulose, and lignin [96], which can allow potentially toxic metals to bind to these compounds. Jalali and Aboulghaez, 2013 [134], investigated the feasibility of sunflower stalks as an adsorbent material for Pb2+ and Cd2+. They obtained a high removal efficiency of 97% and 87%, respectively, and a good adsorption capacity of 182.9 mg g−1 and 69.8 mg g−1. The kinetic models demonstrated improved retention capacity for banana stalks [138], cotton stalks [139], and cassava stalks [143]. Lower uptake capacities were obtained for corn stalks [84], sweet sorghum stalks [141], and grape stalks [140]. As concerning mitigation performance, Yang et al., 2022 [136], demonstrated that 28.67% of Cr6+ ions were removed using corn stalk, while Dehgani et al., 2014 [91], revealed that cotton stalk removed between 90–100% of Cr6+, Mn2+, and Zn2+ ions. Also, Prokopov et al., 2019 [145], reported a removal efficiency of higher than 95% for Cr6+ by using the stalks of three species of tobacco.
Various fruit stones have been used as natural adsorbents in many research studies, after removing their central fleshy parts [96], such as apricot stones [146,147], olive stones [148], peach stones [149,150], lime stones [151], plum stones [152], mango stones [154], and cherry stones [155]. These byproducts are rich in elemental carbon (40–45 wt%) [214]. El-Saharty et al., 2018 [147], reported, by using isotherm and kinetic models, adsorption capacities of apricot stones, for Al3+ and Zn2+, of 333.3 mg g−1 and 500 mg g−1, respectively. Good results were revealed by Abbas et al., 2014 [146], for apricot stone, of 111.11 mg g−1 (Pb2+); Yan et al., 2018 [149], for peach stone, of 118.76 mg g−1 (Pb2+); and by Parlayici and Pehlivan, 2017 [152], of 80.65 mg g−1 (Pb2+). Low uptake capacities were obtained in the studies conducted by Amar et al., 2020 [148], of 0.557 mg g−1, 0.3 mg g−1, 0.581 mg g−1, 2.345 mg g−1, for Cu2+, Cd2+, Pb2+, and Cr6+ (olive stone); Pap et al., 2017 [153], for Pb2+, Cd2+, Ni2+, of 9.93 mg g−1, 12.45 mg g−1, 5.63 mg g−1 (raw plum stone); and Olu-Qwolabi et al., 2012 [154], of 1.90 mg g−1 for Pb2+ (mango stone). Regarding the removal efficiencies, Ku et al., 2012 [151], reported percentages of 8.9–30.8% for Pb2+, Cu2+, Cr6+, and As3+ by using limestone; Amar et al., 2020 [147], of 46–94.5% for Cr6+, Cu2+, Cd2+, and Pb2+ and 81.3% for Cr6+ removal using cherry stones [155]. Improved mitigation performance was demonstrated by Khemmari and Benrachedi, 2019 [150], for Cr6+ by using peach stone.
Vegetable waste is used as a low-cost adsorbent material because it is easily available and has no economic use, leading to zero waste discharge to the environment [96]. Kinetic models have demonstrated higher uptake capacities for tomato waste (150 mg g−1 for Pb2+) [160], carrot waste (86.65 mg g−1 and 88.27 mg g−1 for Cr3+, and Cr6+ respectively) [161], and cabbage waste (54.945 mg g−1 for Pb2+) [156]. Lower retention capacities were obtained for onion waste and garlic waste for As3+, Fe2+, Pb2+, Sn2+, Cd2+, and Hg2+ [163]. Yusuff et al., 2019 [162], obtained removal efficiencies higher than 92% for Pb2+ and Cd2+ by using onion waste, while Yargic et al., 2016 [159], reported uptake efficiency higher than 92% for Cu2+, using tomato waste as an adsorbent material.
Seeds are used as an adsorbent material due to their high lignin content, which is a highly branched polymer insoluble in water and capable of binding toxic metal ions [96]. Many researchers have used moringa seeds as adsorbent materials, but good results have also been obtained with avocado seeds, grape seeds, and papaya seeds. As shown in Table 6, kinetic models have demonstrated higher adsorption capacities for papaya seeds, with values of 97.55 mg g−1 and 99.96 mg g−1 for Cu2+ and Pb2+, respectively [169]. Watermelon seed biochar has also shown good results, with an adsorption capacity of 44.32 mg g−1 for Pb2+ [71]. Contrarily, low absorbability was obtained for Mangifera indica seeds (0.365 mg g−1 for As3+), for Schizizium commune seeds (0.360 mg g−1 for As3+) [166], and for Allium cepa seeds, between 1.40 mg g−1 and 1.78 mg g−1, in the case of Cr6+, Cd2+, Zn2+, Cu2+, and Pb2+ [168].
In case of cakes, many studies on using of olive or cotton cake for potentially toxic metal removal are reported in the literature, due to their high cellulose content [96]. Ucar et al., 2015 [176], examined the applicability of rapeseed oil cake for adsorption of Pb2+ and Ni2+, obtaining an adsorption capacity of 129.87 mg g−1 and 133.33 mg g−1, respectively. Contrarily, Mazurek et al., 2021 [177], obtained lower uptake capacity for Cu2+ and Zn2+ by using rapeseed cake. Kinetic models revealed good sorption capacity for black cumin seeds (106.38 mg g−1 for Cu2+) [178], oil palm cake (125.51 mg g−1 for Pb2+) [173], and gingelly oil cake (105.26 mg g−1 for Pb2+) [170]. Lower adsorption capacity was reported by Meneghel et al., 2013 [172], for moringa seed cake (3.191 mg g−1 for Cr6+) and by Fernandez-Gonzalez et al., 2018 [171], for olive cake (3.571 mg g−1 and 5.851 mg g−1, for Mn2+ and Ni2+, respectively). Regarding removal efficiency, Sireesha and Sreedhar, 2022 [174], demonstrated a capacity of removal higher than 85% for Cu2+, Cr6+, and Ni2+, while Thirugnanasambandham and Sivakumar, 2015 [215], obtained a mitigation performance of 88% for Cu2+.
Leaves are used as natural adsorbent materials due to their abundance in nature and their high lignocellulosic compound content [96]. Ge et al. (2020) [182] demonstrated adsorption capacities of 81.03 mg g−1 and 129.87 mg g−1 for Cu2+ and Pb2+, respectively, using kinetic models. However, low adsorption capacities were observed for pine leaf powder (3.27 mg g−1 for As5+) [181], mango leaves (4.08 mg g−1 for Cd2+) [186], and cabbage leaves (6.307 mg g−1 for Pb2+) [183]. The removal performances reported by Alimohammadi et al., 2017 [187]; Nur-E-Alam et al., 2018 [185]; Malik et al., 2015 [184]; and Kamar et al., 2015 [183] were higher than 94% for As3+ and Hg2+ [187], 95.42% for Cr6+ [185], 95.2% for Pb2+ [184], and 95.67% for Pb2+ [183] using eucalyptus leaves, tea leaves, aloe vera leaf powder, and cauliflower leaves biochar, respectively, as natural adsorbents.
Bagasse is a common waste material resulting from the sugarcane industry and is used for adsorption of various heavy metal ions from wastewater [96]. In addition to sugarcane bagasse, grape bagasse [190], agave bagasse [191,192], mango bagasse [194], and palm bagasse [64] are also used for this purpose. Villabona-Ortiz et al., 2022 [195], and Tejada-Tovar et al., 2015 [64], reported improved uptake capacities for the removal of Cr6+ and Pb2+ using oil palm bagasse and palm bagasse, respectively, with values of 111.45 mg g−1 and 162 mg g−1. However, low sorption capacities of 1.35 mg g−1 and 2.0 mg g−1 were obtained by Mohan et al., 2019 [194], and Aloma et al., 2012 [189], for the mitigation of As3+ and Ni2+ using mango bagasse and sugarcane bagasse, respectively.
The use of hulls has been investigated in many research studies, including peanut hulls [196], rice hulls [197], groundnut hulls [198], maize hulls [199], pistachio hulls [200], soybean hulls [201], hazelnut hulls [202], cotton hulls [203], almond hulls [204], and buckwheat hulls [205], due to their high cellulose and hemicellulose contents, which are suitable for binding toxic metal ions. Wang et al., 2013 [205], and Owalude and Tella, 2016 [198], reported good adsorption capacities for Hg2+ and Cr6+ of 243.9 mg g−1 and 90 mg g−1, respectively, using buckwheat hulls and groundnut hulls. In contrast, Osman et al., 2010 [197], reported very low uptake capacities of 1.3367 mg g−1 and 0.137 mg g−1 for Zn2+ and Cd2+, respectively, in the case of rice hulls, as revealed by using isotherm and kinetic models. As for removal performances, the percentages obtained by Ghasemi et al., 2017 [199]; Beidokhti et al., 2019 [200]; Sheibani et al., 2012 [202]; Sheng-Quan et al., 2012 [201]; and Sahranavard et al., 2011 [204] were 50% for Cu2+ (maize hull), >75% for Ni2+ (pistachio hull), 83.5% for Fe3+ (hazelnut hull), 91.99% for Cr6+ (soybean hull), and >94.14% for Cr6+ (almond hull).
Generally, the physical characteristics of the adsorbent are influenced by the chemical composition and physical structure of the precursor materials. For example, the surface area of an adsorbent may increase as the porosity of the precursor material increases. This can occur when using materials such as straw, which typically have high porosity and therefore can result in an adsorbent with a large surface area. In contrast, using more compact materials such as peels may lead to adsorbents with lower surface area. Furthermore, the size and shape of the precursor materials can also influence the physical characteristics of the prepared adsorbents. For instance, husks with irregular shapes and varying sizes may result in an adsorbent with a less uniform surface area than straw, which has a more uniform shape and size. Overall, the choice of precursor material can have a significant impact on the physical characteristics of the prepared adsorbent, including surface area, porosity, and morphology. Therefore, it is essential to carefully consider the precursor materials used when preparing adsorbents for specific applications.
In addition to these categories of agricultural waste, studies have reported the use of other natural materials, such as green tea waste [117], tea waste biochar [206], hemp fibers [207], corn cob, sunflower achene head [119], cocoa pod [154], coffee waste [208], and mango bark [194]. Khalil et al., 2020 [206], reported higher uptake capacities of 198 mg g−1 using tea waste biochar to remove Cr6+ ions, and Heraldy et al., 2018 [160], reported 108 mg g−1 for Pb2+ (apple juice residue). Celebi et al., 2020 [42], reported reduced absorbability between 1.163 and 2.468 mg g−1 for the removal of Pb2+, Cd2+, Ni2+, and Zn2+ using brewed tea waste. Also, Mohan et al., 2019 [194], obtained a lower uptake capacity of 1.25 mg g−1 for As3+ ions using mango bark. The values of adsorption capacities and removal performances presented in Table 6 depend, first of all, on the type of adsorbent used, the modification of the material (e.g., using alkaline, acidic, or other agents), but also on the parameters of the process (initial concentration, temperature, pH value, and adsorbent dosage). However, these parameters can also be influenced by chemical and physical properties of the adsorbent surface, as well as the presence of soluble substances [55].
Optimal characteristics of a suitable adsorbent material must include a large surface area, small pore diameter, high porosity, good mechanical strength, and good chemical and thermal stability, which can enhance the adsorption capacity due to a large exposed surface area and a suitable surface chemistry [55].

7. Recyclability and Regeneration of Agricultural Adsorbents

The regeneration ability of the adsorbent materials prove that the adsorbents are sustainable and economically viable. For this reason, the circularity of adsorbents has received significant attention. If the adsorbed pollutants can be desorbed from the absorbent surface, the process becomes more feasible due to the repeated use of the adsorbent and pollutants [216]. The regeneration capacity of adsorbents is influenced by the pH of the solution, complexation, and rate of oxidation and degradation [217].
The main regeneration techniques are decomposition and desorption methods. Decomposition techniques include oxidation, electrochemistry, ultrasound, and microbiological methods. Desorption methods can be grouped into thermal methods (steam, inert gas, hot water, microwave) and nonthermal methods (surfactants, chemical, supercritical) [218]. The advantages and drawbacks of these techniques are presented in Table 7.
Desorption mechanisms involve the treatment of exhaust adsorbent with eluting agents (solvents or solutions, or a mixture of both), which breaks the bond formed between adsorbent–adsorbate and liberates the adsorbate from the solution. This technique is represented by the following steps: (i) washing the spent absorbent with water; (ii) treating the washed adsorbent with eluting agent in batch/column treatment, stirring the spent adsorbent with eluting agent at different speed, time, and temperature (for batch process) or passing the eluting agent through the spent adsorbent at a predefined flow rate (column process); (iii) washing of treated adsorbent with water to remove the eluting agent [219].
Table 7. Advantages and disadvantages of different regeneration methods.
Table 7. Advantages and disadvantages of different regeneration methods.
