Next Article in Journal
Food Systems in Informal Urban Settlements—Exploring Differences in Livelihood Welfare Factors across Kibera, Nairobi
Next Article in Special Issue
Utilization of Colored Extracts for the Formulation of Ecological Friendly Plant-Based Green Products
Previous Article in Journal
A Heuristic Algorithm Based on Travel Demand for Transit Network Design
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Sustainable Adsorbents from Plant-Derived Agricultural Wastes for Anionic Dye Removal: A Review

by
Abu Naser Md Ahsanul Haque
1,*,
Nigar Sultana
2,
Abu Sadat Muhammad Sayem
3 and
Shamima Akter Smriti
4
1
Institute for Frontier Materials, Deakin University, Waurn Ponds, VIC 3216, Australia
2
Department of Chemistry, Jagannath University, Dhaka 1100, Bangladesh
3
Manchester Fashion Institute, Manchester Metropolitan University, Manchester M15 6BG, UK
4
Department of Fabric Engineering, Bangladesh University of Textiles, Dhaka 1208, Bangladesh
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(17), 11098; https://doi.org/10.3390/su141711098
Submission received: 4 August 2022 / Revised: 31 August 2022 / Accepted: 31 August 2022 / Published: 5 September 2022
(This article belongs to the Special Issue Advances in Sustainable Valorization of Natural Waste and Biomass)

Abstract

:
The extensive use of dyes in numerous industries results in massive dye discharge in the wastewater, which is a major cause of water pollution. Globally, the consumption of dyes is near seven hundred thousand tons across different sectors, of which around 10–15% goes into the wastewater. Among the dye kinds, anionic dyes make up the main proportion, having a 32–90% share in the wastewater. Different plant-derived wastes, which are sustainable given their natural abundance, effectiveness, and low cost, are frequently proposed for dye separation. However, these adsorbents are inherently more suitable for cationic dyes than anionic dyes. In recent years, the modification of these wastes has been progressively considered to suit them to anionic dye removal. These modifications involve mechanical, thermal, or chemical treatments, or combinations. These attempts propose two-way benefits, as one abundant waste is being used to cure another severe problem, and eventually both could be diminished. This review has a key focus on the evaluation of plant-derived adsorbents and their modifications, and particularly for anionic dye adsorption. Overall, the mechanism of adsorption and the suitability of the current methods are discussed, and their future potential is explored.

