Next Article in Journal
Sunlight-Driven AO7 Degradation with Perovskites (La,Ba)(Fe,Ti)O3 as Heterogeneous Photocatalysts
Previous Article in Journal
Exendin-4-Conjugated Manganese Magnetism-Engineered Iron Oxide Nanoparticles as a Potential Magnetic Resonance Imaging Contrast Agent for Tracking Transplanted β-Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Surface Modified Activated Carbons: Sustainable Bio-Based Materials for Environmental Remediation

1
Nanomaterial Lab, Department of Applied Sciences and Chemical Engineering, IOE, Tribhuvan Universtiy, Kathmandu, Lalitpur 44700, Nepal
2
Department of Healthcare Management, Woosong University, Daejeon 34606, Korea
3
Carbon Composite Energy Nanomaterials Research Center, Woosuk University, Wanju 55338, Korea
4
Woosuk Institute of Smart Convergence Life Care (WSCLC), Woosuk University, Wanju 55338, Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(11), 3140; https://doi.org/10.3390/nano11113140
Submission received: 26 September 2021 / Revised: 15 November 2021 / Accepted: 19 November 2021 / Published: 21 November 2021
(This article belongs to the Section Environmental Nanoscience and Nanotechnology)

Abstract

:
Global warming and water/air contamination caused by human activities are major challenges in environmental pollution and climate change. The improper discharge of a large amount of agro-forest byproduct is accelerating these issues mainly in developing countries. The burning of agricultural byproducts causes global warming, whereas their improper waste management causes water/air pollution. The conversion of these waste materials into effective smart materials can be considered as a promising strategy in waste management and environmental remediation. Over the past decades, activated carbons (ACs) have been prepared from various agricultural wastes and extensively used as adsorbents. The adsorption capacity of ACs is linked to a well-developed porous structure, large specific surface area, and rich surface functional moieties. Activated carbon needs to increase their adsorption capacity, especially for specific adsorbates, making them suitable for specific applications, and this is possible by surface modifications of their surface chemistry. The modifications of surface chemistry involve the introduction of surface functional groups which can be carried out by various methods such as acid treatment, alkaline treatment, impregnation, ozone treatment, plasma treatment, and so on. Depending on the treatment methods, surface modification mainly affects surface chemistry. In this review, we summarized several modification methods for agricultural-waste-based ACs. In addition, the applications of AC for the adsorption of various pollutants are highlighted.

1. Introduction

Agricultural wastes have become a global issue in recent years due to inefficient disposal and poor management techniques mainly in developing countries. The majority of agricultural biomass residues are discarded to decompose naturally or are burned openly, which can generate harmful gases, smoke, and dust, thereby resulting in air, soil, and water pollution [1]. Therefore, it is important to address this environmental problem caused by agricultural byproducts (wastes). Utilization of agricultural wastes or biomasses for solving environmental problems could be considered as an effective and green strategy in environmental protection as they are inexpensive, renewable, and locally available in large quantities [2,3,4]. In the literature, agricultural wastes have been proposed as an economic and eco-friendly precursor for ACs [5,6,7,8,9]. The agricultural wastes are rich in lignocelluloses, mainly composed of polysaccharides such as cellulose (35–50%), hemicellulose (20–35%), and an aromatic polymer, lignin (10–25%) [10]. Hemicellulose and lignin act as matrix and encrusting ingredients, respectively, and surround cellulose to generate a skeleton [11]. Being lignocellulosic materials, they have unique chemical compositions with a low content of inorganic materials and a relatively high content of volatile organic materials. Such chemical composition allows them to synthesize ACs with a well-defined pore structure/network and extended surface area. This conversion could solve problems related to undesirable agricultural wastes by turning them into a useful, valuable adsorbent that can address problems associated with environmental pollution and waste management. The facile conversion of pollutant materials into useful ones that can be applied as an adsorbent to remove other pollutants present in the water/air becomes a blessing for material scientists and environmental engineers. Additionally, AC derived from lignocellulosic biomass wastes has the potential to be a low-cost substitute for non-renewable coal-based granular ACs, presuming they have equal or greater adsorption efficiency [12,13]. The biochemical compositions of some agriculture lignocellulosic biomasses are given in Table 1.
AC is a versatile adsorbent having a complex pore structure, a large specific surface area, exceptional chemical stability, and extended functionalities on the outward [15]. These exceptional features have made it widely used in various applications such as water purification, gas separation and storage, catalysis, and supercapacitors. ACs can be synthesized from bio-wastes by a two-stage approach: (i) carbonization of a carbon-rich precursor and (ii) activation of the carbonized product. ACs can be prepared from the carbon-rich precursor by using thermal or chemical methods. By suitable thermal or chemical treatment, the non-carbon and volatile carbon species are removed, thereby leaving behind a fixed carbon mass with a primitive pore structure. The activation process involves exposing the ACs to an oxidizing agent at a high temperature, which forms new pores and vessels. Pores in ACs are usually categorized into three groups; (i) micropores (diameter less than 2 nm), mesopores (diameter ranging from 2 to 50 nm), and macropores (diameter greater than 50 nm) as shown in Figure 1A. The TEM image shown in Figure 1B also shows the hierarchical pore distribution of AC.
Another important characteristic of AC is the chemical composition of its surface. Heterogeneity of the chemical surface is crucial to the surface composition of the AC. The heteroatoms are generally oxygen, nitrogen, halogen, and other elements that are linked to the carbon layers’ edges. These heteroatoms constitute functional groups at the surface of ACs which consequently affect the hydrophilicity and hydrophobicity of the ACs [18]. Since the edges are the prime adsorbing surface, the presence of these heteroatoms at the edges highly influences the adsorptivity of the ACs. Hence, by modifying the morphological features along with the generation of diverse surface functionalities, the performance of ACs can be enhanced.
The functional moieties at the surface determine several characteristics of ACs such as wettability, surface charge, the electron density of the graphitic layers, and catalytic performance. These characteristics make the ACs suitable for various applications such as sensors [19], energy storage [20], catalytic reactions [21], and adsorptions [22,23]. The thermal or chemical post-treatments can change the nature and application of superficial groups [24]. A modified surface functional group on the surface of ACs enhances its adsorption capacity. Carboxyl, carbonyl, phenols, lactones, and quinines are some key functional groups that are helpful for pollutant adsorption. Modification of porous carbons’ surface chemistry (the process of customizing the surface of the AC) might be a promising path toward the specific application of these materials [25]. To regulate the hydrophilic/hydrophobic characteristics of ACs and increase their binding capacity with metals (or metal oxide), physical or chemical techniques are used to alter the superficial effective group on ACs and the ions and molecules put on the surface, making it the active site in particular adsorption. A modified AC with various functional groups can be applied for extracting metallic cations from any kind of solutions, catalysis, waste and hazardous effluent treatment, and so on. Some methods for changing the chemical composition of AC’s surface include surface oxidation, surface reduction, load material alteration, microwave processing modification, acid/alkali treatment, and grafting of different moieties [26,27,28]. The surface modification of ACs is carried out after the activation process. There are three different sorts of changes that may be done, viz. physical modification, chemical modification, and biological modification. In addition, oxidative and non-oxidative treatments have also been reported in the literature [29,30]. In the chemical process, acid, alkali, or salt are used to improve the surface functionalities, whereas the physical process is carried out to improve the physical characteristics. Moreover, the surface modification of ACs can be achieved by the combination of mechano-chemical and thermo-chemical processes. In addition, different functional moieties and metal/metal oxide NPs can be also grafted on the surface of AC for its surface modification [31,32,33,34,35].
Figure 2. Synthesis of activated carbon by various processes and activating reagents [36].
Figure 2. Synthesis of activated carbon by various processes and activating reagents [36].
Nanomaterials 11 03140 g002

2. Synthesis Methods of Activated Carbon

The use of waste biomass as a precursor of porous ACs has gained great attention worldwide. It is due to its bio-waste abundancy and low cost, and it has proved to be the best alternative to conventional raw materials such as petroleum residues, coal, peat, and lignite which are expensive and the majority are non-renewable [37]. There are two main methods in practice for synthesizing carbon, namely pyrolysis carbonization and hydrothermal carbonization [38]. In general practice, hard bio-waste materials are carbonized using pyrolysis carbonization, whereas soft bio-based materials are carbonized using hydrothermal carbonization. The activated carbon can be prepared by maintaining two basic steps; i) carbonization and ii) activation. Figure 2 depicts the primary activation process for the preparation of activated carbon from biomass.

2.1. Pyrolysis Carbonization

Carbonization is the oldest process of altering biomass into carbon material for the service of humankind. It requires a relatively high temperature and is a slow and long process, and it converts organic substances into carbon or a carbon-containing residue (biochar) through pyrolysis of raw material in a furnace under an inert gas atmosphere by removing volatile, non-carbon species such as nitrogen, oxygen, and hydrogen, intensifying the carbon content [39]. During the degasification process, narrow pore structures of precursors start to develop followed by the removal of lingering substances formed when rising the temperature. Moreover, sometimes this accumulation may be the cause of collision of some lingering substances and collapse the walls of pores resulting in hydro cracking and carbon deposition [40]. In this process, temperature has the most remarkable effect along with the heating rate, the presence of an inert atmosphere and its rate, and finally, the process duration. Normally, a carbonization temperature higher than 600 °C results in a reduced yield of char and an increased liquid and gas release rate [41]. A higher temperature will also increase ash and fixed-carbon content and lower the amount of volatile matter. Thus, high temperatures result in better-quality char but a decreased yield. The decrease in the yield is thought to be caused by the primary decomposition (de-volatilization) of biomass at a high temperature and the secondary decomposition (cracking) of biochar residue. Consequently, higher temperature produces a better quality of biochar [42]. Undoubtedly, pyrolysis carbonization is the best method to produce large amounts of porous carbon products with excellent properties. Currently, this process, followed by chemical activation, is widely used to obtain bio-based hierarchical porous ACs for commercial production. The formation of well-ordered porous ACs employing this process is shown in Figure 3 where tabah bamboo is used as a carbon precursor [43]. SEM images (Figure 3F,G clearly show that the macro-, meso-, and micropores of ACs are well ordered.