ProcessAdvantageDisadvantageReference
Microbiological regenerationEcofriendly, efficient, time effective, cost effectiveVery slow, blocking of the pore, slow regeneration, not applicable for surfactant-modified adsorbents, low efficiency[217,220,221]
Thermal regenerationEfficient, useful for adsorbents
loaded with
heterogeneous mixture of
adsorbates
Expensive, time-consuming, modification of structure adsorption pore, high energy consumes, air pollution problems associated with gases released during the process, loss of carbon surface area[217,220,221]
Oxidative regenerationEfficientExpensive, degradation of pores, time-consuming[217,221]
Microwave regenerationEcofriendly, efficient, time effective, short regeneration time, suitable for multicomponent adsorbents, energy-saving processExpensive, degradation of pores, further secondary treatments are needed[217,221]
Ultrasound regenerationEcofriendly, efficient, time effective, low energy requirement, low carbon lossExpensive, degradation of pores[217,220,221]
Chemical regenerationCost effective, can be coupled easily with other techniques, fast regeneration, high regeneration efficiency, almost zero carbon loss, quick regeneration, possible adsorbate recoveryModification of structure of the adsorbent, oxidant wastage, low solubility of the adsorbate, occasional sludge generation, toxicity issues, require of further treatment[217,220,221]
OzonationEfficientSurface modification of adsorbents, increasing costs due to calcination of adsorbents before use, acidification of adsorption surface which interfere with anionic pollutants[219,220]
Photo-assisted oxidationEcofriendly, fast degradation of organic pollutantsRelease of some harmful byproducts, long regeneration required, decreasing of efficiency with increasing of regeneration cycles[220]
Supercritical fluids regenerationVery fast regeneration times, efficient, less time consumingCost of pressure vessels, degradation of pores of adsorbent material[219,220]
Ultrasonic regenerationHigher desorption efficiency, low energy consumption, simple equipment required, efficient, eco-friendly, less time consumingHighly expensive, degradation of pores of adsorbent[219,220]
Chemical regeneration is one of the most used techniques for the regeneration of adsorbent materials, which uses solvents and chemical reagents for desorption of the adsorbate from the adsorbent [218], or by decomposition of adsorbed species using chemicals that are oxidants under supercritical or subcritical conditions [221]. The regeneration efficiency of chemical regeneration is influenced by the solubility of adsorbates in the solvents, as the solubility of an adsorbent in a solvent is low, the regeneration efficiency is minimal. This process involves three principles: the change in pH of the solution, degradation of adsorbent via oxidation, and complexation (useful for inorganic pollutants) [221]. Menia et al., 2018 [222], demonstrated that HCl 0.5 M is more effective for pea peel compared with NaOH 0.5 M. The authors concluded that the elimination of Zn2+ by desorption is carried out using ion exchange processes. Kalavathy and Miranda, 2010 [223], used different desorbing agents, such as demineralized H2O, HCl, H2SO4, HNO3, Na2CO3, EDTA, NaHCO3, NH4Cl, and CH3COOH. A lower desorption capacity was obtained for demineralized H2O (<1%), followed by Na2CO3, NaHCO3, and NH4Cl (<25%). In the case of CH3COOH, the desorption efficiencies were 45.32% for Cu2+, 47.12% for Ni2+, and 49.56% for Zn2+. Higher recovery efficiency, between 82.11 % and 99.03% for Cu2+, Zn2+, and Ni2+, were obtained after treating the adsorbent with HCl, HNO3, and H2SO4. Complexing agent EDTA revealed higher desorption efficiency (99%), most probably due to the strong metal ion complexing ability to. Contrary to high regeneration capacity, adsorbent treated with EDTA presented the disadvantages of high cost, high stability of metal–EDTA complex, and difficulty in recovering the metal ions from the EDTA solution. In the study of Kyzas, 2012 [224], coffee waste was treated with strong acid and alkaline desorption agents, and the results obtained demonstrated that the desorption percent decreased with the alkalinization of aqueous solution. The decline in adsorption efficiency is a result of both the gradual saturation of the active sites and the deterioration of the material caused by the highly acidic or alkaline pH conditions. Also, Ezeonuegbu et al., 2021 [225], demonstrated that the most effective desorption agent was HNO3, followed by HCl and NaOH. Acidic media produce protonation of the sorbent surface, which allows for the desorption of positively charged metal ions from the adsorbent.
Adsorption capacities of different agricultural byproducts after adsorption-desorption processes are presented in Table 8.

8. Challenges and Future Perspectives on Agricultural Waste Adsorbents

China is currently the country with the highest waste resulting from its agricultural sector in the world. Through the reuse of agricultural byproducts, the air pollution caused by plant incineration can be reduced. Also, the soil and water contamination caused by the residues of heavy metals and pesticides in plants, respectively, by the long-term decay of wastes, can be prevented, in this way improving the soil quality of cultivated land and the ecological environment [233].
This sector faces challenges related to the generation and refinement of surface-modified agricultural waste due to the difficulty of surface-modified procedures. In this manner, Vardhan et al., 2019 [21], recommended the production of techniques fitted to obtain agricultural waste at substantial scale production. An important research step concerning agricultural waste adsorbent is the prediction of the structure of surface-modified adsorbent and enhancing its adsorption capacity. Research on the impact of functional groups on the stability of modified adsorbent, on number of its active sites, and the relation between functional groups and adsorption capacity was considered. They also considered the necessary incorporation of more functional groups, which can retain a higher amount of metal ions, and validation of obtained adsorbent for contaminated industrial wastewaters [21].
Dai et al., 2018 [233], proposed, as future directions, (i) the research and develop of agricultural waste biomass carbonization technology to promote the industrialization of biomass carbon; (ii) development of a green modifier and modification process; and (iii) extending the use of agricultural waste adsorbents in concordance with the engineering problems of pollution scaling.
In addition to the challenges and prospects discussed earlier, several future perspectives must be mentioned. One such perspective involves researching and developing a modification method that can be applied to a blend of two or more agricultural adsorbents. The goal is to create a synergetic effect between the functional groups of the agricultural waste used, resulting in improved adsorption capacities. Another perspective is focused on the development of filter cartridges based on agricultural waste adsorbents. The process would begin with pilot plant trials and then progress to large-scale manufacturing. A third perspective involves researching and developing a green eluent and a green method of regeneration. This would entail finding environmentally friendly solutions for regeneration processes. Another area of research is focused on creating an agricultural adsorbent that can maintain the improved adsorption capacity after many adsorption–desorption cycles. It is also suggested to test the agricultural waste capacity of other harmful substances, such as endocrine disruptor chemicals, e.g., phthalates, plasticizers, flame retardants, parabens, and microplastics. Finally, we propose testing the adsorption capability of agricultural waste adsorbents for solutions that contain a mixture of pollutants. These future perspectives will help in advancing the use of agricultural waste adsorbents and improve their effectiveness in treating contaminated water.
Implementing these perspectives in an efficient and economical manner would require careful planning and the consideration of various factors. Some potential strategies that could be employed are collaboration, optimization, recycling, scaling up, and regulatory support.
Collaboration between academic researchers, industry experts, and policymakers can help to create a more cohesive approach to implementing these perspectives. Working together, stakeholders can leverage their respective areas of expertise to develop efficient and cost-effective methods for implementing these perspectives.
Optimizing the existing processes and technologies used in agricultural waste adsorption can help to reduce costs and improve efficiency. For example, using machine learning algorithms to optimize the design of filter cartridges can help to reduce material waste and improve the performance of the adsorbents.
Developing a closed-loop system for the production and use of agricultural waste adsorbents can help to reduce waste and improve efficiency. For example, spent adsorbents can be recycled and used to produce new filter cartridges or other products, reducing the need for new materials.
Scaling up the production of agricultural waste adsorbents can help to reduce costs by taking advantage of economies of scale. For example, investing in large-scale manufacturing equipment can help to reduce the cost per unit of filter cartridges.
Providing regulatory support for the development and use of agricultural waste adsorbents can help to create a more favorable environment for these technologies. This can include funding for research and development, tax incentives for companies that use these technologies, and streamlined regulatory approval processes.
Implementing these strategies can help to create a more efficient and economical approach to the development and use of agricultural waste adsorbents. By working together and leveraging existing technologies, we can help to create a more sustainable and environmentally friendly approach to water treatment.

9. Conclusions

Agricultural byproducts such as fruit and vegetable peels, hulls, husks, straws, bagasse, stalks, leaves, vegetable and fruit stones, cakes, and vegetable waste are commonly used for chemical contaminants retention due to their low cost, ease of operation, and large surface area. However, the adsorption process is influenced by various factors related to target ions, adsorbent material, wastewater characteristics, and adsorption parameters. Modified adsorbents have remarkable metal uptake capacities compared to unprocessed biosorbents. Agricultural byproducts are suitable for the retention of various heavy metals. There is a need for further research to develop versatile biosorbents for the simultaneous retention of toxic metals from wastewater sources.
This article’s findings indicate that the adsorption capacity of adsorbents is dependent on both the type of agricultural waste utilized and the modifying agent employed to enhance the adsorbent’s surface area. Unprocessed biosorbents contain various functional groups that allow them to bind only a small amount of toxic metal ions in comparison to modified adsorbents, which exhibit significant metal uptake capacities after undergoing chemical pretreatments. Most biosorption studies have demonstrated the suitability of agricultural byproducts for various heavy metals (e.g., Pb2+, Cd2+, Cu2+, Hg2+, Cr3+, Cr6+, As3+, As5+, Mn2+, Sn2+, Co2+, Zn2+, Ni2+, Fe3+). There is a pressing need to research and develop versatile biosorbents based on agricultural byproducts that can effectively remove toxic metals from wastewater sources.

Author Contributions

Conceptualization, E.L.U. and G.M.; methodology, G.M. and L.B.; investigation, E.L.U., A.L.M. and R.M.S.; resources, G.M.; data curation, C.M.P. and C.A.S.; writing—original draft preparation, E.L.U.; writing—review and editing, E.L.U., L.B. and G.M.; visualization, A.L.M.; supervision, E.L.U.; project administration, G.M.; funding acquisition, G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Romanian Ministry of Research, Innovation, and Digitalization, as Intermediate Body for the Competitiveness Operational Program 2014–2020, Projects SMIS number 136213 (METROFOOD-RO) and SMIS number 108234 (IBA SUPORT).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This paper was supported by the Romanian Ministry of Research, Innovation, and Digitalization, as Intermediate Body for the Competitiveness Operational Program 2014–2020, call POC/80/1/2/, Project SMIS number 108234, acronym IBA SUPORT.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nieder, R.; Benbi, D.K.; Reichl, F.X. Role of Potentially Toxic Elements in Soils. In Soil Components and Human Health; Nieder, R., Benbi, D.K., Reichl, F.X., Eds.; Springer: Dordrecht, Germany, 2018; pp. 375–450. [Google Scholar]
  2. Briffa, J.; Sinagra, E.; Blundell, R. Heavy metal pollution in the environment and their toxicological effects on humans. Heliyon 2020, 6, e04691. [Google Scholar] [PubMed]
  3. Khan, S.; Cao, Q.; Zheng, Y.M.; Huang, Y.Z.; Zhu, Y.G. Health risks of heavy metals in contaminated soils and food crops irrigated with wastewater in Beijing. China Environ. Pollut. 2008, 152, 686–692. [Google Scholar] [PubMed]
  4. Rehman, K.; Fatima, F.; Waheed, I.; Akash, M.S.H. Prevalence of exposure of heavy metals and their impact on health consequences. J. Cell. Biochem. 2017, 119, 157–184. [Google Scholar]
  5. Chakraborty, R.; Asthana, A.; Singh, A.K.; Jain, B.; Susan, A.B.H. Adsorption of heavy metal ions by various low-cost adsorbents: A review. Int. Environ. Anal. Chem. 2020, 102, 342–379. [Google Scholar]
  6. Afroze, S.; Sen, T.K. A Review on Heavy Metal Ions and Dye Adsorption from Water by Agricultural Solid Waste Adsorbents. Water Air Soil Poll. 2018, 229, 225. [Google Scholar]
  7. Li, C.; Zhou, K.; Qin, W.; Tian, C.; Qi, M.; Yan, X.; Han, W. A review on heavy metals contamination in soil: Effects, sources and remediation techniques. Soil Sediment Contam Int. J. 2019, 28, 380–394. [Google Scholar]
  8. Rafique, M.; Hajra, S.; Tahir, M.B.; Gillani, S.S.A.; Irshad, M. A review on sources of heavy metals, their toxicity and removal technique using physico-chemical processes from wastewater. Environ. Sci. Pollut. Res. 2022, 29, 16772–16781. [Google Scholar]
  9. Wessling-Resnick, M. Excess iron: Considerations related to development and early growth. Am. J. Clin. Nutr. 2017, 106, 1600–1605. [Google Scholar]
  10. Angelova, I.; Ivanov, I.; Venelinov, T. Origin of Aluminum in the Raw Drinking Water of Sofia City, Bulgaria. Wat. Air Soil Poll. 2020, 231, 455. [Google Scholar]
  11. Rehman, A.U.; Nazir, S.; Irshad, R.; Tahir, K.; Rehman, K.U.; Islam, R.U.; Wahab, Z. Toxicity of heavy metals in plants and animals and their uptake by magnetic iron oxide nanoparticles. J. Mol. Liq. 2021, 321, 114455. [Google Scholar]
  12. Kwikima, M.M.; Mateso, S.; Chebude, Y. Potentials of agricultural wastes as the ultimate alternative adsorbent for cadmium removal from wastewater. A review. Sci. Afr. 2021, 13, e00934. [Google Scholar]