1. Introduction

Dyes are comprehensively used in food, pharmaceutical, textile, plastic, metal, and many other manufacturing industries to impart the colors of choice to the end products. Currently, more than seven hundred thousand tons of dyes are required each year across different sectors, of which at least 10–15% are discarded into the wastewater [1]. Although the use of dyes is unavoidable, it is also responsible for massive water pollution, which is a major concern. Anionic dyes are the known main fraction (nearly 32–90%), compared with all other dye kinds (such as cationic or nonionic dye), in the wastewater. These include the acid, direct, reactive, and soluble forms of some nonionic dyes, which are difficult to eliminate due to their high solubility in water [2]. These dyes not only alter the color of water, but they also inhibit light penetration and reduce the rate of photosynthesis and the oxygen level, causing damage to the entire aquatic ecosystem [2]. Often, these dyes are carcinogenic and initiate various diseases in humans, such as the dysfunction of the kidneys, reproductive system, liver, brain, and central nervous system, and they bring hazards to other living organisms [3]. Therefore, it is essential to remove these dyes from wastewater to attain a sustainable environment. In the recent past, there has been a growing interest in the removal of anionic dyes from wastewater, and particularly in the last few years. For instance, a search of the keyword ‘anionic dye removal’ in the Web of Science database (title, abstract, keywords only) recalled 4981 articles starting from 1975 (this number is 3637 from Scopus): 3938 have been published from 2015 to the present (~79%), and 2457 have been published from just 2019 to the present date (~49%). These studies have been performed across varied disciplines, such as from the engineering, environmental, chemistry, and materials perspectives, which shows the urge to solve the problem from multiple sectors (Figure 1).
Many techniques are proposed for removing anionic dyes from wastewater, such as physical (membrane filtration, ion exchange, irradiation, coagulation), chemical (oxidation, ozonation, photochemical, electrochemical destruction), and biological (aerobic and anaerobic microbial degradation) treatments [4]. However, the adsorption method is still a preferred method of dye removal due to its simplicity and efficiency [5]. There is an ongoing interest in preparing more sustainable adsorbents from agricultural and other natural plant-derived wastes, as they are widely available at a low cost. These resources are chemically known as lignocellulose, and they are often collectible at a negligible cost. The composition of lignocellulose biomasses (such as wheat straw, rice straw, oat straw, bagasse, cotton gin trash, and so on) is mostly cellulose (40–50%), intimately associated with hemicellulose (20–30%) and lignin (10–25%) [6]. However, due to their chemical nature, they are more prone to adsorb cationic dyes than anionic dyes. This is because the lignocellulose structure is chemically composed of different anions as its active sites, such as the hydroxyl and carbonyl groups, and it generally shows a negative surface charge [7]. Thus, it is important to modify them to take advantage of their abundance and utilize them for anionic dye separation. Over the years, several modification methods of these wastes have been proposed, including combinations of mechanical, thermal, or chemical treatments, to enhance the adsorption ability of these resources.
Because the use of agricultural wastes for dye separation is a massive area of research and the interest is increasing, currently, there are several reviews available that summarize these findings. However, these are mostly based on all classes of dyes [4,8,9,10], rather than with any particular focus on the anionicdye class. Due to the natural attraction of cationic dyes to the lignocellulosic structure, the removal of cationic dyes has received more interest compared with the removal of anionic dyes, which limits the discussions for the anionic dye adsorption in reviews in which both classes are discussed. However, as mentioned above, anionic dyes are the major proportion of all other dyes, and the separation of these dyes needs specific attention. The adsorption of anionic dyes is essential, but more challenging, compared with cationic dyes. In recent years, numerous common lignocellulosic wastes, such as rice husk [11], wheat straw [12], sawdust [13], cotton waste [7], banana waste [14], corn stalks [15], orange peel [16], eucalyptus bark [17], coffee waste [18], and many other resources, have been proposed for the successful separation of anionic dyes, showing the variety and possibility of these raw materials for anionicdye removal. Therefore, this review aims to evaluate the findings reported for anionicdye adsorption from plant-derived agricultural wastes, and to explore the suitability of the proposed methods and reveal their future potential.

2. Types of Anionic Dyes

Based on their ionic nature, dyes can be classified into two categories: ionic and nonionic (Figure 2). Anionic dyes are a type of ionic dye in which anions are the key active groups. Some nonionic dyes, such as vat and disperse dyes, need to be converted into soluble anionic forms before being applied on a substrate [7]. Anionic dyes can be classified into reactive, direct, and acid dyes based on their application process.
Historically, the chemistry of dyes and pigments has mostly been studied in the context of textile materials, as the coloration of textiles demands that extensive quality parameters be met, which are not relevant to other industries. Understanding their mechanism of interaction with fibers (such as cellulose) can also provide an indication of their interaction ability when adsorbents from agricultural wastes are used for dye removal, as the chemical nature of these wastes is also mostly cellulosic or lignocellulosic.

2.1. Reactive Dye

Reactive dyes can form a covalent bond with certain fiber active sites, such as the hydroxyl group (-OH) or amino group (-NH2) [19]. Lignocellulosic biomasses contain plenty of hydroxyl groups around their structures, thus making them accessible to interact with reactive dyes. Commercial reactive dyes commonly contain the dichlorotriazine, trichloropyrimidine, aminochloro-s-triazine, sulphatoethylsulphone, dichloroquinoxaline, aminofluoro-s-triazine, and difluorochloropyrimide groups [20]. Because reactive dyes and the surface of cellulosic fibers become negatively charged in water, electrolytes (such as sodium chloride or Glauber salt) are often used to neutralize the fiber surface. No matter how reactive the chemical groups of these dyes are, it is hard to achieve over 80% dye fixation in actual dyeing, which causes the obvious discharge of unfixed dyes as hazardous effluent [21].

2.2. Direct Dye

Direct dyes are named from their high affinity towards cellulosic fiber, and they are thus said to be directly appliable onto cellulose-based fiber systems during dyeing. The major chromophores in the structure of direct dyes are the azo, anthraquinoid, phthalocyanine, and metal complexes [19]. These dyes are highly substantive, although not reactive. The use of electrolytes is also required for the successful attachment of direct dyes to cellulosic fibers.