2.2. Hydrothermal Carbonization

Hydrothermal carbonization is an alternative auspicious thermochemical process that can transmute biomass into ACs economically and environmentally friendly. Hydrothermal carbonization embraces heating raw material disseminated in an aqueous solution and autoclaving it at temperatures between 150 and 350 °C for about 2–24 h at saturated pressure (Figure 4). During this step, water-soluble components and a carbon-rich hydrophilic solid called ‘hydro char’ are formed [17,44]. The hydrothermal carbonization process has limited perquisites in terms of preparation and treatment of biomass, mainly no pre-drying needed for wet waste as in other thermal treatment processes which makes it an economically attractive alternative. The final product formation is pendent on different process parameters such as reaction time, nature of pressure, and temperature [45]. This process is led by dehydration and decarboxylation reactions which are exothermic and render the process self-subsistent after activation. At subcritical conditions, biomass is first converted into monomers by hydrolysis and then into soluble organics by dehydration and fragmentation. A reduction in pH catalyzes the initial hydrolysis step (acid catalysis) [46], and the solution concentration increases due to polymerization and condensation of material and hence nucleation, and thus the growth takes place. The chemical structure of hydro char is almost similar to that of natural coal (core–shell structure with a hydrophobic core and a stabilizing hydrophilic shell) containing oxygen functional groups such as hydroxyl, phenol, carbonyl, or carboxyl which makes it a good adsorbent. The main superiority of this process over other thermochemical conversion technologies is its potentiality to convert the wet feed stock into a solid carbonaceous product (hydro char) with high yields without exposing it to dehydration and drying [47].
Soft bio-based materials can be easily converted into biochar by this process. Thus obtained biochar can be activated by different activating agents using a tubular furnace for its surface modification. Recently, Zhao et al. reported the fabrication of ACs from tobacco rods (soft bio-waste) which produce a hierarchical porous structure with the coexistence of micro-, meso-, and macropores [48]. Moreover, the obtained ACs have numerous oxygen and nitrogen-based functional groups with excellent electrical conductivity. The hydrothermal process is considered as a simple and facile technique for the fabrication of novel materials. The process can be used for doping different metal or metal oxides on the surface of ACs. A three-step process, i.e., pyrolysis carbonization, chemical activation, and hydrothermal treatment, can be used to obtain an effective photocatalyst from bio-waste materials. Recently, our research group reported the fabrication of ZnO-rod-decorated ACs using the hydrothermal process [35]. In this work, hard bio-waste material (Lapsi seed stone) was used for pyrolysis carbonization with phosphoric acid activation followed by hydrothermal treatment with a ZnO precursor as shown in Figure 4. Phosphoric acid activation can provide the functionalities on the surface of ACs which can act as a nucleation site for the growth of ZnO nanorods on the surface of porous ACs during hydrothermal treatment. The stable incorporation of ZnO nanorods into ACs prevents the loss of the photocatalyst during recovery which can be effectively used in water treatment.

3. Modification Methods of Activated Carbon

Surface modification of ACs can develop proper physicochemical properties suitable for their application in environmental remediation. It is the action of tailoring the surface of ACs by adjusting the physical, chemical, or biological characteristics on their surface to produce desired physical and chemical properties [49]. Different activation methods with various activating agents can introduce unique physicochemical properties to the ACs. Synthesis of ACs by various activating reagents is presented in Figure 1, and their detailed discussion is provided in the following sections.

3.1. Acid Modification

It is a typical wet oxidation process applied for the fabrication of ACs from biomass. During acid oxidation, the porous carbon surfaces can eliminate mineral components and boost the surface’s hydrophilic character [24]. Strong mineral acids and powerful oxidants such as HNO3, H2SO4, HCl, H3PO4, H2O2, and HClO are applied for this modification [50]. Nitric acid and sulfuric acid are prominent, whereas weak organic acids are seldom used because of their weak ionization ability. Acidification of ACs enhances their acidic behavior and their surface hydrophilicity by reducing the minerals present on them. Such adsorbent materials with an acidic surface can introduce oxygen-containing functional groups such as hydroxyl, carbonyl, carboxyl, quinone, lactone, and carboxylic anhydride, etc. The presence of such functional groups on the outer surfaces or margins of the basal plane on ACs can play a notable role in adapting the material’s chemical nature [51].
An acid-treated AC can generate a positive charge on its surface, enhancing the adsorption of metal cations due to its ability to form metal complexes with anionic acid groups. Different research groups have adopted this basic principle for the adsorption of heavy metals using acid-activated carbon with various precursors of AC (derived from either coal-based materials or from biomass). Huang et al. recently reported that nitric acid is an effective activating agent for lignite which can be effectively used for the adsorption of Pb [52]. Increasing the amount of polar oxygen-containing functional groups such as hydroxyl, carbonyl, and carboxyl and the introduction of nitro groups on the surface of activated lignite make it a perfect adsorbent for heavy metals. Such chemical moieties improve the charged characteristics, the polarity of the adsorbent surface, and the metal ion adsorption capability. Along with the nature and oxidizing strength of acid, the degassing caused by heating at a higher temperature can alter the adsorption capacity of ACs. Aggarwal et al. [53] used nitric acid, ammonium persulphate, hydrogen peroxide, and gaseous oxygen at 350 °C to oxidize granular ACs and fibrous ACs. The ACs were well outgassed at different temperatures to eliminate surface chemical structures. They found that the adsorption of Cr (III) increased with oxidation and reduced with degassing. Cr (VI) adsorption, on the other hand, reduced with oxidation and rose with degassing. This result shows that metal ions having different oxidation states may have different affinities of adsorption on the same surface. Park and Jang performed HCl treatment on carbon and reported that this treatment could increase the number of surface oxygen complexes, which enhanced the active adsorbent site and, therefore, improved the reduction rates of Cr (VI) [54]. The insertion of acidic oxygen effective groups by HNO3 oxidation into AC was reported by Jia and Thomas [55], and the potential of modified carbon for cadmium adsorption from water was investigated. The main surface species absorbed were carboxylic acid groups, with phenol and quinone groups added during the oxidation process significantly increasing cadmium adsorption [56].
Acid treatment can show different adsorption affinities toward different heavy metals. For example, the surface of Chemviron F400, a commercial AC, was modified by the oxidation with HNO3 to introduce a range of functional groups on its surface [57]. After oxidation, the exterior surface area and pore void were found to be reduced. The carbon surface, on the other hand, developed an acidic nature, with carboxylic groups dominating the surface functional groups. The acid-treated sample demonstrated poly-functionality and cation-exchange characteristics throughout a wide pH range. Nickel (II), cobalt (II), copper (II), zinc (II), and manganese (II) were removed from both carbon samples. It was discovered that the affinity order Mn2+ < Co2+ < Ni2+ < Cu2+ > Zn2+ corresponded to the stability of their metal complexes. Modified ACs were used to study the removal of colors in addition to metal ions. Ayan et al. [58] used origanum onites stalks in their pristine and acid-modified forms using HNO3 and H3PO4 solutions as adsorbents to remove Basic Red 18 and methylene blue from water. They recorded that the modified ACs increased the adsorption capacity by 56–63% for Red 18 dye and 125% for methylene blue dye. Carvajal-Bernal et al. [59] performed a surface modification of ACs with H3PO4 without subsequent heating which enhanced the adsorption rate of 2, 4-DNP. It reflects the changes in the surface chemistry of carbon associated with the interaction between H3PO4 and oxygen groups present on the surface. The modification causes the acid to be dispersed on the carbon matrix and thus facilitates the retention of 2, 4-DNP.
The adsorption capacity of acid-treated ACs may depend upon the process temperature. Pradhananga et al. [60] fabricated nanoporous activated carbon materials from bamboo (Bambusa vulgaris) cane by chemical treatment with H3PO4 at different temperatures (400, 500, and 600 °C) for removal of wool carpet dyes, Lanasyn orange and Lanasyn gray. Irrespective of the carbonization temperature, it was reported that all the ACs contained oxygenated surface functional groups such as –OH, –CO, -COOH, and lactones. They found that the bamboo-derived AC prepared at 600 °C is a favorable adsorbent for the wool carpet dyes. Shrestha et al. [61] investigated the efficiency of AC prepared from Lapsi (Choerospondias axillaris) seeds by carbonization and subsequent acid treatment with concentrated H2SO4 and conc. HNO3 for heavy metal adsorption. The FTIR results confirm that the chemical treatment using concentrated sulfuric acid carbonized the cellulose structure of the raw material differently than the physical treatment and further oxidized carbon with concentrated HNO3 considerably increased the number of functional groups. Furthermore, the FTIR spectrum of the pristine carbon was significantly different from the spectra of acid-treated carbon. They demonstrated that heavy metal ions form complexes with the surface carboxyl and lactone groups in the midst of hydrogen ion development. The sorption capacities of the carbon for removal of lead and nickel, respectively, were found to be (qm for Pb = 424 mg g−1 and qm Ni = 70 mg g−1 at pH = 5) in the model solution. Some current research conducted on the alternation of ACs using acid treatment is summarized in Table 2.
According to the above discussion, acidic modification of AC can produce a large number of effective moieties on the surface of carbon which has been found to improve the sorption of different pollutants due to changes in its surface chemistry. As a result, the capacity of acidic treatment is directly linked to the quantity of acidic functional groups [68].

3.2. Alkaline Modification

A base (alkaline) treatment of AC generates a positive charge on its surface which aids in the adsorption of different negatively charged species. The easiest way of enhancing the basic surface properties in porous carbons is the treatment of ACs at high temperatures (400–900 °C) in an inert hydrogen or ammonia atmosphere [69,70]. Alkali treatment can improve the relative content of alkali groups as well as improve the surface non-polarity. Therefore, this process can enhance the adsorption capacity of ACs for non-polar substances.
The surface of ACs can be modified by treatment with NaOH, KOH, Na2CO3, Na2SiO3, and oxides. When AC is treated with NH3 at 400–900 °C, it develops the basic nitrogen functionalities on the carbon surface [71,72]. Doping of nitrogen functionalities can be achieved by treating it with nitrogen precursors (such as ammonia and amines) or activating it in a nitrogen-rich environment [73,74]. The induced amide, imide, lactone, pyrrolic, and pyridinic groups usually provide the suitable property that can uplift the interplay linking adsorbents and acidic breed by dipole–dipole interchange, hydrogen, and covalent bonding. Chen et al. [75] investigated the refitting of the surface of ACs through thermal treatment in the presence of ammonia. They reported a fourfold increase of perchlorate adsorption by the AC samples prepared at a temperature of 650–700 °C.
Recent important findings on the modification of the ACs by alkaline treatment are found to be effective for adsorption applications. Zheng et al. [76] found that the effect of NaOH modification on ACs significantly reduces the oxygen-containing functional groups on the surface of ACs. The pore void and the distinct surface area are increased with alkali concentration leading to an increase in the adsorption capacity of methane. When treating AC with NaOH, Chiang et al. [77] discovered that the AC enhanced the congregation of phenolic functional sites on the surface. J. Przepiorski [78] reported the impact of the heating situation on the adsorbent through ammonia treatment and found that, at an optimal temperature of 700 °C, the adsorbent’s absorption capability toward phenol rose by 29%. P-Chlorophenol (PCP) adsorption from an aqueous solution on ACs with basic characteristics has been investigated [79]. The preparation of ACs can be carried out by two methods. In the first method, a commercially available activated carbon (CWZ22) was modified at high temperatures in an ammonia-, nitrogen-, and hydrogen-rich environment. In the second method, AC was carbonized and subsequently activated to make chars using CO2 and steam as activating agents. The fundamental of broiled CWZ activated carbon was primarily caused by the deletion of oxygen functions, whereas the basicity of N-polymer-based ACs was caused by the high nitrogen content (2.42–5.42 wt %). The results show that π-π interaction allying the phenol ring and the graphene layers was primarily responsible for aqueous solution adsorption on the surface of ACs. PCP uptake was shown to have a modest contribution from nitrogen-derived basic sites. Kasnejad et al. [80] introduced a new method for nitrogenating commercial AC by treating the adsorbent with heat: (i) under NH3 atmosphere, after pre-oxidation with HNO3, and (ii) without pre-oxidation, for Cu (II) adsorption. The result shows that the pre-oxidation of adsorbent increased the amount of nitrogen functional groups on the structure of the adsorbent, and the modified adsorbent showed a higher Cu (II) adsorption capacity. Similarly, Ofudje et al. [81] employed an alkaline modified coconut shaft as an adsorbent removal of Pb2+ from an aqueous solution. They reported that the optimum removal of Pb2+ was 17.6 and 22.1 mg/g by pristine and alkaline-treated biomass, respectively, at pH 4.0. The authors claimed that this improvement in the adsorption capacity of alkaline-treated AC was due to an appropriately refined feature of the material. Joshi et al. [82] evaluated the characteristics of NaOH-treated nanoporous activated carbons derived from Lapsi (Choerospondias axillaris) seed stone, synthesized under different conditions of activation/carbonization. FTIR data confirmed that despite the different preparation conditions used, the ACs contained common oxygen-containing surface functional groups such as –OH, –C O, COOH, and lactones. The prepared carbon with carbonization at 400 °C for 3 h had the highest surface area of around 1000 m2 g−1 and showed a high adsorption capacity of around 200 mg g−1 for removing methylene blue dye from an aqueous solution. Modification of the AC surface using ammonia gas is also a popular strategy to improve the adsorption capacity of ACs. Different reports show that the CO2 adsorption capacity of NH3-modified ACs is remarkably increased which is probably the affinity of CO2 molecules (acidic nature) toward the basic surface of ACs. Moreover, other pollutants are also effectively removed by NH3-treated ACs. Some reports of the use of gaseous ammonia to modify activated carbon and its application in adsorption are provided in Table 3.
Altogether, one can understand that under alkaline conditions, hydroxyl ion reacts with surface functional groups of ACs. This change also increases the positive charge on the surface of the ACs which are capable of intensifying the negatively charged moieties from water.