  13. Lawinsider. Available online: https://www.lawinsider.com/dictionary/percent-removal (accessed on 1 December 2022).
  14. Jaber, L.; Ihsanullah, I.; Almanassra, I.W.; Backer, S.N.; Abushawish, A.; Khalil, A.K.A.; Alawadhi, H.; Shanableh, A.; Atieh, M.A. Adsorptive Removal of Lead and Chromate Ions from Water by Using Iron-Doped Granular Activated Carbon Obtained from Coconut Shells. Sustainability 2022, 14, 10877. [Google Scholar]
  15. Mokhatab, S.; Poe, W.A.; Mak, J.Y. Natural Gas Dehydration and Mercaptans Removal. In Handbook of Natural Gas Transmission and Processing, 4th ed.; Gulf Professional Publishing: Oxford, UK, 2019; pp. 307–348. [Google Scholar]
  16. Hussain, A.; Madan, S.; Madan, R. Removal of heavy metals from wastewater by adsorption. In Heavy metals—Their Environmental Impacts and Mitigation; Nazal, M.K., Zhao, H., Eds.; IntechOpen Publishing: London, UK, 2021; pp. 1–24. [Google Scholar]
  17. Abin-Bazaine, A.; Trujillo, A.C.; Olmos-Marquez, M. Adsorption Isotherms: Enlightenment of the Phenomenon of Adsorption. In Wastewater Treatment; Ince, M., Ince, O.K., Eds.; IntechOpen Publishing: London, UK, 2022; pp. 1–15. [Google Scholar]
  18. Tcheka, C.; Abia, D.; Iyedjolbo, B.; Akpomie, K.G.; Harouna, M.; Conradie, J. Biosorption of cadmium ions from aqueous solution onto alkaline-treated coconut shell powder: Kinetics, isotherm, and thermodynamics studies. Biomass Convers. Biorefinery 2022, 1–12. [Google Scholar] [CrossRef]
  19. Chen, X.; Hossain, M.F.; Duan, C.; Lu, J.; Tsang, Y.F.; Islam, M.S.; Zhou, Y. Isotherm models for adsorption of heavy metals from water—A review. Chemosphere 2022, 307, 135545. [Google Scholar] [PubMed]
  20. Bilal, M.; Ihsanullah, I.; Younas, M.; Shah, M.; Ul, H. Recent advances in application of low-cost adsorbents for the removal of heavy metals from water: A critical review. Sep. Purif. Technol. 2022, 278, 119510. [Google Scholar]
  21. Vardhan, K.H.; Kumar, P.S.; Panda, R.C. A review on heavy metal pollution, toxicity and remedial measures: Current trends and future perspectives. J. Mol. Liq. 2019, 290, 111197. [Google Scholar]
  22. Shrestha, R.; Ban, S.; Devkkota, S.; Sharma, S.; Joshi, R.; Tiwari, P.; Kim, H.Y.; Joshi, K. Technological trends in heavy metals removal from industrial wastewater: A review. J. Environ. Chem. Eng. 2021, 9, 105688. [Google Scholar]
  23. Xiang, H.; Min, x.; Tang, C.J.; Sillanpaa, M.; Zhao, F. Recent advances in membrane filtration for heavy metal removal from wastewater: A mini review. JWPE 2022, 49, 103023. [Google Scholar]
  24. Zhao, M.; Xu, Y.; Zhang, C.; Rong, H.; Zeng, G. New trends in removing heavy metals from wastewater. Appl. Microbiol. Biotechnol. 2016, 100, 6509–6518. [Google Scholar]
  25. Gunatilake, S.K. Methods of removing heavy metals from industrial wastewater. J. Multidiscip. Eng. Sci. Technol. 2015, 1, 12–18. [Google Scholar]
  26. Benalia, M.C.; Youcef, L.; Bouaziz, M.G.; Archour, S.; Menasra, H. Removal of Heavy Metals from Industrial Wastewater by Chemical Precipitation: Mechanisms and Sludge Characterization. Arab. J. Sci. Eng. 2021, 47, 5587–5599. [Google Scholar]
  27. Ungureanu, N.; Vladut, V.; Cristea, M.; Cujbescu, D. Wastewater electrooxidation using stainless steel electrodes. E3S Web Conf. 2020, 180, 03015. [Google Scholar] [CrossRef]
  28. Al-Nuaim, M.A.; Alwasiti, A.A.; Shnain, Z.Y. The photocatalytic process in the treatment of polluted water. Chem. Pap. 2023, 77, 677–701. [Google Scholar]
  29. Qasem, N.A.A.; Mohammed, R.H.; Lawal, D.U. Removal of heavy metals ions from wastewater: A comprehensive and critical review. NPJ Clean Water 2021, 4, 36. [Google Scholar]
  30. Xu, Z.; Zhang, Q.; Li, X.; Huang, X. A critical review on chemical analysis of heavy metal complexes in water/wastewater and the mechanism of treatment methods. Chem. Eng. J. 2022, 429, 131688. [Google Scholar]
  31. Zhang, X.; Yan, Y.; Li, N.; Yang, P.; Yang, Y.; Duan, G.; Wang, X. A robust and 3D-printed solar evaporator based on naturally occurring molecules. Sci. Bull. 2023, 68, 203–213. [Google Scholar]
  32. Fan, Z.; Ren, J.; Bai, H.; He, P.; Hao, L.; Liu, N.; Chen, B.; Niu, R.; Gong, J. Shape-controlled fabrication of MnO/C hybrid nanoparticle from waste polyester for solar evaporation and thermoelectricity generation. Chem. Eng. J. 2023, 451, 138534. [Google Scholar]
  33. Carbotecnia. Available online: https://www.carbotecnia.info/learning-center/filtration-methods/what-is-microfiltration/?lang=en (accessed on 11 January 2023).
  34. Barakat, M.A. New trends in removing heavy metals from industrial wastewater. Arab. J. Chem. 2011, 4, 361–377. [Google Scholar]
  35. The, C.Y.; Budiman, P.M.; Shak, K.P.Y.; Wu, T.Y. Recent Advancement of coagulation-flocculation and its application in wastewater treatment. Ind. Eng. Chem. Res. 2016, 55, 4363–4389. [Google Scholar]
  36. Chen, L.; Ren, J.; Gong, J.; Qu, J.; Niu, R. Cost-effective, scalable fabrication of self-floating xerogel foam for simultaneous photothermal water evaporation and thermoelectric power generation. Chem. Eng. J. 2023, 454, 140383. [Google Scholar]
  37. Esmaeili, H.; Foroutan, R. Investigation into ion exchange and adsorption methods for removing heavy metals from aqueous solutions. Int. J. Biol. Pharm. Allied Sci. 2015, 4, 620–629. [Google Scholar]
  38. Joseph, L.; Jun, B.M.; Flora, J.R.V.; Park, C.M. Removal of heavy metals from water sources in the developing world using low-cost materials: A review. Chemosphere 2019, 229, 142–159. [Google Scholar] [PubMed]
  39. Sud, D.; Mahajan, G.; Kaur, M.P. Agricultural waste material as potential adsorbent for sequestering heavy metal ions from aqueous solutions—A review. Bioresour. Technol. 2008, 99, 6017–6027. [Google Scholar] [CrossRef] [PubMed]
  40. Ambaye, T.G.; Vaccari, M.; van Hullebusch, E.D.; Amrane, A.; Rtimi, S. Mechanisms and adsorption capacities of biochar for the removal of organicand inorganic pollutants from industrial wastewater. IJEST 2020, 18, 3273–3294. [Google Scholar]
  41. Rudi, N.N.; Muhamad, M.S.; Chuan, L.T.; Alipal, J.; Omar, S.; Hamidon, N.; Hamid, N.H.A.; Sunar, N.M.; Ali, R.; Harun, H. Evolution of adsorption process for manganese removal in water via agricultural waste adsorbents. Heliyon 2020, 6, e05049. [Google Scholar]
  42. Celebi, H.; Gok, G.; Gok, O. Adsorption capability of brewed tea waste in waters containing toxic lead (II), cadmium (II), nickel (II), and zinc (II) heavy metal ions. Sci. Rep. 2020, 10, 17570. [Google Scholar]
  43. Wan, S.; Ma, Z.; Xue, Y.; Ma, M.; Xu, S.; Qian, L.; Zhang, Q. Sorption of Lead(II), Cadmium(II), and Copper(II) Ions from Aqueous Solutions Using Tea Waste. Ind. Eng. Chem. Res. 2014, 53, 3629–3635. [Google Scholar]
  44. Celebi, H.; Gok, O. Evaluation of Lead Adsorption Kinetics and Isotherms from Aqueous Solution Using Natural Walnut Shell. Int. J. Environ. Res. 2017, 11, 83–90. [Google Scholar]
  45. Ghasemi, S.; Gholami, R.M.; Yazdanian, M. Biosorption of heavy metal from cadmium rich aqueous solutions by tea waste as a low-cost bio-adsorbent. Jundishapur J. Health Sci. 2017, 9, e37301. [Google Scholar] [CrossRef]
  46. Malakahmad, A.; Tan, S.; Yavari, S. Valorization of wasted black tea as a low-cost adsorbent for nickel and zinc removal from aqueous solution. J. Chem. 2016, 2016, 5680983. [Google Scholar]
  47. Futalan, C.M.; Kim, J.; Yee, J.J. Adsorptive treatment via simultaneous removal of copper, lead and zinc from soil washing wastewater using spent cofee grounds. Water Sci. Technol. 2019, 79, 1029–1041. [Google Scholar] [PubMed]
  48. Yang, S.; Wu, Y.; Aierken, A.; Zhang, M.; Fang, P.; Fan, Y.; Ming, Z. Mono/competitive adsorption of Arsenic(III) and Nickel(II) using modified green tea waste. J. Taiwan Inst. Chem. Eng. 2016, 60, 213–221. [Google Scholar]
  49. Mutongo, F.; Kuipa, O.; Kuipa, P.K. Removal of Cr(VI) from aqueous solutions using powder of potato peelings as a low cost sorbent. Bioinorg. Chem. Appl. 2014, 2014, 973153. [Google Scholar]
  50. Sabri, M.U.; Qayyum, A.A.; Akhtar, M.; Munawar, Z. Adsorption kinetics of iron(II) from waste/aqueous solution by using potato peel as carbonaceous material. Int. J. Biosci. 2018, 13, 212–220. [Google Scholar]
  51. Witek-Krowiak, A.; Szafran, R.G.; Modelski, S. Biosorption of heavy metals from aqueous solutions onto peanut shell as a low-cost biosorbent. Desalination 2011, 265, 126–134. [Google Scholar]
  52. Abdelfattah, I.; Ismail, A.A.; Al Sayed, F.; Almedolab, A.; Aboelghait, K. Biosorption of heavy metals ions in real industrial wastewater using peanuthusk as efficient and cost-effective adsorbent. Environ. Nanotechnol. Monit. Manag. 2016, 6, 176–183. [Google Scholar]
  53. Liu, Y.; Sun, X.; Li, B. Adsorption of Hg2+ and Cd2+ by ethylenediamine modified peanut shells. Carbohydr. Polym. 2010, 81, 335–339. [Google Scholar]
  54. Wasewar, K.L.; Atif, M.; Prasad, B.; Mishra, I.M. Batch adsorption of zinc on tea factory waste. Desalination 2009, 244, 66–71. [Google Scholar]
  55. Abegunde, S.M.; Idowu, K.S.; Adejuwon, O.M.; Adeyemi-Adejolu, T. A review on the influence of chemical modification on the performance of adsorbents. Resour. Environ. Sustain. 2020, 1, 100001. [Google Scholar]
  56. Liu, G.; Liao, L.; Dai, Z.; Qi, Q.; Wu, J.; Ma, L.Q.; Tang, C.; Xu, J. Organic adsorbents modified with citric acid and Fe3O4 enhance the removal of Cd and Pb in contaminated solutions. Chem. Eng. J. 2020, 395, 125108. [Google Scholar]
  57. Huang, B.; Liu, G.; Wang, P.; Zhao, X.; Xu, H. Effect of nitric acid modification on characteristics and adsorption properties of lignite. Processes 2019, 7, 167. [Google Scholar] [CrossRef] [Green Version]
  58. Lesoana, M.; Mlaba, R.P.V.; Mtunzi, F.M.; Klink, M.J.; Ejidike, P.; Pakade, V.E. Influence of inorganic acid modification on Cr (VI) adsorption performance and the physicochemical properties of activated carbon. S. Afr. J. Chem. Eng. 2019, 28, 8–18. [Google Scholar] [CrossRef]
  59. Boeykens, S.P.; Redondo, N.; Obeso, R.A.; Caracciolo, N.; Vázquez, C. Chromium and Lead adsorption by avocado seed biomass study through the use of Total Reflection X-Ray Fluorescence analysis. Appl. Radiat. Isot. 2019, 153, 108809. [Google Scholar] [CrossRef] [PubMed]
  60. Kim, H.; Ko, R.-A.; Lee, S.; Chon, K. Removal Efficiencies of Manganese and Iron Using Pristine and Phosphoric Acid Pre-Treated Biochars Made from Banana Peels. Water 2020, 12, 1173. [Google Scholar] [CrossRef] [Green Version]
  61. Thabede, P.M.; Shooto, N.D.; Naidoo, E.B. Adsorption studies of toxic cadmium (II) and chromium (VI) ions from aqueous solution by activated black cumin (Nigella sativa) seeds. J. Environ. Chem. Eng. 2020, 8, 104045. [Google Scholar] [CrossRef]
  62. Wanja, N.E.; Murungi, J.; Hassanali, A. Application of chemically modified avocado seed for removal of Copper (II), Lead(II), and Cadmium(II) ions from aqueous solutions. Int. J. Appl. Sci. Eng. 2016, 6, 1–15. [Google Scholar]
  63. Abid, M.; Niazi, N.K.; Bibi, I.; Farroqi, A.; Ok, Y.S.; Kunhikrishnan, A.; Ali, F.; Ali, S.; Igalavithana, A.D.; Arshad, M. Arsenic (V) biosorption by charred orange peel in aqueous environments. Int. J. Phytoremediation 2016, 18, 442–449. [Google Scholar] [CrossRef]
  64. Tejada-Tovar, C.; Ruiz Paternina, E.; Gallo Mercado, J.; Moscote Bohorquez, J. Evaluation of the biosorption with african palm bagasse for the removal of Pb (II) in solution. Prospect 2015, 13, 59–67. [Google Scholar]
  65. Ding, Z.; Yu, R.; Hu, X.; Chen, Y. Adsorptive Removal of Hg(II) Ions from Aqueous Solutions Using Chemical-Modified Peanut Hull Powder. Pol. J. Environ. Stud. 2014, 23, 1115–1121. [Google Scholar]
  66. Shooto, N.D.; Thabede, P.M.; Naidoo, E.B. Simultaneous adsorptive study of toxic metal ions in quaternary system from aqueous solution using low-cost black cumin seeds (Nigella sativa) adsorbents. S. Afr. J. Chem. Eng. 2019, 30, 15–27. [Google Scholar] [CrossRef]