2.3. Acid Dye

Acid dyes have acidic groups in their structures and are mainly applied on noncellulosic fibers, such as protein and polyamides in acidic conditions. The chromophoric systems in acid dyes are sulphonated azo, anthraquinone, nitrodiphenylamine, triphenylmethane, and xanthenes [22]. Cationized fibers can produce a bond with an anion of acid dyes through an electrostatic force [23].

3. Chemical Nature of Plant-Derived Agricultural Wastes

Common plant-based agricultural wastes are a complex combination of cellulose, hemicellulose, and lignin, and are thus also widely known as lignocellulosic biomass. As shown in Figure 3, these constituents are complexly attached in the lignocellulosic structure. Cellulose is the major component that is linearly formed from the D-glucose subunits, and it is commonly known for its higher crystallinity [24]. However, the hemicellulose part (such as pentose and hexose) is randomly oriented within cellulose chains and is known to impact negatively on the crystallinity. Their chain length is much shorter than the typical chain length of cellulose. The other part, lignin, is the outer layer that holds together both cellulose and hemicellulose [25]. It is a complex structure of cross-linked phenolic subunits, although its actual structure is still to be identified [26]. Throughout the lignocellulosic structure, there are abundant hydroxyl groups, which make this material highly hydrophilic and ideal for dye adsorption from water. For the same reason, their surface is electronegative, and they attract cationic dyes more as the cations of dye become easily attached to the anionic hydroxyl groups. Therefore, their properties need to be altered for efficient anionic dye removal, which has frequently been attempted by researchers over the years.

4. Adsorbents for Anionic Dyes

As mentioned above, due to the electronegative nature of the surface of agricultural wastes (i.e., lignocellulosic structure), anionic dyes show a repulsive effect during the attachment through ionic interactions, if the adsorbent surface is not modified [7]. The modification of these materials can be divided into three main broad concepts: (1) mechanical, (2) chemical, and (3) thermal modifications. The concept of mechanical modification is mainly to reduce the particle size of the initial material by cutting or mechanical milling [27]. This may involve some chemical treatment (e.g., for cleaning purposes), but it will not necessarily alter the chemical groups or ionic behavior. However, the principle of the second concept (i.e., chemical modification) is exactly the opposite, where the modification of the chemical groups is performed to make them chemically more interactive with anionic dyes [16]. This method may involve mechanical milling as a preliminary step, although chemical treatments are the main driving factor for the dye adsorption. The other concept is thermal modification, where both mechanical milling and chemical treatment may be present, although the main principle is to convert the biomass into a carbonaceous material (such as activated carbon) by high-temperature carbonization processes [28], which often result in a high surface area, which leads to the key improvement in the adsorption property. All three methods are promising from different aspects and have their benefits and drawbacks.

4.1. Mechanical Modifications

The mechanical modification of biomass mainly involves the mechanical-milling operation, where the materials are crushed into fine or coarse powder (Figure 4 and Table 1). Although the chemical properties are not altered, the adsorption capacity largely depends on bath conditions and is often found as a complex synergistic influence from many parameters. Although anionic dyes have a general repulsive effect on the lignocellulosic surface, at the initial stage, the dye adsorption largely depends on the force generated by the dye-concentration gradient from the solution [29]. At the beginning of adsorption, this force helps the dye molecules to reach the adsorbent surface, overcoming the repulsive effect. However, after a certain level, when the concentration gradient is reduced by some of the dyes adsorbed onto the adsorbent, the repulsive effect starts to take the active role and limits the adsorption rate. Hence, even with a similar surface area, mechanically modified biomass can often adsorb more cationic dyes than anionic dyes [30]. This indicates that there is the possibility of remaining vacant sites inside the adsorbent (not completely saturated), even when the equilibrium is reached.
It is almost a rule of thumb that a smaller-size particle possesses more surface area, and hence, it can hold a greater amount of dye. This has been found to be true in countless studies, and it has been proven for anionic dye adsorption as well. For example, the adsorption capacity of eucalyptus bark towards Solar Red BA (a direct dye) was below 10 mg/g when a particle size of 0.51–0.71 mm was used, although it reached 18.15 mg/g when a <0.255 mm size was used [17]. A similar observation was also found in another report in which the adsorption of Eriochrome Black T was significantly improved (~22.8%) when the particle size of an almond-shell adsorbent was reduced from >500 µm to <100 µm [37].