3.3. Impregnation

Impregnation is defined as the uniform distribution of other species (in nano/micro scale) on the surface of carbon materials [35,89]. It is a process of surface decoration of porous material with metal/metal oxide/chemicals. Dry, as well as wet, impregnation methods can be employed for this process. In the dry impregnation method, a solvent can be applied to pack the aperture of the adsorbent. However, in the wet impregnation method, an excess solvent is added after the pores are filled. Metals or polymeric compounds that do not affect pH are often used as impregnating materials. According to Henning and Schafer [90], impregnation of ACs can improve the material’s prevailing properties by increasing catalytic oxidation abilities, promoting harmony among the ACs, and enhancing the material’s adsorption capabilities of the component as a passive penetrable bearer. Hydroxides, carbonates, chromates, and nitrates are common examples of impregnating materials. It is widely accepted that impregnation with appropriate chemicals significantly enhanced the adsorption abilities of AC for eliminating hazardous chemicals such as heavy metals (arsenic, mercury, and cadmium), fluoride, and cyanide in water [91]. The impregnation of ACs with metals such as silver [92], copper [93], aluminum [94], and iron [95] has gained wide interest due to their higher adsorption capacity. Moreover, the impregnated adsorbent can also provide the advantage of stability and favorable regeneration capacity [96].
Compounds of iron have been mostly utilized in fabricating the impregnated ACs. Huang and Vane [97] remodeled the ACs with a solution of iron salt to enhance their arsenic adsorption capacity. They reported that Fe-modified AC exhibits 10 times greater removal efficiency compared to pristine ACs. The efficacy of arsenic withdrawal was allocated to surface assimilation of ferrous ions and arsenate complexes formation.
Chang et al. [98] described a novel multistep process for impregnating granular activated carbon (GAC) with ferrous chloride as a precursor for excellent removal efficiency of arsenic from wastewater. Impregnation of Fe on ACs can form nano-sized particles that existed in both amorphous and crystalline forms. Iron-permeated GACs (Fe-GACs) were served with NaOH to sustain the Fe in the GAC, and the impregnated iron was found to be extremely stable in water treatments across a wide pH range. To evaluate arsenic adsorption capacities, an isotherm test was performed using synthetic-arsenate-contaminated drinking water. When the iron concentration was less than 4.22%, the iron usage efficiency remained high (40–46 mg As/g Fe). When the iron concentration was raised to 12.13%, the iron utilization efficiency dropped to 14 mg As/g. For arsenate adsorption, Fe-GACs showed better performance in an acidic environment compared to a basic one. With a pH less than 6.0, the arsenate clearance rate was about 100 percent, but with a pH of more than 7.0, it dropped rapidly. Vitela-Rodriguez et al. [99] created a variety of ACs that were treated with iron hydro (oxide) nanoparticles and evaluated for their capacity for the adsorption of arsenic from water. The maximum arsenic adsorption capability of the modified ACs ranged from 370 to 1250 g/g. Meanwhile, if the pH of the solution was increased from 6.0 to 8.0, the arsenic adsorption magnitude decreased by 32%. In contrast, the temperature had no mastery on the adsorption of arsenic. Furthermore, the presence of competing anions such as SO42−, Cl, and F significantly reduced arsenic adsorption when groundwater was utilized in the experiment.
Monser and Adhoum [100] modified ACs by impregnating them with tetrabutyl ammonium (TBA) and sodium diethyl dithiocarbamate (SDDC) for the intake of Cr (VI), Cu (II), Zn (II), and CN ions. They found from the experimental results that TBA-modified ACs display a five times greater removal performance than pristine ACs. Similarly, SDDC-modified ACs show four times greater intake of Zn (II) and Cu (II) and two times greater adsorption of Cr (VI) than pristine ACs. It is caused by the ion exchange mechanism by the impregnated chemical species on the surface of ACs. Dastgheib et al. [74] reported the effectiveness of the elimination of iron-impregnated ACs’ dissolved organic matter (DOM) from natural water. They found that ammonia treatment of oxidized carbons at a high temperature enhances DOM absorption efficiency by up to 120%. Shah et al. [101] introduced iron on AC by wet oxidation and applied both impregnated and unimpregnated ACs for the elimination of methylene blue (MB). They reported that Fe-AC exhibits MB removal of up to 95%, higher than unimpregnated ACs in the pH range of 7–10. In addition to iron, some other metals such as Al, Mn, Ni, Ag, and Zr are also used to impregnate AC. Sonal et al. [102] used zirconium-impregnated ACs to remove a reactive dye and reported a higher adsorption capacity of 500 mg/g by the pristine ACs compared to 506 mg/g by zirconium ACs. Some chemically impregnated species on AC and their potential applications are presented in Table 4.
It can be concluded that the primary advantages of impregnated ACs involve the expansion of the catalytic characteristic and enhanced synergic effect between ACs and impregnating species. However, to ensure the feasibility of the process, more research in the field of leaching of impregnated materials on ACs should be emphasized.

3.4. Ozone Treatment

One way to change the carbon’s structure is by its exposure to ozone due to oxidation. The combined ozone AC treatment modifies the surface groups and chemical composition of ACs [107,108]. The single procedure of interaction of ozone and ACs may be the best alternative for the removal of toxic organic compounds [109]. Some previous studies have reported that, besides adsorption, some other phenomena such as reactions between ozone, the adsorbed organic matter, and the generated free radicals can enhance the efficiency of the system [110,111].
The influence of ozonation on the surface topography and chemical alteration of two granular ACs, F400 and AQ 40, and their capacity for the adsorption of phenol (P), P–Nitro Phenol (PNP), and P–chloro phenol (PCP) from an aqueous solution have been reported [112]. It is found that the ozone-treated carbons’ porous structure remained almost unaltered compared to the plain AC. The ozone-treated AC at either room temperature (25 °C) or 100 °C produced acidic surface oxygen groups on its surface. At 25 °C, carboxylic groups predominated, but at 100 °C, a more uniform dispersal of hydroxyl, carbonyl, carboxylic, and lactonic groups emerged. Physical adsorption was caused by dispersive interactions between the electrons of the aromatic ring and those of the carbon basal planes, whereas irreversible adsorption was caused by the decay incorporation of phenolic compounds brought about by basic surface oxygen groups (SOGs). The exposure of GAC to ozone reduced its capacity to adsorb P, PNP, and PCP at ambient temperature, but when ozone was administered at 100 °C, carbon adsorption was not inhibited, and in certain cases (P and PNP on F400), it was significantly accelerated. Another research study looked at the use of ozone with heat treatment to alter the surface chemistry of ACs [113]. Cherrystones were carbonized at 900 °C and activated in CO2 or steam at 850 °C to create ACs. This product was ozonized at room temperature. It was concluded that the use of ozone in combination with heat treatment can make ACs a potential candidate in water treatment. Valdes et al. [114] reported the effect of ozone modification on the surface properties of ACs. Commercial AC, namely FILTRASORB 400 (Calgon Carbon Corp.), was treated with ozone with a constant flow (76 mg of O3/min) at 25 °C and 1 atm for different exposure times. Its adsorption proficiency was appraised by employing methylene blue adsorption. It is reported that using ozone and ACs together in wastewater treatment could change the chemical characteristics of the surface of ACs [115]. When activated carbon is exposed to ozone gas for a relevant time, the chemical entities of the carbon surface change, resulting in acid groups such as carboxylic acid, anhydride, and lactones. The adsorptive behaviors of the original carbon are influenced by changes in the oxygenated group distribution on the surface of ACs. The ozone activity has an effect on the textural properties of carbon and enhances the adsorptive properties.

3.5. Plasma Treatment

Plasma treatment is becoming a prominent technology for the nano-scale surface modification of different materials. It is a source of a strong electric field, charged particles with sufficient energy, powerful oxidizing as well as reducing species, UV light, ultrasound, electrohydraulic cavitations, and shock waves [116,117,118] which are responsible for carrying out the plasma treatment. One of the most widespread processes for plasma treatment of AC is the plasma oxidation process. In this process, the ACs can be subjected to plasma in the presence of regulated air or oxygen under vacuum or atmospheric pressure. During plasma oxidation, chemical addition of oxygen takes place at the carbon surfaces as oxygen free radicals and increases the surface acidity which alters the surface chemistry of ACs significantly.
Recently, the use of plasma modification of ACs for the removal of hazardous organic compounds from wastewater has gained significant development. The treatment of ACs with oxygen plasma can be carried out to produce oxygen-containing functional groups onto the surface of ACs [119]. It is reported that oxygen functional groups such as C6H5OH and O-C = O are increased while the specific surface area is slightly decreased during a plasma treatment. Because of the high oxygen concentration and well-developed micropores, plasma treatment is very facile and economic for removing highly toxic materials.
The surface modification of ACs by low (cold) oxygen plasma for the study of adsorption of dibenzothiophene (DBT) in diesel fuel was studied by Zhang et al. [120]. They found that, owing to an improvement in the surface congregation of acidic oxygen-containing functional groups, the adsorption efficiency of the plasma-treated ACs rose by 49.1% compared to the pristine ACs. In another study, the surface modification of AC fiber was redesigned by gilding arc discharge technology for the adsorption of Acid Orange II from wastewater [121]. It was claimed that the plasma-modified AC’s adsorption efficiency increased by 20.9% compared to the original one. This is owing to improvement in the amount of surface oxygen-containing functional groups.