  67. Zheng, C.; Ling, Z.; Xiaobai, Z.; Zhimin, F.; An, L. Treatment technologies for organic wastewater. Water Treat. 2013, 11, 250–286. [Google Scholar]
  68. Gao, Y.; Yue, Q.; Gao, B.; Sun, Y.; Wang, W.; Li, Q.; Wang, Y. Comparisons of porous, surface chemistry and adsorption properties of carbon derived from Enteromorpha prolifera activated by H4P2O7 and KOH. Chem. Eng. J. 2013, 232, 582–590. [Google Scholar] [CrossRef]
  69. Enaime, G.; Baçaoui, A.; Yaacoubi, A.; Lübken, M. Biochar for wastewater treatment—Conversion technologies and applications. Appl. Sci. 2020, 10, 3492. [Google Scholar] [CrossRef]
  70. Wang, H.; Gao, B.; Wang, S.; Fang, J.; Xue, Y.; Yang, K. Removal of Pb (II), Cu(II), and Cd (II) from aqueous solutions by biochar derived from KMnO4 treated hickory wood. Bioresour. Technol. 2015, 197, 356–362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Ahmed, W.; Mehmood, S.; Núñez-Delgado, A.; Ali, S.; Qaswar, M.; Shakoor, A.; Mahmood, M.; Chen, D.Y. Enhanced adsorption of aqueous Pb(II) by modified biochar produced through pyrolysis of watermelon seeds. Sci. Total Environ. 2021, 784, 147136. [Google Scholar] [CrossRef]
  72. Tejada-Tovar, C.; Herrera-Barros, A.; Villabona-Ortiz, A.; Gonzalez-Delgado, A.; Nunez-Zarur, J. Hexavalent chromium adsorption from aqueous solution using orange peel modified with calcium chloride: Equilibrium and kinetics study. Indian J. Sci. Technol. 2018, 11, 1–10. [Google Scholar] [CrossRef]
  73. Shao, Y.; Gao, Y.; Yue, Q.; Kong, W.; Gao, B.; Wang, W.; Jiang, W. Degradation of chlortetracycline with simultaneous removal of copper (II) from aqueous solution using wheat straw-supported nanoscale zero-valent iron. Chem. Eng. J. 2020, 379, 122384. [Google Scholar] [CrossRef]
  74. Soldatkina, L.; Zavrichko, M. Equilibrium, kinetic, and thermodynamic studies of anionic dyes adsorption on corn stalks modified by cetylpyridinium bromide. Colloids Interfaces 2019, 3, 4. [Google Scholar] [CrossRef] [Green Version]
  75. Zhao, C.; Tang, J.; Si, H.; Wang, B. Environmental effects and carbon sequestration potential of returning agricultural waste to field by carbonization. WJASS 2022, 8, 1–7. [Google Scholar]
  76. Goswami, M.; Phukan, P. Enhanced adsorption of cationic dyes using sulfonic acid modified activated carbon. J. Environ. Chem. Eng. 2017, 5, 3508–3517. [Google Scholar] [CrossRef]
  77. Yu, D.; Wang, L.; Wu, M. Simultaneous removal of dye and heavy metal by banana peels derived hierarchically porous carbons. J. Taiwan Inst. Chem. 2018, 93, 543–553. [Google Scholar] [CrossRef]
  78. Ali, A.; Saeed, K. Decontamination of Cr (VI) and Mn (II) from aqueous media by untreated and chemically treated banana peel: A comparative study. Desalin. Water Treat. 2015, 53, 3586–3591. [Google Scholar] [CrossRef]
  79. Guo, Y.; Zhu, W.; Li, G.; Wang, X.; Zhu, L. Effect of Alkali Treatment of Wheat Straw on Adsorption of Cu(II) under Acidic Condition. J. Chem. 2016, 2016, 6326372. [Google Scholar] [CrossRef] [Green Version]
  80. Ding, Z.; Hu, X.; Zimmerman, A.R.; Gao, B. Sorption and cosorption of lead (II) and methylene blue on chemically modified biomass. Bioresour. Technol. 2014, 167, 569–573. [Google Scholar] [CrossRef]
  81. Pehlivan, E.; Altun, T.; Parlayici, S. Modified barley straw as a potential biosorbent for removal of copper ions from aqueous solutions. Food chem. 2012, 135, 2229–2234. [Google Scholar] [CrossRef]
  82. Kumar, R.; Sharma, R.K. Synthesis and characterization of cellulose based adsorbents for removal of Ni(II), Cu(II) and Pb(II) ions from aqueous solutions. React. Funct. Polym. 2019, 140, 82–92. [Google Scholar] [CrossRef]
  83. Wang, F.; Pan, Y.; Cai, P.; Guo, T.; Xiao, H. Single and binary adsorption of heavy metal ions from aqueous solutions using sugarcane cellulose-based adsorbent. Bioresour. Technol. 2017, 241, 482–490. [Google Scholar] [CrossRef]
  84. Zheng, L.; Dang, Z.; Yi, X.; Zhang, H. Equilibrium and kinetic studies of adsorption of Cd(II) from aqueous solution using modified corn stalk. J. Hazard. Mater. 2010, 176, 650–656. [Google Scholar] [CrossRef]
  85. Wu, Y.; Fan, Y.; Zhang, M.; Ming, Z.; Yang, S.; Arkin, A.; Fang, P. Functionalized agricultural biomass as a low-cost adsorbent: Utilization of rice straw incorporated with amine groups for the adsorption of Cr(VI) and Ni(II) from single and binary systems. Biochem. Eng. J. 2016, 105, 27–35. [Google Scholar] [CrossRef]
  86. Chen, S.; Yue, Q.; Gao, B.; Li, Q.; Xu, X. Removal of Cr(VI) from aqueous solution using modified corn stalks: Characteristic, equilibrium, kinetic and thermodynamic study. Chem. Eng. J. 2011, 168, 909–917. [Google Scholar] [CrossRef]
  87. Khandanlou, R.; Ahmad, M.B.; Fard Masoumi, H.R.; Shameli, K.; Basri, M.; Kalantari, K. Rapid adsorption of copper(II) and Lead(II) by rice straw/Fe3O4 nanocomposite: Optimization, equilibriumisotherms, and adsorption kinetics study. PLoS ONE 2015, 10, e120264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Pehlivan, E.; Tran, H.T.; Ouédraogo, W.K.I.; Schmidt, C.; Zachmann, D.; Bahadir, M. Sugarcane bagasse treated with hydrous ferric oxide as a potential adsorbent for the removal of As(V) from aqueous solutions. Food Chem. 2013, 138, 133–138. [Google Scholar] [CrossRef] [PubMed]
  89. Yue, M.; Zhang, M.; Liu, B.; Xu, X.; Li, X.; Yue, Q.; Ma, C. Characteristics of amine surfactant modified peanut shell and its sorption property for Cr(VI). Chin. J. Chem. Eng. 2013, 21, 1260–1268. [Google Scholar] [CrossRef]
  90. Namasivayam, C.; Sureshkumar, M.V. Removal of chromium(VI) from water and wastewater using surfactant modified coconut coir pith as a biosorbent. Bioresour. Technol. 2008, 99, 2218–2225. [Google Scholar] [CrossRef] [PubMed]
  91. Dehghani, S.; Esteki, M.; Mousavi, S.F.; Mostafazadeh-Fard, B. Removal of Ni(II), Cd(II) and Cr(III) from industrial wastewater by raw and modified cotton stalk. J. Biodivers. Environ. Sci. 2014, 5, 579–582. [Google Scholar]
  92. Xue, Y.; Gao, B.; Yao, Y.; Inyang, M.; Zhang, M.; Zimmerman, A.R.; Ro, K.S. Hydrogen peroxide modification enhances the ability of biochar (hydrochar) produced from hydrothermal carbonization of peanut hull to remove aqueous heavy metals: Batch and column tests. Chem. Eng. J. 2012, 200–202, 673–680. [Google Scholar] [CrossRef]
  93. Bashir, S.; Zhu, J.; Fu, Q.; Hu, H. Comparing the adsorption mechanism of cd by rice straw pristine and KOH-modified biochar. Environ. Sci. Pollut. Res. 2018, 25, 11875–11883. [Google Scholar] [CrossRef]
  94. Fouladi Tajar, A.; Kaghazchi, T.; Soleimani, M. Adsorption of cadmium from aqueous solutions on sulfurized activated carbon prepared from nut shells. J. Hazard. Mater. 2009, 165, 1159–1164. [Google Scholar] [CrossRef]
  95. Karic, N.; Maia, A.S.; Teodorovic, A.; Atanasova, N.; Langergraber, G.; Crini, G.; Ribeiro, A.R.L.; Dolic, M. Bio-waste valorisation: Agricultural wastes as biosorbents for removal of (in)organic pollutants in wastewater treatment. Chem. Eng. J. Adv. 2022, 9, 100239. [Google Scholar] [CrossRef]
  96. Malik, D.S.; Jain, C.K.; Yadav, A.K. Removal of heavy metals from emerging cellulosic low-cost adsorbents: A review. Appl. Water Sci. 2017, 7, 2113–2136. [Google Scholar] [CrossRef] [Green Version]
  97. Cheng, Q.; Huang, Q.; Khan, S.; Liu, Y.; Liao, Z.; Li, G.; Ok, Y.S. Adsorption of Cd by peanut husks and peanut husk biochar from aqueous solutions. Ecol. Eng. 2016, 87, 240–245. [Google Scholar] [CrossRef]
  98. Radenkovic, M.; Momcilovic, M.; Petrovic, J.; Mrakovic, A.; Relic, D.; Popovic, A. Removal of heavy metals from aqueous media by sunflower husk: A comparative study of biosorption efficiency by using ICP-OES and LIBS. J. Serb. Chem. Soc. 2022, 87, 939–952. [Google Scholar] [CrossRef]
  99. Ismail, M.S.; Yahya, M.D.; Anuta, M.; Obayomi, K.S. Facile preparation of amine -functionalized corn husk derived activated carbon for effective removal of selected heavy metals from battery recycling wastewater. Heliyon 2022, 8, e09516. [Google Scholar] [CrossRef] [PubMed]
  100. Guevara-Bernal, D.F.; Ortiz, M.Y.C.; Gutierrez Cifuentes, J.A.; Bastos-Arrieta, J.; Palet, C.; Candela, A.M. Coffee Husk and Lignin Revalorization: Modification with Ag Nanoparticles for Heavy Metals Removal and Antifungal Assays. Water 2022, 14, 1796. [Google Scholar] [CrossRef]
  101. Dalali, N.; Hagghi, A. Removal of cadmium from aqueous solutions by walnut green husk as a low-cost biosorbent. Desalin. Water Treat. 2015, 57, 13782–13794. [Google Scholar] [CrossRef]
  102. Olguin, M.T.; Lopez-Gonzalez, H.; Serrano-Gomez, J. Hexavalent chromium removal from aqueous solutions by Fe-modified peanut husk. Water Air Soil Pollut. 2013, 224, 1654. [Google Scholar] [CrossRef]
  103. Sugashini, S.; Begum, K.M.M.S. Preparation of activated carbon from carbonized rice husk by ozone activation for Cr(VI) removal. Xinxing Tan Cailiao/New Carbon Mater. 2015, 30, 252–261. [Google Scholar] [CrossRef]
  104. Priya, A.K.; Yogeshwaran, V.; Rajendran, S.; Hoang, T.K.A.; Soto-Moscoso, M.; Ghfar, A.A.; Bathula, C. Investigation of mechanism of heavy metals (Cr6+, Pb2+ & Zn2+) adsorption from aqueous medium using rice husk ash: Kinetic and thermodynamic approach. Chemosphere 2022, 286, 131796. [Google Scholar]
  105. Sobhanardakani, S.; Parvizimosaed, H.; Olyaie, E. Heavy metals removal from wastewaters using organic solid waste—Rice husk. Environ. Sci. Pollut. Res. 2013, 20, 5265–5271. [Google Scholar] [CrossRef]
  106. Lawal, O.S.; Ayanda, O.S.; Rabiu, O.O.; Adebowale, K.O. Application of black walnut (Juglan nigra) husk for the removal of lead (II) ion from aqueous solution. Water Sci. Technol. 2017, 75, 2454–2464. [Google Scholar] [CrossRef]
  107. Dang, V.B.H.; Doan, H.D.; Dang-Vu, T.; Lohi, A. Equilibrium and kinetics of biosorption of cadmium (II) and copper (II) ions by wheat straw. Bioresour. Technol. 2009, 100, 211–219. [Google Scholar] [CrossRef] [PubMed]
  108. Dhir, B.; Kumar, R. Adsorption of heavy metals by salvinia biomass and agricultural residues. Int. J. Environ. Res. 2010, 4, 427–432. [Google Scholar]
  109. Farooq, U.; Khan, M.A.; Atharc, M.; Kozinskia, J.A. Effect of modification of environmentally friendly bioadsorbents wheat (Triticum aestivum) on the biosorptive removal of cadmium(II) ions from aqueous solution. Chem. Eng. J. 2011, 171, 400–441. [Google Scholar] [CrossRef]
  110. Pehlivan, E.; Altun, T.; Parlayici, S. Utilization of barley straws as biosorbents for Cu2+ and Pb2+ ions. J. Hazard. Mater. 2009, 164, 982–986. [Google Scholar] [CrossRef]
  111. Arshadi, M.; Amiri, M.J.; Mousavi, S. Kinetic, equilibrium and thermodynamic investigations of Ni(II), Cd(II), Cu(II) and Co(II) adsorption on barley straw ash. Water Resour. Ind. 2014, 6, 1–17. [Google Scholar] [CrossRef] [Green Version]
  112. Sakhiya, A.K.; Vijay, V.K.; Kaushal, P. Efficacy of rice straw derived biochar for removal of Pb+2 and Zn+2 from aqueous: Adsorption, thermodynamic and cost analysis. Bioresour. Technol. Rep. 2022, 17, 100920. [Google Scholar] [CrossRef]
  113. Jia, D.; Li, C. Adsorption of Pb(II) from aqueous solutions using corn straw. Desalin. Water Treat. 2015, 56, 223–231. [Google Scholar] [CrossRef]
  114. Chi, T.; Zuo, J.; Liu, F. Performance and mechanism for cadmium and lead adsorption from water and soil by corn straw biochar. Front. Environ. Sci. Eng. 2017, 11, 15. [Google Scholar] [CrossRef]
  115. Li, B.; Yang, L.; Wang, C.Q.; Zhang, Q.P.; Liu, Q.C.; Li, Y.D.; Xiao, R. Adsorption of Cd(II) from aqueous solutions by rape straw biochar derived from different modification processes. Chemosphere 2017, 175, 332–340. [Google Scholar] [CrossRef]
  116. Wu, Y.; Wang, L. Kinetic and Thermodynamic Studies of the Biosorption of Ni(II) by Modified Rape Straw. Procedia Environ. Sci. 2016, 31, 75–80. [Google Scholar] [CrossRef] [Green Version]
  117. El-Razik, E.M.A.; Abdel-Karim, A.M. Uses of Some Nano-Sized Organic Wastes to Treat Industrial Wastewater. Egypt. J. Chem. 2021, 64, 3413–3421. [Google Scholar]