4.2. Chemical Modifications

Due to the poor interaction of lignocellulosic biomass with anionic dyes [1], the chemical modification of their surface has received keen attention. In most cases, the main idea was to bring cationic groups to the surface, which facilitated the dye adsorption. Several methods of achieving the cationized adsorbent are reported in varied dimensions. As shown in Figure 5, these can be divided into four main categories: treatment with amine-based chemicals, cationic surfactants, quaternary ammonium compounds, and acid solutions.
Different amine-rich compounds have mostly been chosen by researchers for the chemical modification of lignocellulosic biomass because they contain a high number of amines that easily become attached to lignocellulose and attract anionic dyes. However, these methods often require multiple steps of chemical treatment to effectively modify the structure. For instance, Wong et al. used polyethylenimine to treat coffee waste for the adsorption of reactive and acid dye [18]. The method included the impregnation of coffee-waste powder in a polyethylenimine solution at a moderate temperature (65 °C) for 6 h, and then cross-linking using glutaraldehyde. This delivered 77.5 mg/g and 34.4 mg/g maximum adsorptions of Reactive Black 5 and Congo Red dyes, respectively. Nevertheless, a comparison with the unmodified coffee waste was not reported to validate the true influence of the cationization. In another study, nanocrystalline cellulose extracted from hardwood bleached pulp was amino-functionalized by ethylenediamine [44]. The process involved the treatment of ethylenediamine at 30 °C for 6 h, and an in situ reduction by NaBH4 at room temperature for 3 h, which resulted in the 555.6 mg/g maximum adsorption of Acid Red GR, although no comparison was shown with unmodified lignocellulose. In a different study, wheat straw was treated with epichlorohydrin and N,N-dimethylformamide at 85 °C for 1 h. Furthermore, an ethylenediamine treatment (45 min at 8 °C) and trimethylamine treatments (120 min at 85 °C) were applied to achieve the final amino-modified wheat-straw powder, which showed maximum adsorption capacities of 714.3 mg/g for anionic dye (Acid Red 73) and 285.7 mg/g for reactive dye (Reactive Red 24) [12].
Coating with amine-based polymer was also performed, where sawdust coated with polyaniline nanofibers removed up to 212.97 mg/g of Acid Red G [45]. These nanofibers were prepared by the polymerization of an aniline monomer and oxidant ((FeCl3). The inclusion of polyaniline nanofibers enhanced the surface area of the biomass (from 1.22 to 16.26 m2/g), which was useful for the adsorption. In one study, an amine-rich biopolymer, (chitosan) was used to modify lignocellulose for anionic dye adsorption [7]. The chitosan was dissolved in an acetic acid solution, and the cotton-gin-trash film was impregnated for 1 h at room temperature. The resultant modified adsorbent showed a significant increase in the adsorption capacity of Acid Blue 25 (151.5 mg/g) compared with that of the unmodified biomass (35.5 mg/g). The amine-based adsorbent was also prepared as a form of a nanocomposite. For instance, banana-peel cellulose was combined with chitosan and silver nanoparticles to investigate its suitability to adsorb Reactive Orange 5, reaching a maximum adsorption capacity of 125 mg/g [14].
In the second category, treatment with cationic surfactants commonly increases the aliphatic carbon content in biomass [15], but most importantly, they induce the cationic groups on the surface, which help anionic dye removal. Several studies have been performed to modify lignocellulosic biomass using different cationic surfactants. These include cetylpyridinium bromide [15], hexadecylpyridinium bromide [46,47], hexadecylpyridinium chloride [48], hexadecylpyridinium chloride monohydrate [49], cetyltrimethylammonium bromide [50], and hexadecyltrimethylammonium bromide [51]. This technique commonly involves the continuous shaking of the crushed lignocellulose particles in the surfactant solutions, and mostly at room temperature. However, the maximum adsorption achieved through