3.6. Biological Modification

Biological AC technology is a combination of degradation by efficient microbes and an AC-based adsorption process [122]. In the biological modification, bacteria are confined within the ACs and begin to accumulate under a suitable temperature and nutrients for growth [123]. It is found that biological modification has a crucial distinct outward area and a well-developed pore structure on ACs. Therefore, they can productively engross solvated oxygen and organic subject in water. In addition, the attached microorganisms convert the degradable part, carbon dioxide, waste products, and biomass prior to material employing adsorption situated on the ACs.
The biological AC system is based on the interaction among AC particles, microorganisms, contaminants, dissolved oxygen, and other pollutants present in the soluble state in wastewater. Microorganism adsorption on AC may be advantageous in various ways. For example, carbon-bed life may be extended in biologically modified ACs by converting a fraction of refractory organics to biodegradable organics. The microorganism then converts the biodegradable portion before this substance may employ adsorption situated on the ACs. Furthermore, the biofilm produced on ACs can alter its surface energy, potentially improving its adsorptivity [124].
Chen, M. et al. discovered that adding polyvinyl alcohol (PVA) to powdered AC as an entrapment agent for immobilized microorganisms accelerates the mass transfer process of the reaction system and has the efficacy of both adsorption and entrapment [125]. Gutierrez et al. [126] found that ACs served as a carrier for microbial immobilization and improved the treatment efficacy of p-phenol by around 30%. As a result, AC has a synergistic effect on microorganisms and their processing capacities, boosting their adsorption efficiency significantly.

3.7. Microwave Treatment

In recent years, microwave treatment has been proposed as a potential strategy for the modification of AC due to its benefits of fast temperature increase, uniform temperature distribution, and energy savings over traditional heating methods [127,128]. Microwaves supply energy to the carbon particles using dipole rotation and ionic conduction and have been long-established in the fabrication and regeneration of ACs [129]. Furthermore, microwave processing networks are also somewhat flattened, convenient, sustainable, and profitable which can improve the adsorption capacity of ACs. Liu et al. [130] prepared modified bamboo-based ACs by microwave treatment under an inert atmosphere of nitrogen. The ACs showed enhanced adsorption of methylene blue after the modification, which was attributed to the increased micropores in the modified ACs. ACs from corncob furfural residue have also been prepared by microwave irradiation using zinc chloride [131]. The impact of the ratio by weight of ZnCl2 and the time of soaking the solution, time of microwave radiation, and the ZnCl2 solution’s pH value was examined. It was reported that the best-fitted parameter for microwave-modified ACs was ZnCl2 to corncob furfural residue with a ratio by weight of 3.5:1, irradiation time 20 min, soaking time 12 h, and pH value of the ZnCl2 solution.

3.8. Grafting of Different Moieties

The modification of activated carbon has received much attention to increase the selective adsorptive properties of ACs. Among the several methods of modification of ACs, grafting is also one of the promising methods. Grafting of chemical agents onto activated carbon surface facilitates the binding of functional groups from the chemical agents to the porous carbon surface assisted by new bond formation caused by chemical reaction. Functionalization of carbon surface by grafting is terminated usually by its outpouring in chemical solvents furnishing functional moieties that are eventually bonded chemically to the carbon chain affording the much-needed temporary stability [132]. The leading advantage of functionalization by grafting on the surface of ACs is the gripping of functional groups by bond formation [133]. However, the disadvantage of grafting functional groups is that it is less sensitive compared to free groups due to the aversion outcome from bond formation [27]. This method is also affected by limited chemical loading accomplished when surface carbon has been utterly occupied.

4. Application of Bio-Based Modified Activated Carbon in Adsorption

Due to the large specific surface area, low production cost, and ease of handling and modification, bio-based ACs can be considered as ideal materials in different applications [134]. The modified ACs are applied for various applications such as gas adsorption, energy storage, and water treatment. It can be exploited in the adsorption of CO2, NOx, organic pollutants, dyes, etc. ACs are increasingly used in water purgation for the removal of pollutants and contaminants. It is reported that around 80% of the AC around the world is being used in a liquid phase or aqueous solution [135]. The adsorption of pollutants from the solution is an exchange process that mainly depends on the porous structure and surface area of the ACs [16,136]. The adsorption mechanism may be influenced by the electrostatic or non-electrostatic interaction between the solute and the carbon surface. Electrostatic interactions arise due to the ionization of electrolytes, whereas the non-electrostatic interactions are largely due to dissipation and aquaphobic interaction.
Compared to pristine ACs, surface-modified activated carbon shows better adsorption and catalytic activity. Different research groups have reported various surface-modified ACs. Recently, Joshi et al. [137] investigated the efficiency of a magnetic Fe3O4/sugarcane bagasse (Saccharum officinarum) AC composite for the adsorption of arsenic (III) from aqueous solutions. AC was prepared from sugarcane bagasse by treatment with H3PO4 at 400 °C. The AC was composited with Fe3O4 particles by facile one-pot hydrothermal treatment. They found that the composite could remove the arsenic from the water far more effectively than plain AC. The highest percentage of arsenic removal was found at pH 8, adsorbent dose of 1.8 g/L, and contact time of 60 min. The equilibrium experimental data showed the best fit to the Langmuir model with a maximal adsorption capacity of 6.69 mg/g.
ACs have been studied by many researchers as promising catalysts for toxic gases. Altering the pore size and surface chemistry of ACs can effectively change the catalytic oxidation of some toxic gases such as NO. Recently, You et al. reported that ACs based on hydrothermally HNO3-treated wood can have better catalytic activity toward NO compared to metal oxide, especially at room temperature. The adsorption of pollutants by ACs from biomass is presented in Table 5.

5. Concluding Remarks

A comprehensive study based on various modification techniques on AC derived from biomass to increase its adsorption ability was presented. Various techniques have been established to develop a new generation of various surface functional groups in ACs, including acid treatment, base treatment, salt treatment, ozone treatment, impregnation treatment, plasma treatment, and microwave treatment. The experimental results indicate that the acidic treatment requires uptake of metal ions while base treatment is applied mainly for the removal of anionic and organic compounds from an aqueous environment. Findings on ACs with surfactant amendments, which are designed to remove contaminants from aqueous solutions, are limited and therefore require additional exploration to impart excellent performance. The plasma technique for generating an appropriate charge on superficial carbon is an efficient method to create the desired charge on the AC surface making it beneficial for the elimination of toxic pollutants. In addition, carbon-based materials with basic surfaces to incorporate CO2 adsorption in order to attenuate global warming are extremely desirable. On the other hand, further investigation is required of the surfactant modification of ACs that can be tailored to eliminate pollutants from an aqueous environment to yield excellent performance as experimental findings in this field are limited. To select a modifying agent, some factors such as the charge (positive, negative, or neutral), state (solid, liquid, gaseous), size, solubility, and pH of the system are required. However, some of the major drawbacks of modification methods include the associated costs involved in the process and the leaching of hazardous chemicals employed in the alternation process into the treated water. Considering the effects of increased/decreased smell of certain contaminants, selective adsorption may be generated, and the reclamation of the adsorbents should be utilized in cyclic measurements and quietly. Considering the above, the area demands novel, sustainable, and environmentally friendly approaches for activated carbon modification. Thus, the authors suggest that existing modification procedures should be improved as these techniques enhance the chemical surfaces of ACs while limiting the degradation of textural features. It would also promote the dominance of the AC adsorption method of water/wastewater treatment and other applications.