  118. Hu, Y.; Cheng, H.; Tao, S. The challenges and solutions for cadmium-contaminated rice in China: A critical review. Environ. Int. 2016, 92, 515–532. [Google Scholar] [CrossRef] [PubMed]
  119. Mahmood-ul-Hassan, M.; Surthor, V.; Rafique, E.; Yasin, M. Removal of Cd, Cr, and Pb from aqueous solution by unmodified and modified agricultural wastes. Environ. Monit. Assess. 2015, 187, 19. [Google Scholar] [CrossRef] [PubMed]
  120. Saha, R.; Mukherjee, K.; Saha, I.; Ghosh, A.; Ghosh, S.K.; Saha, B. Removal of hexavalent chromium from water by adsorption on mosambi (Citrus limetta) peel. Res. Chem. Intermed. 2013, 39, 2245–2257. [Google Scholar] [CrossRef]
  121. Lasheen, M.R.; Ammar, N.S.; Ibrahim, H.S. Adsorption/desorption of Cd(II), Cu(II) and Pb(II) using chemically modified orange peel: Equilibrium and kinetic studies. Solid State Sci. 2012, 14, 202–210. [Google Scholar] [CrossRef]
  122. Yirga, A.; Yadav, O.P.; Dey, T. Waste Orange Peel Adsorbent for Heavy Metal Removal from Water. Pollution 2022, 8, 553–566. [Google Scholar]
  123. Guechi, E.K.; Hamdaowio, O. Evaluation of potato peel as a novel adsorbent for the removal of Cu(II) from aqueous solutions: Equilibrium, kinetics and thermodynamic studies. Desalin. Water Treat. 2015, 57, 10677–10688. [Google Scholar] [CrossRef]
  124. Abdic, S.; Memic, M.; Sabanovic, E.; Sulejmanovic, J.; Begic, S. Adsorptive removal of eight heavy metals from aqueous solution by unmodifed and modifed agricultural waste: Tangerine peel. Int. J. Environ. Sci. Technol. 2017, 15, 2511–2518. [Google Scholar] [CrossRef]
  125. Sabanovic, E.; Memic, M.; Sulejmanovic, J.; Selovic, A. Simultaneous adsorption of heavy metals from water by novel lemon-peel based biomaterial. Pol. J. Chem. Technol. 2020, 22, 46–53. [Google Scholar] [CrossRef] [Green Version]
  126. Ahmadi, H.; Hafiz, S.S.; Sharifi, H.; Rene, N.N.; Habibi, S.S.; Hussain, S. Low cost biosorbent (Melon Peel) for effective removal of Cu (II), Cd (II), and Pb (II) ions from aqueous solution. CSCEE 2022, 6, 100242. [Google Scholar] [CrossRef]
  127. Kumar, P.S.; Ramalingam, S.; Sathyaselvabala, V.; Kirupha, S.D.; Murugesa, A.; Sivanesan, S. Removal of cadmium(II) from aqueous solution by agricultural waste cashew nut shell. Korean J. Chem. Eng. 2012, 29, 756–768. [Google Scholar] [CrossRef]
  128. Pehlivan, E.; Altun, T.; Cetin, S.; Bhanger, M.I. Lead sorption by waste biomass of hazelnut and almond shell. J. Hazard. Mater. 2009, 167, 1203–2128. [Google Scholar] [CrossRef] [PubMed]
  129. Chen, S.; Zhong, M.; Zhou, S.; Li, W.; Wang, T.; Li, J. Study on adsorption of Cu2+, Pb2+, Cd2+, and Zn2+ by the KMnO4 modified biochar derived from walnut shell. Int. J. Environ. Sci. Technol. 2022, 20, 1551–1568. [Google Scholar] [CrossRef]
  130. Kumar, P.S.; Ramalingam, S.; Abhinaya, R.V.; Kirupha, S.D.; Murugesan, A.; Sivanesan, S. Adsorption of Metal Ions onto the Chemically Modified Agricultural Waste. Clean—Soil Air Water 2012, 40, 188–197. [Google Scholar] [CrossRef]
  131. Anandkumar, J.; Mandal, B. Removal of Cr(VI) from aqueous solution using Bael fruit (Aegle marmelos correa) shell as an adsorbent. J. Hazard. Mater 2009, 168, 633–640. [Google Scholar] [CrossRef]
  132. Okoye, A.I.; Ejikeme, P.M.; Onukwuli, O.D. Lead removal from wastewater using fluted pumpkin seed hull activated carbon: Adsorption modelling and kinetics. Int. J. Environ. Sci. Technol. 2010, 7, 793–800. [Google Scholar] [CrossRef] [Green Version]
  133. Ercag, E.; Kanmaz, N.; Bugdayci, M.; Hizal, J. Cr(VI) adsorption on binary and ternary composites of raw cocoa shell with magnetic nanoparticle and Prussian blue. Environ. Nanotechnol. Monit. Manag. 2022, 17, 100625. [Google Scholar]
  134. Jalali, M.; Aboulghazi, F. Sunflower stalk, an agricultural waste, as an adsorbent for the removal of lead and cadmium from aqueous solutions. J. Mater. Cycles Waste Manag. 2013, 15, 548–555. [Google Scholar] [CrossRef]
  135. Zheng, L.; Zhu, C.; Dang, Z.; Zhang, H.; Yi, X.; Liu, C. Preparation of cellulose derived from corn stalk and its application for cadmium ion adsorption from aqueous solution. Carbohydr. Polym. 2012, 90, 1008–1015. [Google Scholar] [CrossRef]
  136. Yang, W.; Lei, G.; Quan, S.; Zhang, L.; Wang, B.; Hu, H.; Li, L.; Ma, H.; Yin, C.; Feng, F.; et al. The Removal of Cr(VI) from Aqueous Solutions with Corn Stalk Biochar. Int. J. Environ. Res. Public Health 2022, 19, 14188. [Google Scholar] [CrossRef]
  137. Gong, Z.; Zhou, W.; Qiu, Z. The study of copper (II) removal from aqueous solutions by adsorption using corn stalk material. Adv. Mater. Res. 2012, 610–613, 1950–1953. [Google Scholar] [CrossRef]
  138. Deng, H.; Li, Q.; Huang, M.; Li, A.; Zang, J.; Li, Y.; Li, S.; Kang, C.; Mo, W. Removal of Zn(II), Mn(II) and Cu(II) by adsorption onto banana stalk biochar: Adsorption process and mechanisms. Water Sci. Technol. 2020, 82, 2962–2974. [Google Scholar] [CrossRef]
  139. Gao, L.; Li, Z.; Yi, W.; Li, Y.; Zhang, P.; Zhang, A.; Wang, L. Impacts of pyrolysis temperature on lead adsorption by cotton stalk-derived biochar and related mechanisms. J. Environ. Chem. Eng. 2021, 9, 105602. [Google Scholar] [CrossRef]
  140. Erdem, M.; Duran, H.; Sahin, M.; Ozdemir, I. Kinetics, thermodynamics, and isotherms studies of Cd(II) adsorption onto grape stalk. Desalin. Water Treat. 2015, 54, 3348–3357. [Google Scholar] [CrossRef]
  141. Dong, J.; Hu, J.; Wang, J. Radiation-induced grafting of sweet sorghum stalk for copper(II) removal from aqueous solution. J. Hazard. Mater. 2013, 262, 845–852. [Google Scholar] [CrossRef] [PubMed]
  142. Kirbiyik, C.; Kilic, M.; Cepeliogullar, O.; Putun, A.E. Use of sesame stalk biomass for the removal of Ni(II) and Zn(II) from aqueous solutions. Water Sci. Technol. 2012, 66, 231–238. [Google Scholar] [CrossRef]
  143. Kang, C.; Li, Q.; Deng, H.; Mo, W.; Meng, M.; Huang, S. EDTAD-modified cassava stalks loaded with Fe3O4: Highly efficient removal of Pb2+ and Zn2+ from aqueous solution. Environ. Sci. Pollut. Res. 2021, 28, 6733–6745. [Google Scholar] [CrossRef] [PubMed]
  144. Mingzhen, Z.; Guijan, L.; Ruijia, L.; Arif, M.; Jinzhao, X. Competitive adsorption of copper and zinc on tea stalk biochar prepared at 2 different pyrolysis temperatures: Mechanism and characteristics. SSRN 2022. [Google Scholar] [CrossRef]
  145. Prokopov, T.; Nikolova, M.; Ivanova, T.; Popova, V.; Dimov, M.; Taneva, D. Equilibrium study of Cr (VI) removal from aqueous solution by stalks from three tobacco species (Nicotiana) grown in Bulgaria. J. Environ. Res. Eng. Manag. 2019, 75, 46–54. [Google Scholar] [CrossRef] [Green Version]
  146. Abbas, M.; Kaddour, S.; Trari, M. Kinetic and equilibrium studies of cobalt adsorption on apricot stone activated carbon. J. Ind. Eng. Eng. Chem. 2014, 20, 745–751. [Google Scholar] [CrossRef]
  147. El-Saharty, A.; Mahmoud, S.N.; Manjood, A.H.; Nassar, A.A.H.; Ahmed, A.M. Effect of Apricot Stone Activated Carbon Adsorbent on the Removal of Toxic Heavy Metals Ions from Aqueous Solutions. IJEE 2018, 3, 51–62. [Google Scholar] [CrossRef]
  148. Amar, M.B.; Walha, K.; Salvado, V. Evaluation of Olive Stones for Cd(II), Cu(II), Pb(II) and Cr(VI) Biosorption from Aqueous Solution: Equilibrium and Kinetics. Int. J. Environ. Res. 2020, 14, 193–204. [Google Scholar] [CrossRef]
  149. Yan, J.; Lan, G.; Qiu, H.; Chen, C.; Liu, Y.; Du, G.; Zhang, J. Adsorption of heavy metals and methylene blue from aqueous solution with citric acid modified peach stone. Sep. Sci. Technol. 2018, 53, 1678–1688. [Google Scholar] [CrossRef] [Green Version]
  150. Khemmari, F.; Benrachedi, K. Peach stones valorized to high efficient biosorbent for hexavalent chromium removal from aqueous solution: Adsorption kinetics, equilibrium and thermodynamic studies. Rev. Roum. Chim. 2019, 64, 603–613. [Google Scholar] [CrossRef]
  151. Ku, K.; Seong-Jik, P.; Woo-Seok, S.; Byoung-Hwan, U.; Young-Kee, K. Removal of Synthetic Heavy Metal (Cr6+, Cu2+, As3+, Pb2+) from Water Using Red Mud and Lime Stone. J. Korean Soc. Environ. Eng. 2012, 34, 566–573. [Google Scholar]
  152. Parlayici, S.; Pehlivan, E. Removal of metals by Fe3O4 loaded activated carbon prepared from plum stone (Prunus nigra): Kinetics and modelling study. Powder Technol. 2017, 317, 23–30. [Google Scholar] [CrossRef]
  153. Pap, S.; Knudsen, T.S.; Radonic, J.; Maletic, S.; Igic, S.M.; Sekulic, M.T. Utilization of fruit processing industry waste as green activated carbon for the treatment of heavy metals and chlorophenols contaminated water. J. Clean. Prod. 2017, 162, 958–972. [Google Scholar] [CrossRef] [Green Version]
  154. Olu-Owolabi, B.I.; Opotu, O.U.; Adebowale, K.O.; Ogunsolu, O.; Alujimi, O.O. 2012. Biosorption of Cd2+ and Pb2+ ions onto mango stone and cocoa pod waste: Kinetic and equilibrium studies. Sci. Res. Essays. 2012, 7, 1614–1629. [Google Scholar] [CrossRef]
  155. Kahraman, H.T.; Pehlivan, E. Cr6+ removal using oleaster (Elaeagnus) seed and cherry (Prunus avium) stone biochar. Powder Technol. 2017, 306, 61–67. [Google Scholar] [CrossRef]
  156. Paneru, K.R.; Jha, V.K. Adsorptive removal of Pb(II) ions from aqueous solution by activated carbon prepared from cabbage waste. Nepal J. Environ. Sci. 2020, 8, 1–10. [Google Scholar] [CrossRef]
  157. Hossain, M.A.; Ngo, H.H.; Guo, W.S.; Nguyan, T.V.; Vigneswaran, S. Performance of cabbage and cauliflower wastes for heavy metal removal. Desalin. Water Treat. 2014, 52, 844–860. [Google Scholar] [CrossRef]
  158. Vafakhah, S.; Bahrololoom, M.E.; Saeedikhani, M. Adsorption Kinetics of Cupric Ions on Mixture of Modified Corn Stalk and Modified Tomato Waste. J. Water Resour. Prot. 2016, 8, 1238–1250. [Google Scholar] [CrossRef] [Green Version]
  159. Yargıç, A.Ş.; Yarbay Şahin, R.Z.; Özbay, N.; Önal, E. Assessment of toxic copper(II) biosorption from aqueous solution by chemically-treated tomato waste. J. Clean. Prod. 2015, 88, 152–159. [Google Scholar] [CrossRef]
  160. Heraldy, E.; Lestati, W.W.; Permatasari, D.; Arimurti, D.D. Biosorbent from tomato waste and apple juice residue for lead removal. J. Environ. Chem. Eng. 2018, 6, 1201–1208. [Google Scholar] [CrossRef]
  161. Bhatti, H.N.; Nasir, A.W.; Hanif, M.A. Efficacy of Daucus carota L. waste biomass for the removal of chromium from aqueous solutions. Desalination 2010, 253, 78–87. [Google Scholar] [CrossRef]
  162. Yusuff, A.S.; Owolabi, J.O.; Igbomezie, C.O. Optimization of process parameters for adsorption of heavy metals from aqueous solutions by alumina-onion skin composite. Chem. Eng. Commun. 2019, 208, 14–28. [Google Scholar] [CrossRef]
  163. Negi, R.; Satpathy, G.; Tyagi, Y.K.; Gupta, R.K. Biosorption of heavy metals by utilizing onion and garlic wastes. Int. J. Environ. Pollut. 2012, 49, 179. [Google Scholar] [CrossRef]
  164. Hegazy, I.; Ali, M.E.A.; Zaghlool, E.H.; Elsheikh, R. Heavy metals adsorption from contaminated water using moringa seeds/ olive pomace byproducts. Appl. Water Sci. 2021, 11, 95. [Google Scholar] [CrossRef]
  165. Muthuraman, R.M.; Murugappan, A.; Soundharajan, B. A sustainable material for removal of heavy metals from water: Adsorption of Cd(II), Pb(II), and Cu(II) using kinetic mechanism. Desalin. Water Treat. 2021, 220, 192–198. [Google Scholar] [CrossRef]
  166. Giri, D.D.; Jha, J.M.; Srivastava, N.; Shah, M.; Almalki, A.H.; Alkhanani, M.F.; Pal, D.B. Waste seeds of Mangifera indica, Artocarpus heterophyllus and Schizizium commune as biochar for heavy metal removal from simulated wastewater. Biomass Convers. Biorefinery 2022. [Google Scholar] [CrossRef]
  167. Fristak, V.; Moreno-Jimenez, E.; Fresno, T.; Diaz, E. Effect of Physical and Chemical Activation on Arsenic Sorption Separation by Grape Seeds-Derived Biochar. Separations 2018, 5, 59. [Google Scholar] [CrossRef] [Green Version]
  168. Sheikh, Z.; Amin, M.; Khan, N.; Khan, M.N.; Sami, S.K.; Khan, S.B.; Hafeez, I.; Khan, S.A.; Bakhsh, E.; Cheng, C.K. Potential application of Allium Cepa seeds as a novel biosorbent for efficient biosorption of heavy metals ions from aqueous solution. Chemosphere 2021, 279, 130545. [Google Scholar] [CrossRef] [PubMed]