this technique is somewhat lower than the amine-based methods. For example, a few of the adsorption capacities reported for this method are 70 mg/g dye by modified wheat straw [47], 30.8–31.1 mg/g dye by modified corn stalks [15], 55–89 mg/g dye by modified coir pith [51], and 146.2 ± 2.4 mg/g dye by modified orange-peel powder [16], which are comparatively lower than most of those reported for amine-based modifications (Table 2). This is probably related to the abundance of amide groups [18] in the amine-based compounds, which are more prone to attach to the lignocellulosic structure, as well as with the anionic dyes, by electrostatic attraction [7,18].
There are few studies on the modification of lignocellulose by quaternary ammonium compounds for anionic dye adsorption, which mostly showed high efficiency. For example, Jiang and Hu reported hydroxypropyloctadecyldimethylammonium-modified rice husk cellulose for reactive (Diamine Green B) and acid-dye (Congo Red and Acid Black 24) adsorption. This modification required multiple steps, which included alkaline and ultrasonic treatments, and treatments with epichlorohydrin and N,N-dimethyl-1-octadecylamine/isopropanol. The resultant adsorbent was able to adsorb up to ~580 mg/g of the anionic dye (Congo Red) [52]. In another study, biomass from palm kernel shell was quaternized using N-(3-chloro-2-hydroxypropyl)trimethylammonium chloride in an alkaline aqueous mixture, which resulted in 207.5 mg/g of maximum adsorption [53].
The fourth category of the chemical modification of the lignocellulose surface is through acidic treatment. For instance, it was claimed that a phosphoric acid treatment worked by increasing the surface area and led to the grafting of phosphate onto the lignocellulose structure. However, the adsorption capacity reported in the proposed method was quite low (15.96 mg/g) [54]. A similar result (adsorption capacity of 13.39 mg/g of Reactive Red 196) was reported when concentrated HCl was used for sawdust treatment [13]. Slightly higher adsorption has also been reported, such as 40.98 mg/g of Drimarine Black CL-B removal by HCl-modified peanut husk [55]. The only exception (a higher removal efficiency) was the treatment of fermentation wastes with nitric acid, where the raw biomass structure was reported to be protonated, replacing the naturally available ionic species, and thereby resulting in a 185.2 mg/g adsorption of Reactive Black 5 [56]. However, there was no following study on the nitric acid treatment.
A combination of amine-based chemicals and acid treatment, such as cationization by dichloroethane and methyl amine, followed by protonation using acetic acid, was also reported in one study to modify the biomass surface [16]. The resultant orange-peel adsorbent showed a maximum adsorption capacity of 163 mg/g of Congo Red. Furthermore, the silylation of biomass was also found to be effective for anionic dye adsorption, where a maximum adsorption of 208.33 mg/g was reported [57]. Moreover, Lin et al. proposed an amphoteric-modification method of delignified wheat straw by a monochloroacetic acid treatment, and then grafting with 2-dimethylamino ethyl methacrylate. The resultant adsorbent could be used to adsorb both anionic and cationic dyes by switching between acidic or alkaline conditions ranging from pH 2 to 10 [58].
Aside from the four major kinds of treatments, some other methods, such as modifications of biomass using oxides and metal salts, have also been reported. For example, the modification of wood biowaste using Al2O3 by the solvo-reaction method showed a high adsorption capacity of Reactive Blue 19 (441.9 mg/g) [59]. Furthermore, the modification of sawdust using FeCl3 (maximum 212.97 mg/g Acid Red G removal) [45], the use of ZnCl2 on pulp waste (maximum 285.7 mg/g Methyl Orange removal) [60], hazelnut bagasse (maximum 450 mg/g Acid Blue 350 removal) [61], and pine cone (maximum 118.06 mg/g Alizarin Red S removal) [62], and the use of CaCl2 with peanut husk (maximum 40.98 mg/g Drimarine Black CL-B removal) [55], have also been reported, showing considerable efficiency.