Author Contributions

Conceptualization, M.K.J. and H.R.P.; methodology, M.K.J., S.J. and R.K.S.; software, M.K.J. and A.A.K.; validation, H.R.P. and M.P.; formal analysis, B.P. and H.R.P.; investigation, H.R.P.; resources, H.R.P. and M.P.; data curation, M.K.J., R.K.S. and S.J.; writing—original draft preparation, M.K.J., B.P., A.A.K. and H.R.P.; writing—review and editing, M.K.J., B.P. and H.R.P.; visualization, H.R.P. and M.P.; supervision, B.P., M.P. and H.R.P.; project administration, H.R.P. and M.P.; funding acquisition, H.R.P. and M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. NRF-2019R1A2C1004467). This research was also supported by the World Academy of Science, Trieste, Italy (TWAS)/Swedish International Development Corporation Agency (Sida) (project number; 18-168RG/CHE/AS_G-FR3240303651).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors are thankful to the National Research Foundation of Korea (NRF-2019R1A2C1004467) and the World Academy of Science, Trieste, Italy (TWAS)/Swedish International Development Corporation Agency (Sida) (project number; 18-168RG/CHE/AS_G-FR3240303651) for supporting this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barbieri, L.; Andreola, F.; Lancellotti, I.; Taurino, R. Management of agricultural biomass wastes: Preliminary study on characterization and valorisation in clay matrix bricks. Waste Manag. 2013, 33, 2307–2315. [Google Scholar] [CrossRef] [PubMed]
  2. Pant, B.; Park, M.; Kim, H.-Y.; Park, S.-J. CdS-TiO2 NPs decorated carbonized eggshell membrane for effective removal of organic pollutants: A novel strategy to use a waste material for environmental remediation. J. Alloys Compd. 2017, 699, 73–78. [Google Scholar] [CrossRef]
  3. Pant, B.; Ojha, G.P.; Acharya, J.; Park, M. Eggshell membrane templated synthesis of Ni/MoC decorated carbon fibers with good electrochemical behavior. Int. J. Hydrog. Energy 2021, 46, 2774–2782. [Google Scholar] [CrossRef]
  4. Roberts, D.A.; Paul, N.A.; Cole, A.J.; de Nys, R. From waste water treatment to land management: Conversion of aquatic biomass to biochar for soil amelioration and the fortification of crops with essential trace elements. J. Environ. Manag. 2015, 157, 60–68. [Google Scholar] [CrossRef] [PubMed]
  5. Yahya, M.A.; Al-Qodah, Z.; Ngah, C.W.Z. Agricultural bio-waste materials as potential sustainable precursors used for activated carbon production: A review. Renew. Sustain. Energy Rev. 2015, 46, 218–235. [Google Scholar] [CrossRef]
  6. Matei, E.; Râpă, M.; Predescu, A.M.; Țurcanu, A.A.; Vidu, R.; Predescu, C.; Bobirica, C.; Bobirica, L.; Orbeci, C. Valorization of Agri-Food Wastes as Sustainable Eco-Materials for Wastewater Treatment: Current State and New Perspectives. Materials 2021, 14, 4581. [Google Scholar] [CrossRef]
  7. El Gamal, M.; Mousa, H.A.; El-Naas, M.H.; Zacharia, R.; Judd, S. Bio-regeneration of activated carbon: A comprehensive review. Sep. Purif. Technol. 2018, 197, 345–359. [Google Scholar] [CrossRef]
  8. Mensah-Darkwa, K.; Zequine, C.; Kahol, P.K.; Gupta, R.K. Supercapacitor Energy Storage Device Using Biowastes: A Sustainable Approach to Green Energy. Sustainability 2019, 11, 414. [Google Scholar] [CrossRef] [Green Version]
  9. Lee, K.-C.; Lim, M.S.W.; Hong, Z.-Y.; Chong, S.; Tiong, T.J.; Pan, G.-T.; Huang, C.-M. Coconut Shell-Derived Activated Carbon for High-Performance Solid-State Supercapacitors. Energies 2021, 14, 4546. [Google Scholar] [CrossRef]
  10. Arora, A.; Nandal, P.; Singh, J.; Verma, M.L. Nanobiotechnological advancements in lignocellulosic biomass pretreatment. Mater. Sci. Energy Technol. 2020, 3, 308–318. [Google Scholar] [CrossRef]
  11. Ingram, L.O.; Doran, J.B. Conversion of cellulosic materials to ethanol. FEMS Microbiol. Rev. 1995, 16, 235–241. [Google Scholar] [CrossRef]
  12. Martin, M.J.; Artola, A.; Balaguer, M.D.; Rigola, M. Activated carbons developed from surplus sewage sludge for the removal of dyes from dilute aqueous solutions. Chem. Eng. J. 2003, 94, 231–239. [Google Scholar] [CrossRef]
  13. Ouyang, J.; Zhou, L.; Liu, Z.; Heng, J.Y.Y.; Chen, W. Biomass-derived activated carbons for the removal of pharmaceutical mircopollutants from wastewater: A review. Sep. Purif. Technol. 2020, 253, 117536. [Google Scholar] [CrossRef]
  14. Jahirul, M.I.; Rasul, M.G.; Chowdhury, A.A.; Ashwath, N. Biofuels Production through Biomass Pyrolysis—A Technological Review. Energies 2012, 5, 4952–5001. [Google Scholar] [CrossRef]
  15. Duan, D.; Chen, D.; Huang, L.; Zhang, Y.; Zhang, Y.; Wang, Q.; Xiao, G.; Zhang, W.; Lei, H.; Ruan, R. Activated carbon from lignocellulosic biomass as catalyst: A review of the applications in fast pyrolysis process. J. Anal. Appl. Pyrolysis 2021, 158, 105246. [Google Scholar] [CrossRef]
  16. Lewoyehu, M. Comprehensive review on synthesis and application of activated carbon from agricultural residues for the remediation of venomous pollutants in wastewater. J. Anal. Appl. Pyrolysis 2021, 159, 105279. [Google Scholar] [CrossRef]
  17. You, F.-T.; Yu, G.-W.; Xing, Z.-J.; Li, J.; Xie, S.-Y.; Li, C.-X.; Wang, G.; Ren, H.-Y.; Wang, Y. Enhancement of NO catalytic oxidation on activated carbon at room temperature by nitric acid hydrothermal treatment. Appl. Surf. Sci. 2019, 471, 633–644. [Google Scholar] [CrossRef]
  18. Wang, J.; Duan, X.; Gao, J.; Shen, Y.; Feng, X.; Yu, Z.; Tan, X.; Liu, S.; Wang, S. Roles of structure defect, oxygen groups and heteroatom doping on carbon in nonradical oxidation of water contaminants. Water Res. 2020, 185, 116244. [Google Scholar] [CrossRef]
  19. Lukaszewicz, J.P. Carbon-film-based humidity sensor containing sodium or potassium. Recovery effect. Sens. Actuators B Chem. 1999, 60, 184–190. [Google Scholar] [CrossRef]
  20. Frackowiak, E.; Béguin, F. Carbon materials for the electrochemical storage of energy in capacitors. Carbon 2001, 39, 937–950. [Google Scholar] [CrossRef]
  21. Rodríguez-reinoso, F. The role of carbon materials in heterogeneous catalysis. Carbon 1998, 36, 159–175. [Google Scholar] [CrossRef]
  22. Ania, C.O.; Parra, J.B.; Pis, J.J. Influence of oxygen-containing functional groups on active carbon adsorption of selected organic compounds. Fuel Process. Technol. 2002, 79, 265–271. [Google Scholar] [CrossRef]
  23. Sharma, R.; Jasrotia, K.; Singh, N.; Ghosh, P.; Srivastava, S.; Sharma, N.R.; Singh, J.; Kanwar, R.; Kumar, A. A Comprehensive Review on Hydrothermal Carbonization of Biomass and its Applications. Chem. Afr. 2020, 3, 1–19. [Google Scholar] [CrossRef] [Green Version]
  24. Wenzhong, S.; Zhijie, L.; Yihong, L. Surface Chemical Functional Groups Modification of Porous Carbon. Recent Pat. Chem. Eng. 2008, 1, 27–40. [Google Scholar] [CrossRef]
  25. Rehman, A.; Park, M.; Park, S.-J. Current Progress on the Surface Chemical Modification of Carbonaceous Materials. Coatings 2019, 9, 103. [Google Scholar] [CrossRef] [Green Version]
  26. Bhatnagar, A.; Hogland, W.; Marques, M.; Sillanpää, M. An overview of the modification methods of activated carbon for its water treatment applications. Chem. Eng. J. 2013, 219, 499–511. [Google Scholar] [CrossRef]
  27. Sun, H.; Cannon, F.S.; He, X. Effective removal of perfluorooctanoate from groundwater using quaternary nitrogen-grafted granular activated carbon. J. Water Process. Eng. 2020, 37, 101416. [Google Scholar] [CrossRef]
  28. Mines, P.D.; Uthuppu, B.; Thirion, D.; Jakobsen, M.H.; Yavuz, C.T.; Andersen, H.R.; Hwang, Y. Granular activated carbon with grafted nanoporous polymer enhances nanoscale zero-valent iron impregnation and water contaminant removal. Chem. Eng. J. 2018, 339, 22–31. [Google Scholar] [CrossRef] [Green Version]
  29. Sato, S.; Yoshihara, K.; Moriyama, K.; Machida, M.; Tatsumoto, H. Influence of activated carbon surface acidity on adsorption of heavy metal ions and aromatics from aqueous solution. Appl. Surf. Sci. 2007, 253, 8554–8559. [Google Scholar] [CrossRef]
  30. Snoeyink, V.L.; Weber, W.J. The surface chemistry of active carbon; a discussion of structure and surface functional groups. Environ. Sci. Technol. 1967, 1, 228–234. [Google Scholar] [CrossRef]
  31. Makó, É.; Senkár, Z.; Kristóf, J.; Vágvölgyi, V. Surface modification of mechanochemically activated kaolinites by selective leaching. J. Colloid Interface Sci. 2006, 294, 362–370. [Google Scholar] [CrossRef] [PubMed]
  32. Ibrahim, M.A.; Saleh, T.A. Partially aminated acrylic acid grafted activated carbon as inexpensive shale hydration inhibitor. Carbohydr. Res. 2020, 491, 107960. [Google Scholar] [CrossRef]
  33. Chen, Y.; Zi, F.; Hu, X.; Yu, H.; Nie, Y.; Yang, P.; Cheng, H.; Wang, Q.; Qin, X.; Chen, S.; et al. Grafting of organic sulfur-containing functional groups on activated carbon for gold(I) adsorption from thiosulfate solution. Hydrometallurgy 2019, 185, 102–110. [Google Scholar] [CrossRef]
  34. Ribeiro, E.; Plantard, G.; Goetz, V. TiO2 grafted activated carbon elaboration by milling: Composition effect on sorption and photocatalytic properties. J. Photochem. Photobiol. A Chem. 2021, 408, 113108. [Google Scholar] [CrossRef]
  35. Shrestha, P.; Jha, M.K.; Ghimire, J.; Koirala, A.R.; Shrestha, R.M.; Sharma, R.K.; Pant, B.; Park, M.; Pant, H.R. Decoration of Zinc Oxide Nanorods into the Surface of Activated Carbon Obtained from Agricultural Waste for Effective Removal of Methylene Blue Dye. Materials 2020, 13, 5667. [Google Scholar] [CrossRef]
  36. Reza, M.S.; Yun, C.S.; Afroze, S.; Radenahmad, N.; Bakar, M.S.A.; Saidur, R.; Taweekun, J.; Azad, A.K. Preparation of activated carbon from biomass and its’ applications in water and gas purification, a review. Arab. J. Basic Appl. Sci. 2020, 27, 208–238. [Google Scholar] [CrossRef]
  37. Yu, L.; Gamliel, D.P.; Markunas, B.; Valla, J.A. A Promising Solution for Food Waste: Preparing Activated Carbons for Phenol Removal from Water Streams. ACS Omega 2021, 6, 8870–8883. [Google Scholar] [CrossRef]