  169. Garba, Z.N.; Bello, I.; Galadima, A.; Lawal, A.Y. Optimization of adsorption conditions using central composite design for the removal of copper (II) and lead (II) by defatted papaya seed. Karbala Int. J. Mod. Sci. 2016, 2, 20–28. [Google Scholar] [CrossRef] [Green Version]
  170. Nagashanmugam, K.B.; Srinivasan, K. Evaluation of carbon derived from Gingelly oil cake for the removal of lead(II) from aqueous solutions. J. Environ. Sci. Eng. 2010, 52, 349–360. [Google Scholar]
  171. Fernandez-Gonzalez, R.; Martin-Lara, M.A.; Ianez-Rodriguez, I.; Calero, M. Removal of heavy metals from acid mining effluents by hydrolyzed olive cake. Bioresour. Technol. 2018, 268, 169–175. [Google Scholar] [CrossRef]
  172. Meneghel, A.P.; Goncalves, A.C., Jr.; Strey, L.; Rubio, T.; Schwantes, D.; Casarin, J. Biosorption and removal of chromium from water by using moringa seed cake (Moringa oleifera Lam.). Quim. Nova 2013, 36, 104–110. [Google Scholar] [CrossRef]
  173. Yusoff, M.E.M.; Idris, J.; Zainal, N.H.; Ibrahim, M.F.; Abd-Aziz, S. Adsorption of Heavy Metal Ions by Oil Palm Decanter Cake Activated Carbon. Makara J. Sci. 2019, 23, 59–64. [Google Scholar] [CrossRef] [Green Version]
  174. Sireesha, S.; Sreedhar, I. Modified low cost engineered biochar prepared from neem de-oiled cake for heavy metal sorption. Mater. Today Proc. 2022, 72, 34–40. [Google Scholar] [CrossRef]
  175. Vishnuvardhan Reddy, T.; Chauhan, S.; Chakraborty, S. Adsorption isotherm and kinetics analysis of hexavalent chromium and mercury on mustard oil cake. Environ. Eng. Res. 2017, 22, 95–107. [Google Scholar] [CrossRef]
  176. Ucar, S.; Erdem, M.; Tay, T.; Karagoz, S. Removal of lead (II) and nickel (II) ions from aqueous solution using activated carbon prepared from rapeseed oil cake by Na2CO3 activation. Clean Technol. Environ. Policy 2015, 14, 747–756. [Google Scholar] [CrossRef]
  177. Mazurek, K.; Druzynski, S.; Kielkowska, U.; Szlyk, E. New Separation Material Obtained from Waste Rapeseed Cake for Copper(II) and Zinc(II) Removal from the Industrial Wastewater. Materials 2021, 14, 2566. [Google Scholar] [CrossRef] [PubMed]
  178. Coskun, Y.I. Biosorption of Copper by a Natural Byproduct Material: Pressed Black Cumin Cakes. Anal. Lett. 2019, 53, 1247–1265. [Google Scholar] [CrossRef]
  179. Mondal, D.K.; Nandi, B.K.; Purkait, M.K. Removal of mercury (II) from aqueous solution using bamboo leaf powder: Equilibrium, thermodynamic and kinetic studies. J. Environ. Chem. Eng. 2013, 1, 891–898. [Google Scholar] [CrossRef]
  180. Al Rmalli, S.W.; Dahmani, A.A.; Abuein, M.M.; Gleza, A.A. Biosorption of mercury from aqueous solutions by powdered leaves of castor tree (Ricinus communis L.). J. Hazard. Mater. 2008, 152, 955–959. [Google Scholar] [CrossRef]
  181. Shafique, U.; Ijaz, A.; Salman, M.; Zaman, W.; Jamil, N.; Rehman, R.; Javaid, A. Removal of arsenic from water using pine leaves. J. Taiwan Inst. Chem. Eng. 2012, 43, 256–263. [Google Scholar] [CrossRef]
  182. Ge, Q.; Tian, Q.; Moeen, M.; Wang, S. Facile Synthesis of Cauliflower Leaves Biochar at Low Temperature in the Air Atmosphere for Cu(II) and Pb(II) Removal from Water. Materials 2020, 13, 3163. [Google Scholar] [CrossRef]
  183. Kamar, F.H.; Nechifor, A.C.; Ridha, M.J.M.; Mohammed Altaieemi, M.B.; Nechifor, G. Study on Adsorption of Lead Ions from Industrial Wastewater by Dry Cabbage Leaves. Rev. Chim. 2015, 66, 921–925. [Google Scholar]
  184. Malik, R.; Lata, S.; Singhal, S. Removal of heavy metal from waste water by the use of modified aloe vera leaf powder. Int. J. Basic Appl. Sci. 2015, 5, 6–17. [Google Scholar]
  185. Nur-E-Alam, M.; Mia, A.S.M.; Ahmad, F.; Rahman, M.M. Adsorption of chromium (Cr) from tannery wastewater using low-cost spent tea leaves adsorbent. Appl. Water Sci. 2018, 8, 129. [Google Scholar] [CrossRef] [Green Version]
  186. Al Prol, A.E.; El Azeem, M.A.; Amer, A.; El-Metwally, M.E.A.; El-Hamid, H.T.A.; El-Moselhy, K.M. Adsorption of Cadmium (II) Ions from Aqueous Solution onto Mango Leaves. Asian J. Chem. Sci. 2017, 2, 1–11. [Google Scholar] [CrossRef]
  187. Alimohammadi, M.; Saeedi, Z.; Akbarpour, B.; Rasoulzadeh, H.; Yetilmezsoy, K.; Al-Ghouti, M.A.; Khraisheh, M.; McKay, G. Adsorptive Removal of Arsenic and Mercury from Aqueous Solutions by Eucalyptus Leaves. Water Air Soil Pollut. 2017, 228, 429. [Google Scholar] [CrossRef]
  188. Sathish, T.; Vinithkumar, N.V.; Dharani, G.; Kirubagarn, R. Efficacy of mangrove leaf powder for bioremediation of chromium (VI) from aqueous solutions: Kinetic and thermodynamic evaluation. Appl. Water Sci. 2014, 5, 153–160. [Google Scholar] [CrossRef] [Green Version]
  189. Aloma, I.; Martın-Lara, M.A.; Rodrıguez, I.L.; Blazquez, G.; Calero, M. Removal of nickel (II) ions from aqueous solutions by biosorption on sugarcane bagasse. J. Taiwan Inst. Chem. Eng. 2012, 43, 275–281. [Google Scholar] [CrossRef]
  190. Demiral, H.; Gungor, C. Adsorption of copper(II) from aqueous solutions on activated carbon prepared from grape bagasse. J. Clean. Prod. 2016, 124, 103–113. [Google Scholar] [CrossRef]
  191. Cholico-Gonzalez, D.; Lara, N.O.; Macedo, A.M.F.; Salas, J.C. Adsorption Behavior of Pb(II), Cd(II), and Zn(II) onto Agave Bagasse, Characterization, and Mechanism. ACS Omega 2020, 5, 3302–3314. [Google Scholar] [CrossRef]
  192. Velazquez-Jimenez, L.H.; Pavlick, A.; Rangel-Mendez, J.R. Chemical characterization of raw and treated agave bagasse and its potential as adsorbent of metal cations from water. Ind. Crop. Prod. 2013, 43, 200–206. [Google Scholar] [CrossRef]
  193. Ghorbani, F.; Kamari, S.; Zamani, S.; Akbari, S.; Salehi, M. Optimization and modeling of aqueous Cr(VI) adsorption onto activated carbon prepared from sugar beet bagasse agricultural waste by application of response surface methodology. Surf. Interfaces 2020, 18, 100444. [Google Scholar] [CrossRef]
  194. Mohan, D.; Markandeya; Dey, S.; Dwivedi, S.B.; Shukla, S.P. Adsorption of arsenic using low-cost adsorbents: Guava leaf biomass, mango bark and bagasse. Curr. Sci. 2019, 117, 649–661. [Google Scholar] [CrossRef]
  195. Villabona-Ortiz, A.; Tejada-Tovar, C.; Gonzalez-Delgado, A.D. Elimination of Chromium (VI) and Nickel (II) Ions in a Packed Column Using Oil Palm Bagasse and Yam Peels. Water 2020, 14, 1240. [Google Scholar] [CrossRef]
  196. Ali, R.M.; Hamad, H.A.; Hussein, M.M.; Malash, G.F. Potential of using green adsorbent of heavy metal removal from aqueous solutions: Adsorption kinetics, isotherm, thermodynamic, mechanism and economic analysis. Ecol. Eng. 2016, 91, 317–332. [Google Scholar] [CrossRef]
  197. Osman, H.E.; Badwy, R.K.; Ahmad, H.F. Usage of some agricultural by-products in the removal of some heavy metals from industrial wastewater. J. Phytol. 2010, 2, 51–62. [Google Scholar]
  198. Owalude, S.O.; Tella, A.C. Removal of hexavalent chromium from aqueous solutions by adsorption on modified groundnut hull. Beni-Suef Univ. J. Basic Appl. Sci. 2016, 5, 377–388. [Google Scholar] [CrossRef] [Green Version]
  199. Ghasemi, S.M.; Mohseni-Bandpei, A.; Ghaderpoori, M.; Fakri, Y.; Keramati, H.; Taghavi, M.; Moradi, B.; Karimyan, K. Application of modified maize hull for removal of Cu(II) ions from aqueous solutions. Environ. Prot. Eng. 2017, 43, 93–103. [Google Scholar] [CrossRef]
  200. Beidokhti, M.Z.; Naeeni, S.T.O.; Abdi Ghahroudi, M.S. Biosorption of Nickel (II) from Aqueous Solutions onto Pistachio Hull Waste as a Low-Cost Biosorbent. Civ. Eng. J. 2019, 5, 447–457. [Google Scholar] [CrossRef] [Green Version]
  201. Sheng-Quan, Y.; Si-Yuan, G.; Yi-Gang, Y.; Hui, W.; Han-Rui. Removal of the heavy metal ion Cr(VI) by soybean hulls in dyehouse wastewater treatment. Desalination Water Treat. 2012, 42, 197–201. [Google Scholar] [CrossRef]
  202. Sheibani, A.; Shishehbor, M.R.; Alaei, H. Removal of Fe(III) ions from aqueous solution by hazelnut hull as an adsorbent. Int. J. Ind. Chem. 2012, 3, 4. [Google Scholar] [CrossRef] [Green Version]
  203. Yahya, M.D.; Yohanna, I.; Auta, M.; Obayomi, K.S. Remediation of Pb (II) ions from Kagara gold mining effluent using cotton hull adsorbent. Sci. Afr. 2020, 8, e00399. [Google Scholar] [CrossRef]
  204. Sahranavard, M.; Ahmadpour, A.; Doosti, M.R. Biosorption of Hexavalent Chromium Ions from Aqueous Solutions using Almond Green Hull as a Low-Cost Biosorbent. Eur. J. Res. 2011, 58, 392–400. [Google Scholar]
  205. Wang, Z.; Yin, P.; Qu, R.; Chen, H.; Wang, C.; Ren, S. Adsorption kinetics, thermodynamics and isotherm of Hg(II) from aqueous solutions using buckwheat hulls from Jiaodong of China. Food Chem. 2013, 136, 1508–1514. [Google Scholar] [CrossRef]
  206. Khalil, U.; Shakoor, M.B.; Ali, S.; Rizwan, M.; Alyemeni, M.N.; Wijaya, L. Adsorption-reduction performance of tea waste and rice husk biochars for Cr(VI) elimination from wastewater. J. Saudi Chem. Soc. 2020, 24, 799–810. [Google Scholar] [CrossRef]
  207. Tofan, L.; Teodosiu, C.; Paduraru, C.; Wenkert, R. Cobalt (II) removal from aqueous solutions by natural hemp fibers: Batch and fixed-bed column studies. Appl. Surf. Sci. 2013, 285, 33–39. [Google Scholar] [CrossRef]
  208. Hu, C.H.; Kuo, C.Y.; Guan, S.S. Adsorption of heavy metals from aqueous solutions by waste coffee residues: Kinetics, equilibrium, and thermodynamics. Desalin. Water Treat. 2014, 57, 5056–5064. [Google Scholar]
  209. Ghasemi, M.; Naushad, M.; Ghasemi, N.; Khosravi-fard, Y. A novel agricultural waste based adsorbent for the removal of Pb(II) from aqueous solution: Kinetics, equilibrium and thermodynamic studies. J. Ind. Eng. Chem. 2014, 20, 454–461. [Google Scholar] [CrossRef]
  210. Asim, N.; Amin, M.H.; Samsudin, N.A.; Badiei, M.; Razali, H.; Akhtaruzzaman, M.D.; Amin, N.; Sopian, K. Development of effective and sustainable adsorbent biomaterial from an agricultural waste material: Cu(II) removal. Mater. Chem. Phys. 2020, 249, 123128. [Google Scholar] [CrossRef]
  211. Rahaman, M.; Das, A.; Bose, S. Development of copper—Iron bimetallic nanoparticle impregnated activated carbon derived from coconut husk and its efficacy as a novel adsorbent toward the removal of chromium (VI) from aqueous solution. Water 2021, 93, 1417–1427. [Google Scholar] [CrossRef] [PubMed]
  212. Obayomi, K.S.; Bello, J.O.; Nnoruka, J.S.; Adediran, A.A.; Olajide, P.O. Development of low-cost bio-adsorbent from agricultural waste composite for Pb(II) and As(III) sorption from aqueous solution. Cogent Eng. 2019, 6, 1687274. [Google Scholar] [CrossRef]
  213. Abd-Talib, N.; Chuong, C.S.; Mohd-Setapar, S.H.; Asli, U.A.; Pa’ee, K.F.; Len, K.Y.T. Trends in Adsorption Mechanisms of Fruit Peel Adsorbents to Remove Wastewater Pollutants (Cu (II), Cd (II) and Pb (II)). J. Water Environ. Technol. 2020, 18, 290–313. [Google Scholar] [CrossRef]
  214. Saleem, J.; Shahid, U.B.; Hijab, M.; Mackey, H.; McKay, G. Production and applications of activated carbons as adsorbents from olive stones. Biomass Convers. Biorefinery 2019, 9, 775–802. [Google Scholar] [CrossRef] [Green Version]
  215. Thirugnanasambandham, K.; Sivakumar, V. An eco-friendly approach for copper (II) ion adsorption onto cotton seed cake and its characterization: Simulation and validation. J. Taiwan Inst. Chem. Eng. 2015, 50, 198–204. [Google Scholar] [CrossRef]
  216. Ahsanul Haque, A.N.; Sultana, N.; Muhammad Sayem, A.S.; Smriti, S.A. Sustainable Adsorbents from Plant-Derived Agricultural Wastes for Anionic Dye Removal: A Review. Sustainability 2022, 14, 17. [Google Scholar]
  217. El-Messaoudi, N.; El Khomri, M.; El Mouden, A.; Bouich, A.; Jada, A.; Lacherai, A.; Iqbal, H.M.N.; Mulla, S.I.; Kumar, V.; Americo-Pinheiro, J.H.P. Regeneration and reusability of non-conventional low-cost adsorbents to remove dyes from wastewaters in multiple consecutive adsorption–desorption cycles: A review. Biomass Convers. Biorefinery 2022, 5, 29. [Google Scholar]