Overall, the trend in Table 2 shows that effective dye removal by chemically modified adsorbents is more related to the modification methods, rather than to the type of dye. For example, the acidic modification of lignocellulose often produced a lower efficiency either for acid or reactive dyes, although the adsorption capacity was reported to be higher for both when amine-based chemicals were used. When the same adsorbent was used to adsorb acid and reactive dyes, mostly the acid dyes showed a higher removal efficiency [12,51,52], although an exception was also seen [48]. Modification with different amine-based chemicals was found to be more effective for anionic dye adsorption, which was probably related to the abundance of cationic amine groups in these compounds.
Table 2. Maximum adsorption capacities (qmax) of anionic dye adsorbents derived from chemically modified agricultural wastes.
Table 2. Maximum adsorption capacities (qmax) of anionic dye adsorbents derived from chemically modified agricultural wastes.
ResourcesChemical ModificationModification TypeDyeqmax (mg/g)Reference
Wheat strawGrafting with 2-dimethylamino ethyl methacrylate monomerAmine-basedOrange II506[58]
Wheat strawTreatment with hexadecylpyridinium bromideCationic surfactantLight Green70.01 ± 3.39[47]
Peanut huskModified by hexadecylpyridinium bromide Cationic surfactantLight Green60.5[46]
Wheat strawTreatment with epichlorohydrin,
N,N-dimethylformamide, ethylenediamine, and trimethylamine
Amine-basedAcid Red 73714.3[12]
Wheat strawTreatment with epichlorohydrin,
N,N-dimethylformamide, ethylenediamine, and trimethylamine
Amine-basedReactive Red 24285.7[12]
Barley strawTreatment with hexadecylpyridinium chloride monohydrateCationic surfactantAcid Blue 4051.95[49]
Barley strawTreatment with hexadecylpyridinium chloride monohydrateCationic surfactantReactive Blue 431.5[49]
Corn stalksTreatment by cetylpyridinium bromideCationic surfactantAcid Red 30.77[15]
Corn stalksTreatment by cetylpyridinium bromideCationic surfactantAcid Orange31.06[15]
Banana peelReinforcement with nanoparticles and chitosan Amine-basedReactive Orange 5125[14]
SawdustCoating with polyaniline Amine-basedAcid Red G212.97[45]
SawdustTreatment with cetyltrimethylammonium bromideCationic surfactantCongo Red9.1[50]
SawdustTreatment with concentrate HClAcid treatmentReactive Red 19613.39[13]
Rice huskTreatment with
hydroxypropyloctadecyldimethylammonium
Quaternary ammonium compoundsDiamine Green B207.15[52]
Rice huskTreatment with
hydroxypropyloctadecyldimethylammonium
Quaternary ammonium compoundsAcid Black 24268.88
[52]
Rice huskTreatment with
hydroxypropyloctadecyldimethylammonium
Quaternary ammonium compoundsCongo
Red
580.09[52]
Peanut huskTreatment with alginate and CaCl2Metal saltDrimarine Black CL-B40.98[55]
Peanut huskTreatment by hexadecylpyridinium bromide in batch modeCationic surfactantLight Green146.2 ± 2.4[63]
Peanut huskHydrochloric acid treatmentAcid treatmentDrimarine Black CL-B51.02[55]
Oil palm empty fruit bunchesSilylationAmine-basedProcion Red 208.33[57]
Orange peel Treatment with dichloroethane, methyl amine, and acetic acidAmine-basedCongo Red163[16]
Cotton gin trashCationized by chitosanAmine-basedAcid Blue 25151.52[7]
Aquatic plantPhosphoric acid treatment and low-temperature activationAcid treatmentDirect Red 8915.96[54]
Coffee wasteTreatment with
polyethylenimine
Amine-basedReactive Black 577.52[18]
Coffee wasteTreatment with
polyethylenimine
Amine-basedCongo Red34.36[18]
Hardwood kraft pulpGrafting cellulose nanocrystals
with ethylenediamine
Amine-basedAcid Red GR555.6[44]
Palm kernel shellQuaternized by N-(3-chloro-2-hydroxypropyl) trimethylammonium chlorideQuaternary ammonium compoundsReactive Black 5207.5[53]
Fermentation wasteProtonated by nitric acidAcid treatmentReactive Black 5185.2[56]
Waste coir pithTreatment with hexadecyltrimethylammonium solutionCationic surfactantAcid Brilliant
Blue
159[51]
Waste coir pithTreatment with hexadecyltrimethylammonium solutionCationic surfactantProcion Orange89[51]
Wood residueAluminum oxide modificationMetal oxideReactive Blue 1929.83[64]
Wood biowasteAluminum oxide modificationMetal oxideReactive Blue 19441.9[59]