  38. Bai, X.; Wang, Z.; Luo, J.; Wu, W.; Liang, Y.; Tong, X.; Zhao, Z. Hierarchical Porous Carbon with Interconnected Ordered Pores from Biowaste for High-Performance Supercapacitor Electrodes. Nanoscale Res. Lett. 2020, 15, 88. [Google Scholar] [CrossRef] [Green Version]
  39. Radenahmad, N.; Azad, A.T.; Saghir, M.; Taweekun, J.; Bakar, M.S.A.; Reza, M.S.; Azad, A.K. A review on biomass derived syngas for SOFC based combined heat and power application. Renew. Sustain. Energy Rev. 2020, 119, 109560. [Google Scholar] [CrossRef]
  40. Lateef, F.A.; Ogunsuyi, H.O. Jatropha curcas L. biomass transformation via torrefaction: Surface chemical groups and morphological characterization. Curr. Res. Green Sustain. Chem. 2021, 4, 100142. [Google Scholar] [CrossRef]
  41. Promraksa, A.; Rakmak, N. Biochar production from palm oil mill residues and application of the biochar to adsorb carbon dioxide. Heliyon 2020, 6, e04019. [Google Scholar] [CrossRef] [PubMed]
  42. Reza, M.S.; Islam, S.N.; Afroze, S.; Abu Bakar, M.S.; Sukri, R.S.; Rahman, S.; Azad, A.K. Evaluation of the bioenergy potential of invasive Pennisetum purpureum through pyrolysis and thermogravimetric analysis. Energy Ecol. Environ. 2020, 5, 118–133. [Google Scholar] [CrossRef]
  43. Negara, D.N.K.P.; Nindhia, T.G.T.; Surata, I.W.; Hidajat, F.; Sucipta, M. Nanopore structures, surface morphology, and adsorption capacity of tabah bamboo-activated carbons. Surf. Interfaces 2019, 16, 22–28. [Google Scholar] [CrossRef]
  44. Usman, M.; Chen, H.; Chen, K.; Ren, S.; Clark, J.H.; Fan, J.; Luo, G.; Zhang, S. Characterization and utilization of aqueous products from hydrothermal conversion of biomass for bio-oil and hydro-char production: A review. Green Chem. 2019, 21, 1553–1572. [Google Scholar] [CrossRef]
  45. Putra, H.E.; Damanhuri, E.; Dewi, K.; Pasek, A.D. Hydrothermal treatment of municipal solid waste into coal-like fuel. IOP Conf. Ser. Earth Environ. Sci. 2020, 483, 012021. [Google Scholar] [CrossRef]
  46. Maniscalco, M.P.; Volpe, M.; Messineo, A. Hydrothermal Carbonization as a Valuable Tool for Energy and Environmental Applications: A Review. Energies 2020, 13, 4098. [Google Scholar] [CrossRef]
  47. Kumar, N.; Weldon, R.; Lynam, J.G. Hydrothermal carbonization of coffee silverskins. Biocatal. Agric. Biotechnol. 2021, 36, 102145. [Google Scholar] [CrossRef]
  48. Zhao, Y.-Q.; Lu, M.; Tao, P.-Y.; Zhang, Y.-J.; Gong, X.-T.; Yang, Z.; Zhang, G.-Q.; Li, H.-L. Hierarchically porous and heteroatom doped carbon derived from tobacco rods for supercapacitors. J. Power Sources 2016, 307, 391–400. [Google Scholar] [CrossRef]
  49. 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] [CrossRef]
  50. Temuujin, J.; Jadambaa, T.; Burmaa, G.; Erdenechimeg, S.; Amarsanaa, J.; MacKenzie, K.J.D. Characterisation of acid activated montmorillonite clay from Tuulant (Mongolia). Ceram. Int. 2004, 30, 251–255. [Google Scholar] [CrossRef]
  51. Tamon, H.; Okazaki, M. Influence of acidic surface oxides of activated carbon on gas adsorption characteristics. Carbon 1996, 34, 741–746. [Google Scholar] [CrossRef]
  52. 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]
  53. Aggarwal, D.; Goyal, M.; Bansal, R.C. Adsorption of chromium by activated carbon from aqueous solution. Carbon 1999, 37, 1989–1997. [Google Scholar] [CrossRef]
  54. Park, S.-J.; Jang, Y.-S. Pore Structure and Surface Properties of Chemically Modified Activated Carbons for Adsorption Mechanism and Rate of Cr(VI). J. Colloid Interface Sci. 2002, 249, 458–463. [Google Scholar] [CrossRef]
  55. Jia, Y.F.; Thomas, K.M. Adsorption of Cadmium Ions on Oxygen Surface Sites in Activated Carbon. Langmuir 2000, 16, 1114–1122. [Google Scholar] [CrossRef]
  56. Gao, L.; Wen, H.; Tian, Q.; Wang, Y.; Li, G. Influence of surface modification by sulfuric acid on coking coal's adsorption of coking wastewater. Water Sci. Technol. 2017, 76, 555–566. [Google Scholar] [CrossRef]
  57. Strelko, V.; Malik, D.J. Characterization and Metal Sorptive Properties of Oxidized Active Carbon. J. Colloid Interface Sci. 2002, 250, 213–220. [Google Scholar] [CrossRef] [PubMed]
  58. Mavioglu Ayan, E.; Secim, P.; Karakaya, S.; Yanik, J. Oreganum Stalks as a New Biosorbent to Remove Textile Dyes from Aqueous Solutions. CLEAN—Soil Air Water 2012, 40, 856–863. [Google Scholar] [CrossRef]
  59. Carvajal-Bernal, A.M.; Gómez, F.; Giraldo, L.; Moreno-Pirajá, J.C. Chemical modification of activated carbons and its effect on the adsorption of phenolic compounds. Ing. Y Compet. 2015, 17, 109–119. [Google Scholar]
  60. Pradhananga, R.R.; Adhikari, L.; Shrestha, R.G.; Adhikari, M.P.; Rajbhandari, R.; Ariga, K.; Shrestha, L.K. Wool Carpet Dye Adsorption on Nanoporous Carbon Materials Derived from Agro-Product. C 2017, 3, 12. [Google Scholar] [CrossRef]
  61. Shrestha, R.M.; Varga, I.; Bajtai, J.; Varga, M. Design of surface functionalization of waste material originated charcoals by an optimized chemical carbonization for the purpose of heavy metal removal from industrial waste waters. Microchem. J. 2013, 108, 224–232. [Google Scholar] [CrossRef]
  62. Martín-Lara, M.A.; Pagnanelli, F.; Mainelli, S.; Calero, M.; Toro, L. Chemical treatment of olive pomace: Effect on acid-basic properties and metal biosorption capacity. J. Hazard. Mater. 2008, 156, 448–457. [Google Scholar] [CrossRef] [PubMed]
  63. Amuda, O.S.; Giwa, A.A.; Bello, I.A. Removal of heavy metal from industrial wastewater using modified activated coconut shell carbon. Biochem. Eng. J. 2007, 36, 174–181. [Google Scholar] [CrossRef]
  64. Elangovan, R.; Philip, L.; Chandraraj, K. Biosorption of chromium species by aquatic weeds: Kinetics and mechanism studies. J. Hazard. Mater. 2008, 152, 100–112. [Google Scholar] [CrossRef]
  65. Mameri, N.; Aioueche, F.; Belhocine, D.; Grib, H.; Lounici, H.; Piron, D.L.; Yahiat, Y. Preparation of activated carbon from olive mill solid residue. J. Chem. Technol. Biotechnol. 2000, 75, 625–631. [Google Scholar] [CrossRef]
  66. Blázquez, G.; Calero, M.; Ronda, A.; Tenorio, G.; Martín-Lara, M.A. Study of kinetics in the biosorption of lead onto native and chemically treated olive stone. J. Ind. Eng. Chem. 2014, 20, 2754–2760. [Google Scholar] [CrossRef]
  67. Montanher, S.F.; Oliveira, E.A.; Rollemberg, M.C. Removal of metal ions from aqueous solutions by sorption onto rice bran. J. Hazard. Mater. 2005, 117, 207–211. [Google Scholar] [CrossRef]
  68. Ahn, C.K.; Kim, Y.M.; Woo, S.H.; Park, J.M. Removal of cadmium using acid-treated activated carbon in the presence of nonionic and/or anionic surfactants. Hydrometallurgy 2009, 99, 209–213. [Google Scholar] [CrossRef]
  69. Menéndez, J.A.; Phillips, J.; Xia, B.; Radovic, L.R. On the Modification and Characterization of Chemical Surface Properties of Activated Carbon:  In the Search of Carbons with Stable Basic Properties. Langmuir 1996, 12, 4404–4410. [Google Scholar] [CrossRef]
  70. Shaarani, F.W.; Hameed, B.H. Ammonia-modified activated carbon for the adsorption of 2,4-dichlorophenol. Chem. Eng. J. 2011, 169, 180–185. [Google Scholar] [CrossRef]
  71. Mangun, C.L.; Benak, K.R.; Economy, J.; Foster, K.L. Surface chemistry, pore sizes and adsorption properties of activated carbon fibers and precursors treated with ammonia. Carbon 2001, 39, 1809–1820. [Google Scholar] [CrossRef]
  72. Jansen, R.J.J.; van Bekkum, H. XPS of nitrogen-containing functional groups on activated carbon. Carbon 1995, 33, 1021–1027. [Google Scholar] [CrossRef]
  73. Jansen, R.J.J.; van Bekkum, H. Amination and ammoxidation of activated carbons. Carbon 1994, 32, 1507–1516. [Google Scholar] [CrossRef]
  74. Dastgheib, S.A.; Karanfil, T.; Cheng, W. Tailoring activated carbons for enhanced removal of natural organic matter from natural waters. Carbon 2004, 42, 547–557. [Google Scholar] [CrossRef]
  75. Chen, W.; Cannon, F.S.; Rangel-Mendez, J.R. Ammonia-tailoring of GAC to enhance perchlorate removal. I: Characterization of NH3 thermally tailored GACs. Carbon 2005, 43, 573–580. [Google Scholar] [CrossRef]
  76. Zheng, C.; Zhao, L.; Zhou, X.; Fu, Z.; Li, A. Treatment technologies for organic wastewater. Water Treat. 2013, 11, 250–286. [Google Scholar]
  77. Chiang, H.-L.; Huang, C.P.; Chiang, P.C. The surface characteristics of activated carbon as affected by ozone and alkaline treatment. Chemosphere 2002, 47, 257–265. [Google Scholar] [CrossRef]
  78. Przepiórski, J. Enhanced adsorption of phenol from water by ammonia-treated activated carbon. J. Hazard. Mater. 2006, 135, 453–456. [Google Scholar] [CrossRef]
  79. Lorenc-Grabowska, E.; Gryglewicz, G.; Machnikowski, J. p-Chlorophenol adsorption on activated carbons with basic surface properties. Appl. Surf. Sci. 2010, 256, 4480–4487. [Google Scholar] [CrossRef]
  80. Kasnejad, M.H.; Esfandiari, A.; Kaghazchi, T.; Asasian, N. Effect of pre-oxidation for introduction of nitrogen containing functional groups into the structure of activated carbons and its influence on Cu (II) adsorption. J. Taiwan Inst. Chem. Eng. 2012, 43, 736–740. [Google Scholar] [CrossRef]
  81. Ofudje, E.A.; Akiode, O.K.; Oladipo, G.O.; Adedapo, A.E.; Adebayo, L.O.; Awotula, A.O. Application of raw and alkaline-modified coconut shaft as a biosorbent for Pb2+ removal. BioResources 2015, 10, 3462–3480. [Google Scholar] [CrossRef] [Green Version]
  82. Joshi, S.; Shrestha, L.K.; Kamachi, Y.; Malgras, V.; Pradhananga, M.A.; Pokhrel, B.P.; Nakato, T.; Pradhananga, R.R.; Ariga, K.; Yamauchi, Y. Synthesis and characterizations of nanoporous carbon derived from Lapsi (Choerospondias axillaris) seed: Effect of carbonization conditions. Adv. Powder Technol. 2015, 26, 894–900. [Google Scholar] [CrossRef] [Green Version]
  83. Plaza, M.G.; Pevida, C.; Arias, B.; Casal, M.D.; Martín, C.F.; Fermoso, J.; Rubiera, F.; Pis, J.J. Different Approaches for the Development of Low-Cost CO2 Adsorbents. J. Environ. Eng. 2009, 135, 426–432. [Google Scholar] [CrossRef] [Green Version]