  218. Alsawy, T.; Rashad, E.; El-Qelish, M.; Mohammed, R.H. A comprehensive review on the chemical regeneration of biochar adsorbent for sustainable wastewater treatment. NPJ Clean Water 2022, 5, 29. [Google Scholar]
  219. Patel, H. Review on solvent desorption study from exhausted adsorbent. J. Saudi Chem. Soc. 2021, 25, 101302. [Google Scholar]
  220. Omorogie, M.O.; Babalola, J.O.; Unuabonah, E.I. Regeneration strategies for spent solid matrices used in adsorption of organic pollutants from surface water: A critical review. Desalin. Water Treat. 2016, 57, 518–544. [Google Scholar] [CrossRef]
  221. Othmani, A.; Magdouli, S.; Kumar, P.S.; Kapoor, A.; Chellam, P.V.; Gokkus, O. Agricultural waste materials for adsorptive removal of phenols, chromium (VI) and cadmium (II) from wastewater: A review. Environ. Res. 2022, 204, 111916. [Google Scholar] [CrossRef] [PubMed]
  222. Menia, S.; Abbaci, A.; Azzouz, N. Regeneration of Peel of Peas (Pisum sativum) After Zinc Adsorption. Green Energy Technol. 2018, 133, 136. [Google Scholar]
  223. Kalavathy, H.M.; Miranda, L.R. Moringa oleifera—A solid phase extractant for the removal of copper, nickel and zinc from aqueous solutions. Chem. Eng. J. 2010, 158, 188–199. [Google Scholar] [CrossRef]
  224. Kyzas, G.Z. Commercial coffee wastes as materials for adsorption of heavy metals from aqueous solutions. Materials 2012, 5, 1826–1840. [Google Scholar] [CrossRef] [Green Version]
  225. Ezeonuegbu, B.A.; Machido, D.A.; Whong, C.M.Z.; Japhet, W.S.; Alexiou, A.; Elazab, S.T.; Qusty, N.; Yaro, C.A.; Batiha, G. Agricultural waste of sugarcane bagasse as efficient adsorbent for lead and nickel removal from untreated wastewater: Biosorption, equilibrium isotherms, kinetics and desorption studies. Biotechnol. Rep. 2021, 30, e00614. [Google Scholar] [CrossRef]
  226. Wen, X.; Sun, N.; Yan, C.; Zhou, S.; Pang, T. Rapid removal of Cr(VI) ions by densely grafted corn stalk fibers: High adsorption capacity and excellent recyclable property. J. Taiwan Inst. Chem. Eng. 2018, 89, 95–104. [Google Scholar] [CrossRef]
  227. Feng, N.; Guo, X.; Liang, S.; Zhu, Y.; Liu, J. Biosorption of heavy metals from aqueous solutions by chemically modified orange peel. J. Hazard. Mater. 2011, 185, 49–54. [Google Scholar] [CrossRef] [PubMed]
  228. Prajapati, A.K.; Verma, P.; Singh, S.; Mondal, M.K. Adsorption-Desorption Surface Bindings, Kinetics, and Mass Transfer Behavior of Thermally and Chemically Treated Great Millet Husk towards Cr(VI) Removal from Synthetic Wastewater. Adsorp. Sci. Technol. 2022, 2022, 3956977. [Google Scholar] [CrossRef]
  229. Iqbal, M.; Saeed, A.; Edyvean, R.G.J. Bioremoval of antimony(III) from contaminated water using several plant wastes: Optimization of batch and dynamic flow conditions for sorption by green bean husk (Vigna radiata). Chem. Eng. J. 2013, 225, 192–201. [Google Scholar] [CrossRef]
  230. Poonam; Bhart, S.K.; Kumar, N. Kinetic study of lead (Pb2+) removal from battery manufacturing wastewater using bagasse biochar as biosorbent. Appl. Water Sci. 2018, 8, 119. [Google Scholar] [CrossRef] [Green Version]
  231. Chowdhury, S.; Saha, P.D. Batch and continuous (fixed-bed column) biosorption of Cu(II) by Tamarindus indica fruit shell. Korean, J. Chem. Eng. 2012, 30, 369–378. [Google Scholar]
  232. Shanmugaprakash, M.; Sivakumar, V. Batch and fixed-bed column studies for biosorption of Zn(II) ions onto pongamia oil cake (Pongamia pinnata) from biodiesel oil extraction. J. Environ. Manag. 2015, 164, 161–170. [Google Scholar] [CrossRef]
  233. Dai, Y.; Sun, Q.; Wang, W.; Lu, L.; Liu, M.; Li, J.; Yang, S.; Sun, Y.; Zhang, K.; Xu, J.; et al. Utilizations of agricultural waste as adsorbent for the removal of contaminants: A review. Chemosphere 2018, 211, 235–253. [Google Scholar]
Figure 1. Overview of heavy metal removal techniques (adapted from Shrestha et al., 2021 [22]).
Figure 1. Overview of heavy metal removal techniques (adapted from Shrestha et al., 2021 [22]).
Sustainability 15 05999 g001
Table 1. Health effects and permissible limits of potentially toxic metals.
Table 1. Health effects and permissible limits of potentially toxic metals.
Toxic MetalHealth HazardPermissible LimitReference
Lead (Pb)Hypertension, anorexia, renal impairment, senile dementia, reduced fertility, cognitive impairment, abdominal painWHO—0.05 mg/L
USEPA—0.005 mg/L
[2,5,6,7,8]
Cadmium (Cd)Bone disease, lung and prostate cancer, pulmonary fibrosis, Itai–Itai disease, hypertension, renal toxicity, testicular atrophyWHO—0.003 mg/L
USEPA—0.005 mg/L
[2,5,6,7,8]
Mercury (Hg)Ataxia, dermatitis, kidney damage, Minamata disease, miscarriages, DNA damage, pulmonary edema, dementia, gingivitisWHO—0.001 mg/L
USEPA—0.002 mg/L
UE—0.001 mg/L
[2,5,7,8]
Arsenic (As)Brain damage, liver tumor, melanosis, lung irritation, infertility and miscarriage, hemolysis, hepatomegaly, conjunctivitisWHO—0.01 mg/L
USEPA—0.05 mg/L
[2,5,6,7,8]
Chromium (Cr)Chronic bronchitis, vomiting, lung cancer, pneumonia, liver damage, nausea and vomiting, renal failure, reproductive toxicityWHO—0.05 mg/L
USEPA—0.1 mg/L
[2,5,6,7,8]
Cobalt (Co)Skin problems, congestion, edema, allergic dermatitis, respiratory problems, pneumonia and fibrosis, cardiac problems, asthma, nausea and vomiting, liver disordersWHO—0.1 mg/L[2,5]
Copper (Cu)Anemia, metabolic disorders, metal fiver, Wilson disease, hepatic and kidney disease, reproductive and developmental toxicity, diabetes, dizzinessWHO—2.5 mg/L
USEPA—1.3 mg/L
[2,5,6,7,8]
Manganese (Mn)Hypertension, pneumonia, neurological problems (dullness, lethargy, weakness), mimicry of Parkinson’s disease, infertility problems, pneumoniaWHO—0.5 mg/L[2,5]
Nickel (Ni)Lung and nasal cancer, dermatitis, nausea and vomiting, kidney diseases, nausea and vomiting, dizziness, heart disorders, asthmaWHO—2.0 mg/L[2,5,7,8]
Iron (Fe)Neurodegenerative diseases (Alzheimer’s disease, Parkinson’s disease), effects on cell proliferation and differentiation, affects cognition and memoryWHO—0.3 mg/L[5,9]
Zinc (Zn)Ataxia, depression, gastrointestinal problems, respiratory disorders, anemia, impaired immune function, metal fume fever, prostate cancerWHO—5.0 mg/L[2,5,7,8]
Aluminum (Al)Dementia, lung damage, pulmonary fibrosis, colitis, neurological diseases (Alzheimer’s, sclerosis, Parkinson’s disease), kidney issuesWHO—0.2 mg/L
USEPA—0.2 mg/L
[2,10]
WHO = World Health Organization; USEPA = United States Environmental Protection Agency.
Table 2. Equation of the most important isotherm and kinetic models used in adsorption studies.
Table 2. Equation of the most important isotherm and kinetic models used in adsorption studies.
Isotherm/Kinetic ModelEquationReference
Langmuir isotherm q e = q m K L C e 1 + K L C e [18]
Freundlich isotherm q e = K F C e 1 / n
Redlich–Peterson isotherm q e = K R C e 1 + α R C e g
Sips isotherm q e = K S C e β S 1 + α S C e β S
PFO model q t = q e ( 1 e k 1 t )
PSO model q t = k 2 q e 2 t 1 + k 2 q e t
Henry’s linear model q e = K p C e [19]
Temkin isotherm q e = R T b ln ( A C e )
Toth isotherm q e = K T C e ( α T + C e z ) 1 / z
Table 3. Advantages and drawbacks of various removal processes.
Table 3. Advantages and drawbacks of various removal processes.
Removal TechniqueAdvantagesDrawbacksReference
UltrafiltrationCost effective, low-pressure demand, great performanceSecondary pollutants, requirement of posttreatments[23]
NanofiltrationEasy to operate, high removal performances, reliability, suitable for industrial practiceHigh energy consumption, lower water permeability[24,25]
MicrofiltrationLow operation pressure, cost effective, easy to operate, low energy demandLow removal performances[29,33]
Reverse osmosisHigh removal performancesMembrane fouling, membrane degradation, high power consumption[24,29]
Forward osmosisEnergy-saving because it does not need hydraulic pressure, environmentally friendly, easy cleaning, low foulingInternal and external concentration polarization, draw solution reconcentration,[29]
ElectrodialysisHigh water recovery, can operate in a wide range of pH, no phase change, no reaction productsMembrane fouling, high operation cost, high cost of membranes, need of electric potential, periodic maintenance[24,29]
Chemical precipitationSimple operation, low cost of precipitant, ease of automatic control, inexpensive and simple, high removal performances, less dissolve solidsRequires a large quantity of precipitates, sludge generation, can produce hydrogen sulfide gas, extra cost for sludge disposal[25,29,34]
Coagulation–flocculationSuitable for broader pH range, simplicity in operation, low energy consumption, high versatilityProduction of sludge, transfer of toxic compound into solid phase, toxicity, hazardousness of inorganic coagulants, results in a high quantity of sludge, selective technique for metals, inefficient for some emerging chemical contaminants, the flocculants are nonbiodegradable[25,29,35]
FlotationLow-energy demand, rapid and compact process, reduced volumes of sludge, moderate costDemand of selective treatments, requirement of efficient and nontoxic surfactants[29]
Electro-depositionSelective method, cost-effective, high capacity to recovery of valuable materialsHigh sensitivity, side reactions affect the removal capacity[29]
ElectrocoagulationSludge resulted is stable, easily removable and nontoxic, easy to operateHigh energy consumption, resulting of harmful byproducts[29]
ElectroflotationShort process time, resulting of stable sludgeDifficulty in controlling system pH[29]
ElectrooxidationOxidizes high toxic compounds, does not result in secondary pollutantsCorrosion of electrodes[29]
Ion exchangeHigh removal efficiency, low cost, large applicability in a wide range of treatments, high selectivity, rapid kinetics, high treatment capacity, recovery of heavy metals, fast kineticsLow longevity, high cost of reduction phase, resins pollution, residual sludge production, demand of resins[25,33]
PhotocatalysisHigh efficiency, rapid destruction, less harmful byproductsHigh cost, long duration time, limited applications[24,34]
Solar steam generationHigh efficiency, low cost, no consumption of additional energy, low heat loss, improves energy utilization of elaborated infrastructuresDifficulty of precise customization of conical arrays, extended processing time[31,32,36]
AdsorptionEasy to operate and use, simplicity, low cost, high effect on contaminants removal, lack of sludge, recycling of adsorbed materials, usage of particles with different sizes, availability of adsorbent materialsLow selectivity[34,37]
Table 4. The optimum parameters of adsorption with different agricultural wastes.
Table 4. The optimum parameters of adsorption with different agricultural wastes.
AdsorbentMetal IonOptimum ParametersR.E. *
(%)
qe*
(mg g−1)
Reference
M (g)pHC0 (mg/L)t (min)T (°C)
Brewed tea wastePb2+, Cd2+, Ni2+, Zn2+1.0; 1.0; 2.0; 2.04.0; 4.0; 5.0; 5.01002; 5; 30; 1020 °C97.97; 84.74; 82; 761.197; 1.457; 1.163; 2.468[42]
Tea wastePb2+, Cd2+, Cu2+0.15.0100; 206025 °C-33.49; 16.87; 21.02[43]
Walnut shellPb2+1.04.0100220 °C909.912[44]
Tea wasteCd2+105.059020 °C99.51.76[45]
Waste black teaZn2+205.010025020 °C80166.7[46]
Coffee groundsZn2+, Cu2+, Pb2+2.55.010030055 °C84.55; 68.73; 57.235.25; 24.28; 48.13[47]
Modified green tea wasteNi2+, As3+0.33.051033 °C-0.3116; 0.4212[48]
Potato peelCr6+4.02.52048-1003.28[49]
Potato peels charcoalFe2+0.8-5020-90.4121.95[50]
Peanut shellsCr3+, Cu2+105.0-6020 °C-27.86;
25.39
[51]
Peanut huskPb2+, Cd2+, Co2+, Mn2+, Ni2+5.06.0100180; 240 (Mn)-99; 62; 30; 45; 3827.03; 11.36; 6.10; 14.29; 56.82[52]
Peanut shellsCd2+, Hg2+0.231030-37 (Hg)6.0; 1.90[53]
Peanut shells100 (Hg)14.17;
30.72
Tea wasteZn2+0.24.22530--8.90[54]
* R.E.—removal efficiency; qe—adsorption capacity.
Table 5. The absorption capacities of unmodified and modified adsorbents.
Table 5. The absorption capacities of unmodified and modified adsorbents.