  84. Radovic, L.R.; Silva, I.F.; Ume, J.I.; Menéndez, J.A.; Leon, C.A.L.Y.; Scaroni, A.W. An experimental and theoretical study of the adsorption of aromatics possessing electron-withdrawing and electron-donating functional groups by chemically modified activated carbons. Carbon 1997, 35, 1339–1348. [Google Scholar] [CrossRef]
  85. Abotsi, G.M.K.; Scaroni, A.W. Reaction of carbons with ammonia: Effects on the surface charge and molybdenum adsorption. Carbon 1990, 28, 79–84. [Google Scholar] [CrossRef]
  86. Plaza, M.G.; Pevida, C.; Arias, B.; Fermoso, J.; Rubiera, F.; Pis, J.J. A comparison of two methods for producing CO2 capture adsorbents. Energy Procedia 2009, 1, 1107–1113. [Google Scholar] [CrossRef] [Green Version]
  87. Stöhr, B.; Boehm, H.P.; Schlögl, R. Enhancement of the catalytic activity of activated carbons in oxidation reactions by thermal treatment with ammonia or hydrogen cyanide and observation of a superoxide species as a possible intermediate. Carbon 1991, 29, 707–720. [Google Scholar] [CrossRef]
  88. Przepiórski, J.; Skrodzewicz, M.; Morawski, A.W. High temperature ammonia treatment of activated carbon for enhancement of CO2 adsorption. Appl. Surf. Sci. 2004, 225, 235–242. [Google Scholar] [CrossRef]
  89. Pant, H.R.; Adhikari, S.P.; Pant, B.; Joshi, M.K.; Kim, H.J.; Park, C.H.; Kim, C.S. Immobilization of TiO2 nanofibers on reduced graphene sheets: Novel strategy in electrospinning. J. Colloid Interface Sci. 2015, 457, 174–179. [Google Scholar] [CrossRef]
  90. Henning, K.D.; Schäfer, S. Impregnated activated carbon for environmental protection. Gas Sep. Purif. 1993, 7, 235–240. [Google Scholar] [CrossRef]
  91. Adhoum, N.; Monser, L. Removal of phthalate on modified activated carbon: Application to the treatment of industrial wastewater. Sep. Purif. Technol. 2004, 38, 233–239. [Google Scholar] [CrossRef]
  92. Miyanaga, S.; Hiwara, A.; Yasuda, H. Preparation and high bacteriostatic action of the activated carbons possessing ultrafine silver particles. Sci. Technol. Adv. Mater. 2002, 3, 103–109. [Google Scholar] [CrossRef]
  93. Yeddou, A.R.; Chergui, S.; Chergui, A.; Halet, F.; Hamza, A.; Nadjemi, B.; Ould-Dris, A.; Belkouch, J. Removal of cyanide in aqueous solution by oxidation with hydrogen peroxide in presence of copper-impregnated activated carbon. Miner. Eng. 2011, 24, 788–793. [Google Scholar] [CrossRef]
  94. Tchomgui-Kamga, E.; Alonzo, V.; Nanseu-Njiki, C.P.; Audebrand, N.; Ngameni, E.; Darchen, A. Preparation and characterization of charcoals that contain dispersed aluminum oxide as adsorbents for removal of fluoride from drinking water. Carbon 2010, 48, 333–343. [Google Scholar] [CrossRef]
  95. Vaughan, R.L.; Reed, B.E. Modeling As(V) removal by a iron oxide impregnated activated carbon using the surface complexation approach. Water Res. 2005, 39, 1005–1014. [Google Scholar] [CrossRef]
  96. Girish, C.R. Various Impregnation Methods Used for the Surface Modification of the Adsorbent: A Review. Int. J. Eng. Technol. 2018, 7, 330–334. [Google Scholar] [CrossRef] [Green Version]
  97. Huang, C.P.; Vane, L.M. Enhancing As5+ Removal by a Fe2+-Treated Activated Carbon. Res. J. Water Pollut. Control. Fed. 1989, 61, 1596–1603. [Google Scholar]
  98. Chang, Q.; Lin, W.; Ying, W.-C. Preparation of iron-impregnated granular activated carbon for arsenic removal from drinking water. J. Hazard. Mater. 2010, 184, 515–522. [Google Scholar] [CrossRef]
  99. Vitela-Rodriguez, A.V.; Rangel-Mendez, J.R. Arsenic removal by modified activated carbons with iron hydro(oxide) nanoparticles. J. Environ. Manag. 2013, 114, 225–231. [Google Scholar] [CrossRef]
  100. Adhoum, N.; Monser, L. Removal of cyanide from aqueous solution using impregnated activated carbon. Chem. Eng. Process. Process. Intensif. 2002, 41, 17–21. [Google Scholar] [CrossRef]
  101. Shah, I.; Adnan, R.; Wan Ngah, W.S.; Mohamed, N. Iron impregnated activated carbon as an efficient adsorbent for the removal of methylene blue: Regeneration and kinetics studies. PLoS ONE 2015, 10, e0122603. [Google Scholar] [CrossRef] [Green Version]
  102. Sonal, S.; Prakash, P.; Mishra, B.K.; Nayak, G.C. Synthesis, characterization and sorption studies of a zirconium(iv) impregnated highly functionalized mesoporous activated carbons. RSC Adv. 2020, 10, 13783–13798. [Google Scholar] [CrossRef] [Green Version]
  103. Lee, S.J.; Seo, Y.-C.; Jurng, J.; Lee, T.G. Removal of gas-phase elemental mercury by iodine- and chlorine-impregnated activated carbons. Atmos. Environ. 2004, 38, 4887–4893. [Google Scholar] [CrossRef]
  104. Korpiel, J.A.; Vidic, R.D. Effect of Sulfur Impregnation Method on Activated Carbon Uptake of Gas-Phase Mercury. Environ. Sci. Technol. 1997, 31, 2319–2325. [Google Scholar] [CrossRef]
  105. Rajaković, L.V. The Sorption of Arsenic onto Activated Carbon Impregnated with Metallic Silver and Copper. Sep. Sci. Technol. 1992, 27, 1423–1433. [Google Scholar] [CrossRef]
  106. Deveci, H.; Yazıcı, E.Y.; Alp, I.; Uslu, T. Removal of cyanide from aqueous solutions by plain and metal-impregnated granular activated carbons. Int. J. Miner. Process. 2006, 79, 198–208. [Google Scholar] [CrossRef]
  107. Sutherland, I.; Sheng, E.; Bradley, R.; Freakley, P. Effects of ozone oxidation on carbon black surfaces. J. Mater. Sci. 1996, 31, 5651–5655. [Google Scholar] [CrossRef]
  108. Zaror, C.; Soto, G.; Valdés, H.; Mansilla, H. Ozonation of 1, 2-dihydroxybenzene in the presence of activated carbon. Water Sci. Technol. 2001, 44, 125–130. [Google Scholar] [CrossRef]
  109. Zaror, C.A. Enhanced oxidation of toxic effluents using simultaneous ozonation and activated carbon treatment. J. Chem. Technol. Biotechnol. 1997, 70, 21–28. [Google Scholar] [CrossRef]
  110. Jans, U.; Hoigné, J. Activated Carbon and Carbon Black Catalyzed Transformation of Aqueous Ozone into OH-Radicals. Ozone Sci. Eng. 1998, 20, 67–90. [Google Scholar] [CrossRef]
  111. Rivera-Utrilla, J.; Sánchez-Polo, M. Ozonation of 1,3,6-naphthalenetrisulphonic acid catalysed by activated carbon in aqueous phase. Appl. Catal. B Environ. 2002, 39, 319–329. [Google Scholar] [CrossRef] [Green Version]
  112. Álvarez, P.M.; García-Araya, J.F.; Beltrán, F.J.; Masa, F.J.; Medina, F. Ozonation of activated carbons: Effect on the adsorption of selected phenolic compounds from aqueous solutions. J. Colloid Interface Sci. 2005, 283, 503–512. [Google Scholar] [CrossRef]
  113. Jaramillo, J.; Álvarez, P.M.; Gómez-Serrano, V. Preparation and ozone-surface modification of activated carbon. Thermal stability of oxygen surface groups. Appl. Surf. Sci. 2010, 256, 5232–5236. [Google Scholar] [CrossRef]
  114. Valdés, H.; Sánchez-Polo, M.; Rivera-Utrilla, J.; Zaror, C.A. Effect of Ozone Treatment on Surface Properties of Activated Carbon. Langmuir 2002, 18, 2111–2116. [Google Scholar] [CrossRef]
  115. Zhang, J.; Huang, G.-Q.; Liu, C.; Zhang, R.-N.; Chen, X.-X.; Zhang, L. Synergistic effect of microbubbles and activated carbon on the ozonation treatment of synthetic dyeing wastewater. Sep. Purif. Technol. 2018, 201, 10–18. [Google Scholar] [CrossRef]
  116. Joshi, A.A.; Locke, B.R.; Arce, P.; Finney, W.C. Formation of hydroxyl radicals, hydrogen peroxide and aqueous electrons by pulsed streamer corona discharge in aqueous solution. J. Hazard. Mater. 1995, 41, 3–30. [Google Scholar] [CrossRef]
  117. Locke, B.R.; Sato, M.; Sunka, P.; Hoffmann, M.R.; Chang, J.S. Electrohydraulic Discharge and Nonthermal Plasma for Water Treatment. Ind. Eng. Chem. Res. 2006, 45, 882–905. [Google Scholar] [CrossRef]
  118. Bogaerts, A.; Neyts, E.; Gijbels, R.; van der Mullen, J. Gas discharge plasmas and their applications. Spectrochim. Acta Part B At. Spectrosc. 2002, 57, 609–658. [Google Scholar] [CrossRef]
  119. Panorel, I.; Preis, S.; Kornev, I.; Hatakka, H.; Louhi-Kultanen, M. Oxidation of aqueous pharmaceuticals by pulsed corona discharge. Environ. Technol. 2013, 34, 923–930. [Google Scholar] [CrossRef]
  120. Zhang, W.; Liu, H.; Xia, Q.; Li, Z. Enhancement of dibenzothiophene adsorption on activated carbons by surface modification using low temperature oxygen plasma. Chem. Eng. J. 2012, 209, 597–600. [Google Scholar] [CrossRef]
  121. ChangMing, D.; DongWei, H.; HongXia, L.; MuDan, X.; Kui, W.; Lu, Z.; ZhiYi, L.; TengFei, C.; JianMin, M.; Dong, G.; et al. Adsorption of Acid Orange II from Aqueous Solution by Plasma Modified Activated Carbon Fibers. Plasma Chem. Plasma Process. 2013, 33, 65–82. [Google Scholar] [CrossRef]
  122. Kee, S.H.; Chiongson, J.B.V.; Saludes, J.P.; Vigneswari, S.; Ramakrishna, S.; Bhubalan, K. Bioconversion of agro-industry sourced biowaste into biomaterials via microbial factories—A viable domain of circular economy. Environ. Pollut. 2021, 271, 116311. [Google Scholar] [CrossRef]
  123. Wilcox, D.; Chang, E.; Dickson, K.; Johansson, K. Microbial growth associated with granular activated carbon in a pilot water treatment facility. Appl. Environ. Microbiol. 1983, 46, 406–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Rivera-Utrilla, J.; Bautista-Toledo, I.; Ferro-García, M.A.; Moreno-Castilla, C. Activated carbon surface modifications by adsorption of bacteria and their effect on aqueous lead adsorption. J. Chem. Technol. Biotechnol. 2001, 76, 1209–1215. [Google Scholar] [CrossRef] [Green Version]
  125. Jiang, C.; Cui, S.; Han, Q.; Li, P.; Zhang, Q.; Song, J.; Li, M. Study on Application of Activated Carbon in Water Treatment. IOP Conf. Ser. Earth Environ. Sci. 2019, 237, 022049. [Google Scholar] [CrossRef]
  126. Gutiérrez, M.; Grillini, V.; Mutavdžić Pavlović, D.; Verlicchi, P. Activated carbon coupled with advanced biological wastewater treatment: A review of the enhancement in micropollutant removal. Sci. Total Environ. 2021, 790, 148050. [Google Scholar] [CrossRef]
  127. Carrott, P.J.M.; Nabais, J.M.V.; Ribeiro Carrott, M.M.L.; Menéndez, J.A. Thermal treatments of activated carbon fibres using a microwave furnace. Microporous Mesoporous Mater. 2001, 47, 243–252. [Google Scholar] [CrossRef] [Green Version]
  128. Foo, K.Y.; Hameed, B.H. Microwave-assisted regeneration of activated carbon. Bioresour. Technol. 2012, 119, 234–240. [Google Scholar] [CrossRef]
  129. Yagmur, E.; Ozmak, M.; Aktas, Z. A novel method for production of activated carbon from waste tea by chemical activation with microwave energy. Fuel 2008, 87, 3278–3285. [Google Scholar] [CrossRef]
  130. Liu, Q.-S.; Zheng, T.; Li, N.; Wang, P.; Abulikemu, G. Modification of bamboo-based activated carbon using microwave radiation and its effects on the adsorption of methylene blue. Appl. Surf. Sci. 2010, 256, 3309–3315. [Google Scholar] [CrossRef]
  131. Chen, C.J.; Wei, L.B.; Zhao, P.C.; Li, Y.; Hu, H.Y.; Qin, Y.B. Study on preparation of activated carbon from corncob furfural residue with ZnCl2 by microwave irradiation. Adv. Mater. Res. 2010, 152–153, 1322–1327. [Google Scholar] [CrossRef]
  132. Abdulrasheed, A.A.; Jalil, A.A.; Triwahyono, S.; Zaini, M.A.A.; Gambo, Y.; Ibrahim, M. Surface modification of activated carbon for adsorption of SO2 and NOX: A review of existing and emerging technologies. Renew. Sustain. Energy Rev. 2018, 94, 1067–1085. [Google Scholar] [CrossRef]
  133. Naushad, M.; Alqadami, A.A.; Ahamad, T. Removal of Cd(II) ion from aqueous environment using triaminotriethoxysilane grafted oxidized activated carbon synthesized via activation and subsequent silanization. Environ. Technol. Innov. 2020, 18, 100686. [Google Scholar] [CrossRef]
  134. Li, Z.; Guo, D.; Liu, Y.; Wang, H.; Wang, L. Recent advances and challenges in biomass-derived porous carbon nanomaterials for supercapacitors. Chem. Eng. J. 2020, 397, 125418. [Google Scholar] [CrossRef]
  135. Rivera-Utrilla, J.; Sánchez-Polo, M.; Ferro-García, M.Á.; Prados-Joya, G.; Ocampo-Pérez, R. Pharmaceuticals as emerging contaminants and their removal from water. A review. Chemosphere 2013, 93, 1268–1287. [Google Scholar] [CrossRef]
  136. Ruiz-Rosas, R.; García-Mateos, F.J.; Gutiérrez, M.D.C.; Rodríguez-Mirasol, J.; Cordero, T. About the Role of Porosity and Surface Chemistry of Phosphorus-Containing Activated Carbons in the Removal of Micropollutants. Front. Mater. 2019, 6, 134. [Google Scholar] [CrossRef] [Green Version]
  137. Joshi, S.; Sharma, M.; Kumari, A.; Shrestha, S.; Shrestha, B. Arsenic Removal from Water by Adsorption onto Iron Oxide/Nano-Porous Carbon Magnetic Composite. Appl. Sci. 2019, 9, 3732. [Google Scholar] [CrossRef] [Green Version]
  138. Sodeinde, O. Preparation of a locally produced activated carbon from coconut shells and its use in reducing hexamine cobalt (III). Int. J. Chem. Eng. Appl. 2012, 3, 67. [Google Scholar] [CrossRef]
  139. Latiff, M.F.P.M.; Abustan, I.; Ahmad, M.A.; Yahaya, N.K.E.M.; Khalid, A.M. Effect of preparation conditions of activated carbon prepared from corncob by CO2 activation for removal of Cu (II) from aqueous solution. AIP Conf. Proc. 2016, 1774, 030001. [Google Scholar] [CrossRef] [Green Version]
  140. Masiya, T.; Gudyanga, F. Investigation of Granular Activated Carbon From Peach Stones for Gold Adsorption in Acidic Thiourea. In Proceedings of the Hydrometallurgy Conference 2009, Muldersdrift, South Africa, 24–26 February 2009; pp. 465–474. [Google Scholar]
  141. Nahil, M.A.; Williams, P.T. Pore characteristics of activated carbons from the phosphoric acid chemical activation of cotton stalks. Biomass Bioenergy 2012, 37, 142–149. [Google Scholar] [CrossRef]
  142. Singh, P.; Raizada, P.; Pathania, D.; Sharma, G.; Sharma, P. Microwave induced KOH activation of guava peel carbon as an adsorbent for congo red dye removal from aqueous phase. IJCT 2013, 20, 305–311. [Google Scholar]
  143. Arami-Niya, A.; Daud, W.M.A.W.; Mjalli, F.S. Comparative study of the textural characteristics of oil palm shell activated carbon produced by chemical and physical activation for methane adsorption. Chem. Eng. Res. Des. 2011, 89, 657–664. [Google Scholar] [CrossRef]
  144. Hassan, L.; Itodo, A.; Umar, U.B. Equilibrium study on the biosorption of malachite green from aqueous solution onto thermochemically cracked groundnut shells. J. Chem. Pharm. Res. 2010, 2, 656–666. [Google Scholar]
  145. Hasar, H. Adsorption of nickel(II) from aqueous solution onto activated carbon prepared from almond husk. J. Hazard. Mater. 2003, 97, 49–57. [Google Scholar] [CrossRef]
  146. Vicinisvarri, I.; Kumar, S.S.; Aimi, N.; Norain, I.; Izza, N. Preparation and characterization of phosphoric acid activated carbon from Canarium Odontophyllum (Dabai) nutshell for methylene blue adsorption. Res. J. Chem. Environ. 2014, 18, 57–62. [Google Scholar]
  147. Hameed, B.H.; Din, A.T.M.; Ahmad, A.L. Adsorption of methylene blue onto bamboo-based activated carbon: Kinetics and equilibrium studies. J. Hazard. Mater. 2007, 141, 819–825. [Google Scholar] [CrossRef]
Figure 1. (A) Schematic diagram representing the three types of pores for pollutant adsorption [16]. Reprinted with permission from the Journal of Analytical and Applied Pyrolysis, Copyright Elsevier 2021. (B) TEM image of AC showing the hierarchical pore distribution [17]. Reprinted with permission from Applied Surface Science, Copyright Elsevier 2019.
Figure 1. (A) Schematic diagram representing the three types of pores for pollutant adsorption [16]. Reprinted with permission from the Journal of Analytical and Applied Pyrolysis, Copyright Elsevier 2021. (B) TEM image of AC showing the hierarchical pore distribution [17]. Reprinted with permission from Applied Surface Science, Copyright Elsevier 2019.
Nanomaterials 11 03140 g001
Figure 3. Illustration of formation of ACs using pyrolysis carbonization followed by chemical activation ((A) tabah bamboo; (B) pieces of tabah bamboo; (C) char; (D) powdered char; and (E) activated carbon) and SEM images obtained from bamboo (F,G) [43]. Reprinted with permission from Surfaces and Interfaces, Copyright Elsevier 2019.
Figure 3. Illustration of formation of ACs using pyrolysis carbonization followed by chemical activation ((A) tabah bamboo; (B) pieces of tabah bamboo; (C) char; (D) powdered char; and (E) activated carbon) and SEM images obtained from bamboo (F,G) [43]. Reprinted with permission from Surfaces and Interfaces, Copyright Elsevier 2019.
Nanomaterials 11 03140 g003
Figure 4. Schematic diagram showing the use of hydrothermal process for the preparation of ZnO/AC composite [35].
Figure 4. Schematic diagram showing the use of hydrothermal process for the preparation of ZnO/AC composite [35].
Nanomaterials 11 03140 g004
Table 1. The biochemical composition of some agriculture lignocellulosic biomasses [14].
Table 1. The biochemical composition of some agriculture lignocellulosic biomasses [14].
Feedstock Lignin (%)Cellulose (%)Hemicellulose (%)
Almond shell20.450.728.9
Bamboo21.3126–4315–26
Banana waste1413.214.8
Barley straw14–1431–3424–29
Rice straw1832.124
Sugarcane bagasse23–3219–2432–48
Tea waste4030.219.9
Walnut shell52.325.622.7
Wheat straw15–2033–4020–30
Wood25–3035–4020–30
Table 2. Some reports on the alteration of AC with acid treatment.
Table 2. Some reports on the alteration of AC with acid treatment.
SamplesAcid UsedSpecies BiosorbedRef.
Agricultural wasteH3PO4, H2O2Cd[62]
Activated coconut shell carbonH3PO4Zn(II)[63]
Aquatic weedsH2SO4Cr(III), Cr(VI)[64]
Olive mill solid residueHClPhenol[65]
Olive stoneH2SO4, HNO3Pb (II)[66]
Rice branHNO3Cd(II), Cu(II), Pb(II), Zn(II)[67]
Table 3. A review of studies regarding the use of gaseous ammonia to modify activated carbon.
Table 3. A review of studies regarding the use of gaseous ammonia to modify activated carbon.
MaterialsAmination TemperatureApplicationsRef.
Carbon materials (biomass residues, sewage, sludge, pet, coke)400 °CCO2 adsorption[83]
Commercial granular activated carbons385 °CAdsorption of aromatic compounds (aniline, nitrobenzene)[84]
Activated carbon from sulfonated styrene-divinyl-benzene copolymer600 °CAdsorption of molybdenum[85]
Carbon adsorbents from biomass residue (almond shells)800 °CCO2 adsorption[86]
Activated carbon from pea900 °CEnhancement of catalytic
activity of AC in
oxidation reaction
[87]
Commercial activated carbon1000 °CCO2 adsorption[88]
Table 4. Chemically impregnated ACs and their potential applications.
Table 4. Chemically impregnated ACs and their potential applications.
SamplesSpecies ImpregnatedSpecies RemovedRef.
Activated carboniodine and chlorinegas-phase elemental mercury[103]
Granular activated carbonsulfurgas-phase elemental mercury[104]
Activated carbonmetallic silver and copperarsenic[105]
Activated carbonsilver and nickelcyanide[100]
Granular activated carboncopper and silvercyanide[106]
Table 5. Applications of activated carbon from biomass.
Table 5. Applications of activated carbon from biomass.
Biomass PrecursorApplicationRef.
Coconut shellReduction of hexamine cobalt (III)[138]
Rice huskAdsorption of Cu[139]
Peach stoneAdsorption of gold[140]
Sugarcane bagasseDecolorization of sugar[141]
Guava peelRemoval of Congo Red dye[142]
Oil palm shellMethane adsorption[143]
Ground nut shellAdsorption of malachite green[144]
Almond huskAdsorption of Ni (II) from aqueous solution[145]
Nut shellMethylene blue adsorption[146]
BambooAdsorption of methylene blue[147]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jha, M.K.; Joshi, S.; Sharma, R.K.; Kim, A.A.; Pant, B.; Park, M.; Pant, H.R. Surface Modified Activated Carbons: Sustainable Bio-Based Materials for Environmental Remediation. Nanomaterials 2021, 11, 3140. https://doi.org/10.3390/nano11113140

AMA Style

Jha MK, Joshi S, Sharma RK, Kim AA, Pant B, Park M, Pant HR. Surface Modified Activated Carbons: Sustainable Bio-Based Materials for Environmental Remediation. Nanomaterials. 2021; 11(11):3140. https://doi.org/10.3390/nano11113140

Chicago/Turabian Style

Jha, Manoj Kumar, Sahira Joshi, Ram Kumar Sharma, Allison A Kim, Bishweshwar Pant, Mira Park, and Hem Raj Pant. 2021. "Surface Modified Activated Carbons: Sustainable Bio-Based Materials for Environmental Remediation" Nanomaterials 11, no. 11: 3140. https://doi.org/10.3390/nano11113140

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