AdsorbentMetal IonsModifying AgentAdsorption Capacity (mg g−1)Reference
Unmodified FormModified Form
Acidification
Avocado seedsPb2+, Cr6+H3PO418.9; 3.3926.6; 5.10[59]
Banana peel biocharMn2+, Fe2+H3PO40.796; 27.3552.319; 29.550[60]
Black cumin seedsCd2+H3PO418.1619.48[61]
Avocado seedsPb2+, Cd2+, Cu2+H2SO421.48; 11.78; 46.02271.02; 237.14; 102.57[62]
Orange peelAs6+H2SO486.9598.03[63]
Peanut hullHg2+Mercaptoacetic acid30.383.3[65]
Black cumin seedsCo2+HCl28.0578.58[66]
Banana peelMn2+HCl2.8644.736[78]
Alkalization
Black cumin seedsCo2+NaOH28.0587.54[66]
Banana peelMn2+NaOH2.8643.601[78]
Wheat strawCu2+NaOH10.1110.59[79]
Peanut hullsPb2+NaOH19.749.6[80]
Oxidizing agents
Watermelon seeds biocharPb2+H2O244.3260.87[71]
Orange peelCr6+CaCl212.294.950[72]
Black cumin seedsCd2+KMnO418.1623.87[61]
Esterification
Raw palm bagassePb2+Citric acid162451[64]
Barley strawCu2+Citric acid4.6431.71[81]
Rice huskCu2+, Pb2+, Ni2+N-isopropylacrylamide and citric acid86.81; 79.75; 69.9884.67; 118.3; 79.87[82]
Etherification
Sugarcane bagassePb2+, Cu2+, Zn2+Ethylenediamine, CS2-558.9; 446.2; 363.3[83]
Corn stalkCd2+Acrylonitrile3.3912.73[84]
Rice strawCr6+, Ni2+Triethylamine-15.82; 3.95[85]
Corn stalkCr6+Diethylenetriamine-200.0[86]
Magnetization
Sugarcane bagassePb2+, Cd2+Citric acid and Fe2O3-33; 117[56]
Rice strawCu2+, Pb2+Fe3O4-16.31; 19.45[87]
Sugarcane bagasseAs5+Fe oxides-22.1[88]
Wheat strawCu2+Nanoscale zero-valent 376.4[73]
Surfactant modification
Peanut huskCr6+Epichlorohydrin-138.34[89]
Coconut coir pithCr6+Hexadecyltrimethylammonium bromide-76.3[90]
Cotton stalkNi2+, Cd2+, Cr3+Sodium dodecyl sulphate31.91; 19.58; 38.2938.23; 24.93; 43.33[91]
Carbonization
Peanut hull biocharPb2+H2O2-22.82[92]
Rice straw biocharCd2+H3PO412.1741.9[93]
Nut shell biocharCd2+SO2104.17142.86[94]
Table 6. Removal efficiency and adsorption capacity of some unmodified agricultural biosorbents.
Table 6. Removal efficiency and adsorption capacity of some unmodified agricultural biosorbents.
Agricultural WasteMetal IonsRemoval Efficiency (%)Adsorption Capacity (mg g−1)Reference
Husks
Peanut huskCr6+-33.1[102]
Rice huskCr3+, Cu2+-22.5; 30.0[105]
Rice huskCr6+-52.1[103]
Rice husk ashCr6+, Pb2+, Zn2+87.12; 88.63; 99.28-[104]
Sunflower huskPb2+, Cu2+, Cd2+, Ni2+98.7; 90.3; 80.0; 94.0-[98]
Corn huskPb2+, Cu2+, Ni2+99.66; 96.01; 92.36-[99]
Coffee huskPb2+, Cd2+, Cr3+, Cu2+-1.470; 0.174; 0.188; 0.259[100]
Walnut huskCd2+96.11-[101]
Walnut seed huskPb2+-4.0[106]
Straw
Wheat strawCd2+, Cu2+-14.56; 11.43[107]
Wheat strawCr6+, Ni2+-47.16; 41.84[108]
Wheat strawCd2+-39.22[109]
Barley strawCu2+, Pb2+-4.64; 23.2[110]
Barley strawCu2+, Ni2+, Co2+, Cd2+-17.8; 8.25; 6.58; 1.42[111]
Rice strawPb2+, Zn2+-17.93; 25.73[112]
Corn strawPb2+-15.0236[113]
Corn strawCd2+, Pb2+99.24; 98.62-[114]
Rape strawCd2+-32.737[115]
Rape strawNi2+99.7-[116]
Peel
Pomegranate peelCr6+98.95-[117]
Grapefruit peelCd2+-42.09[118]
Banana peelCd2+, Cr6+, Pb2+-3.66; 6.85; 20.90[119]
Sweet lime peelCr6+-250[120]
Orange peelCu2+, Pb2+-15.27; 73.53[121]
Orange peelCu2+, Cd2+96.9; 98.12.78; 2.57[122]
Potato peelCu2+-84.74[123]
Tangerine peelCr3+, Cu2+, Mn2+, Co2+, Ni2+, Pb2+, Cd2+, Zn2+88.92; 97.04; 92.48; 94.70; 93.50; 93.0; 97.90; 96.80-[124]
Lemon peelCd2+, Co2+, Cr6+, Cu2+, Mn2+, Ni2+, Pb2+-7.34; 5.63; 7.56; 7.17; 5.17; 5.73; 8.17[125]
Melon peelCu2+, Cd2+, Pb2+-77.76; 76.16; 191.93[126]
Shells
Cashew nut shellCd2+-22.11[127]
Almond shellPb2+-8.08;[128]
Hazelnut shell28.18
Walnut shellCu2+, Pb2+, Cd2+, Zn2+-30.18; 70.37; 44.94; 58.96[129]
Walnut shellCr6+-200[130]
Cashew nut shellCu2+, Cd2+, Zn2+, Ni2+-406.6; 436.7; 455.7; 456.3
Bael fruit shellCr6+-17.27[131]
Fluted pumpkin seed shellPb2+-14.286[132]
Coconut shellCd2+-14.22[18]
Cocoa shellCr6+-23.36[133]
Stalk
Sunflower stalkPb2+, Cd2+97; 87182.90; 69.80[134]
Corn stalkCd2+-12.73[84]
Corn stalkCd2+-21.37[135]
Corn stalkCr6+28.67-[136]
Corn stalkCu2+-54.05[137]
Banana stalkCu2+, Mn2+, Zn2+-134.88; 109.10; 108.10[138]
Cotton stalkPb2+-146.78[139]
Cotton stalkCr3+, Ni2+, Cd2+100; >90; >9031.91; 19.58; 38.29[91]
Grape stalkCd2+-21.5[140]
Sweet sorghum stalkCu2+-13.32[141]
Sesame stalkNi2+, Zn2+-47.62; 100[142]
Cassava stalksPb2+, Zn2+-163.93; 84.74[143]
Tea stalkCu2+, Zn2+-50.34; 37.87[144]
Stalks of tobacco speciesCr6+99.13; 98.33; 95.0-[145]
Stone
Apricot stonePb2+-111.11[146]
Apricot stoneAl3+, Zn2+-333.3; 500[147]
Olive stoneCu2+, Cd2+, Pb2+, Cr6+77.4; 80.5; 94.5; 460.557; 0.3; 0.581; 2.345[148]
Peach stonePb2+, Cd2+, Cu2+-118.76; 37.48; 32.22[149]
Peach stoneCr6+97-[150]
Lime stonePb2+, Cu2+, Cr6+, As3+30.8; 16.5; 11.5; 8.9-[151]
Activated plum stoneCu2+, Pb2+-48.31; 80.65[152]
Raw plum stonePb2+, Cd2+, Ni2+-9.93; 12.45; 5.63[153]
Mango stoneCd2+, Pb2+-21.05; 1.90[154]
Cherry stonesCr6+81.3-[155]
Vegetable waste
Cabbage wastePb2+-54.945[156]
Cabbage wastePb2+, Cd2+-60.57; 20.57[157]
Cauliflower wastePb2+, Cd2+-47.63; 21.32
Tomato wasteCu2+-25[158]
Tomato wasteCu2+92.0834.48[159]
Tomato wastePb2+-152[160]
Carrot wasteCr3+, Cr6+-86.65; 88.27[161]
Onion wastePb2+, Cd2+92.05; 94.899.74; 14.17[162]
Onion wasteAs3+, Fe2+, Pb2+, Sn2+, Cd2+, Hg2+-2.56; 8.012; 9.957; 7.812; 1.36; 4.95[163]
Garlic wasteAs3+, Fe2+, Pb2+, Sn2+, Cd2+, Hg2+-2.304; 8.459; 10.496; 7.132; 1.47; 5.12
Seeds
Moringa seedsFe2+, Mn2+80.5; 9310.28; 11.641[164]
Avocado seedsCd2+, Cu2+, Pb2+98.23; 99.12; 99.29-[165]
Avocado seedsPb2+, Cr6+-18.9; 3.39[59]
Mangifera indica seedsAs3+940.365[166]
Artocarpus heterophyllus seedsAs3+9329.25
Schizizium commune seedsAs3+920.360
Grape seeds biocharAs3+-5.082[167]
Allium Cepa seedsCr6+, Cd2+, Zn2+, Cu2+, Pb2+-1.78; 1.52; 1.40; 1.62; 1.68[168]
Watermelon seeds biocharPb2+-44.32[71]
Papaya seedsCu2+, Pb2+-97.55; 99.96[169]
Cakes
Gingelly oil cakePb2+-105.26[170]
Neem cakeCu2+, Cr6+, Ni2+93.3; 85.4; 96.6-[171]
Moringa seed cakeCr6+-3.191[172]
Oil palm cakeCu2+, Pb2+, Zn2+-45.01; 125.51; 39.21[173]
Cottom seed cakeCu2+88-[174]
Mustard oil cakeCr6+, Hg2+-29; 48[175]
Rapeseed oil cakePb2+, Ni2+-129.87; 133.33[176]
Rapeseed cakeCu2+, Zn2+-52.196; 29.043[177]
Olive cakeCu2+, Mn2+, Zn2+,Ni2+, Pb2+, Cr3+-30.031; 3.571; 12.693; 5.851; 41.539; 22.193[141]
Black cumin cakeCu2+-106.38[178]
Leaves
Bamboo leaf powderHg2+-27.11[179]
Castor leaf powderHg2+-37.20[180]
Pine leaf powderAs5+-3.27[181]
Cauliflower leaf biocharCu2+, Pb2+-81.43; 224.60[182]
Cabbage leavesPb2+95.676.307[183]
Aloe vera leaf powderPb2+96.2-[184]
Tea leavesCr6+95.4210.64[185]
Mango leavesCd2+-4.08[186]
Eucalyptus leavesAs3+, Hg2+> 94%84.03; 129.87[187]
Mangrove leaf powderCr6+-60.24[188]
Bagasse
Sugarcane bagasseNi2+-2.0[189]
Grape bagasseCu2+-43.47[190]
Agave bagassePb2+, Cd2+, Zn2+-93.14; 28.50; 24.66[191]
Agave bagasseZn2+, Cd2+, Pb2+-8; 14; 36[192]
Sugar beet bagasseCr6+-52.87[193]
Mango bagasseAs3+-1.35[194]
Oil palm bagasseCr6+-111.45[195]
Palm bagassePb2+-162[64]
Hulls
Peanut hullCu2+-14.13[196]
Rice hullZn2+, Cd2+, Fe3+91.017; 84.848; 94.6671.3367; 0.137; 25.403[197]
Groundnut hullCr6+-90[198]
Maize hullCu2+50-[199]
Pistachio hullNi2+> 7514[200]
Soybean hullsCr6+91.991-[201]
Hazelnut hullFe3+83.513.59[202]
Cotton hullPb2+-27.65[203]
Almond hullCr6+> 94.14-[204]
Buckwheat hullsHg2+-243.90[205]
Other agricultural wastes
Green teaCr6+99.98-[117]
Tea waste biocharCr6+-198.0[206]
Brewed tea wastePb2+, Cd2+, Ni2+, Zn2+-1.1947; 1.457; 1.163; 2.468[42]
Hemp fiberCo2+-13.58[207]
Corn cobCd2+, Cr3+, Pb2+-13.577; 18.782; 29.168[119]
Sunflower achene head11.404; 12.206; 22.644
Cocoa podCd2+, Pb2+-12.15; 5.31[154]
Apple juice residuePb2+-108[160]
Coffee wasteCu2+, Pb2+, Zn2+-8.2; 27.6; 8.0[208]
Mango barkAs3+-1.25[194]
Fig sawdustPb2+95.880.645[209]
Coconut coirCu2+13.43–41.560.5–1.15[210]
Coconut husk-activated carbonCr6+95.28173.9[211]
Agricultural waste-activated carbonAs3+, Pb2+-200; 250[212]
Table 8. Regeneration of adsorbents produced from agricultural byproducts. The symbol ↘ means “decreasing”.
Table 8. Regeneration of adsorbents produced from agricultural byproducts. The symbol ↘ means “decreasing”.
AdsorbentMetal IonsAmount DesorbedNo. of CyclesDesorbing AgentReference
Corn stalk fibersCr6+-20NaOH 0.1 M[226]
Pea peelZn2+10 mg g−15HCl 0.5 M[222]
5 mg g−15NaOH 0.5 M
Moringa oleifera biocharCu2+, Zn2+, Ni2+↘ with 9.15%, ↘ with 10.3%, ↘ with 11.78%6; 7; 7H2SO4 0.1 M[223]
Coffee wasteCu2+, Cr6+70–75%,
55–60%
10-[224]
Modified orange peelPb2+, Cd2+, Ni2+43.5 mg L−1, 43.1 mg L−1, 41.3 mg L−130.05 mol/L HCl[227]
Millet huskCr6+74.48%60.1 M NaOH, followed by 0.1 M HCl[228]
Green bean huskSb3+83.7%70.1 M HCl[229]
Sugarcane bagassePb2+, Ni2+89.9%, 96.11%10.1 M HNO3[225]
75.28%, 79.6%1HCl
44.7%, 55.3%1NaOH
Sugarcane bagassePb2+90.05%; 83.97%; 77.92 %; 70.01%10.1 M HNO3, 0.1 M HCl, 0.1 M H2SO4, 0.1 M NaOH[230]
Tamarind fruit seed powderCu2+90%-0.5 N HCl[231]
Pongamia oil cakeZn2+90.14%60.05–0.1 mM HCl; 0.05–0.1 mM H2SO4; 0.01–0.1 mM EDTA[232]
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

Ungureanu, E.L.; Mocanu, A.L.; Stroe, C.A.; Panciu, C.M.; Berca, L.; Sionel, R.M.; Mustatea, G. Agricultural Byproducts Used as Low-Cost Adsorbents for Removal of Potentially Toxic Elements from Wastewater: A Comprehensive Review. Sustainability 2023, 15, 5999. https://doi.org/10.3390/su15075999

AMA Style

Ungureanu EL, Mocanu AL, Stroe CA, Panciu CM, Berca L, Sionel RM, Mustatea G. Agricultural Byproducts Used as Low-Cost Adsorbents for Removal of Potentially Toxic Elements from Wastewater: A Comprehensive Review. Sustainability. 2023; 15(7):5999. https://doi.org/10.3390/su15075999

Chicago/Turabian Style

Ungureanu, Elena L., Andreea L. Mocanu, Corina A. Stroe, Corina M. Panciu, Laurentiu Berca, Robert M. Sionel, and Gabriel Mustatea. 2023. "Agricultural Byproducts Used as Low-Cost Adsorbents for Removal of Potentially Toxic Elements from Wastewater: A Comprehensive Review" Sustainability 15, no. 7: 5999. https://doi.org/10.3390/su15075999

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