4.3. Thermal Modifications

The thermal modification of plant-derived wastes is mainly performed to prepare activated carbon (AC). AC is an advanced form of charcoal that is basically a porous carbonaceous structure with an extended surface area [65]. It has been accepted as an effective adsorbent and has been applied widely in water and wastewater treatment, air purification, and solvent recovery, as well as in the pharmaceutical and medical industries and industrial processes. According to a market research report, the market demand for AC was estimated to be USD 5.7 billion worldwide in 2021, and it is predicted to reach USD 8.9 billion by 2026 [66]. A variety of plant-derived wastes have been used as raw materials for the preparation of AC for anionic dye adsorption, as these resources are rich in carbon content, cost effective, renewable, and abundant [67].
Carbonization and activation are basically the two steps that are involved in preparing AC [68]. Activation can be performed by different methods, such as mechanical or thermal, chemical, or a combination of both (i.e., physiochemical) [67], and it can also be biological [69]. These include several methods and their combinations, such as chemical treatment, air oxidation, electrochemical oxidation, and plasma, microwave, and ozone treatment for enhancing the adsorption performance of AC [70,71,72,73,74]. However, for AC prepared for anionic dye adsorption, the chemical-activation technique (using activating agents, such as zinc chloride or phosphoric acid) [75] is mostly preferred (Figure 6) because of its faster activation time, higher carbon yield, simplicity, lower operating temperature, and well-developed pore structure. The proper selection of materials can enhance the adsorption potentiality of AC through the development of its pore structure (total pore volume, pore size distribution) and surface functionalities [76].
The main functional groups on the surface of AC that are responsible for removing dyes are carboxyl, carbonyl, phenols, lactones, and quinones. A positive surface charge on AC can be attained by applying an alkaline treatment, which can accelerate the adsorption of anionic dyes. The porous carbon of alkali-treated AC can enhance the reaction with acid-dye molecules by dipole–dipole, H-bonding, and covalent bonding [69]. As listed in Table 3, ACs prepared from many agricultural byproducts, such as peanut hull, coir pith, rice husk, coffee husk, and water hyacinth, have successfully been used for eliminating anionic dyes from wastewater.
The dye-adsorption capacity of different ACs is substantially controlled by the pH of the dye solution, possibly because the electrostatic interaction between ionized dye molecules and carbon surface functionalities that change the solution pH and rate of adsorption also varies. In the case of anionic dye, high dye uptake is favored at a lower pH or acidic medium, as the complete protonation of the carbon surface functionalities enhances the electrostatic attraction with anionic dyes, thereby increasing the dye uptake. From previous studies, it was observed that the OH group in the solution increased with an increasing pH, producing electrostatic repulsion with anionic dye, and thus decreasing the dye uptake. Therefore, the maximum adsorption of anionic dyes was observed in a pH range of 2–6 [79,84]. Although the surface area of the AC is known as a key factor for dye adsorption because it provides more spaces for the dye molecules to be attached, it can be perceived from Table 3 that this was not always true for anionic dye adsorption due to ionic interactions, which could have a positive or negative impact, depending on the specific adsorption system. For example, the maximum anionic dye absorption (960 mg/g) was found in AC obtained from pink shower seed pods, although it contained a moderate surface area (283.4 m2/g) [75], where the strong electrostatic attraction between the functional groups of the AC surface and anionic dye was facilitated with π–π interactions between the free electrons of the aromatic rings of the Congo Red structure, and the delocalized π-electrons on the basal planes of carbon.

4.4. The Effect of Process Conditions

Other than the modification performed on the biomass, if the amount of adsorbent is kept constant, then different process parameters, which include the pH, temperature, dye concentration, and chemical composition of the biomass, can impact the overall adsorption.
The pH of the solution plays a vital role during the adsorption. In numerous studies, for anionic dye adsorption, the favorable pH was reported as 2 [27,30,39,43], and in some cases, even 1 [31,35]. This is directly related to the change in the surface charge in lignocellulose that is altered at a lower pH. A lower pH reduces the electronegativity of the surface, and can even alter it to a positive surface charge [37], thus possibly attracting anionic dyes more conveniently. This was found to be true in cases of different kinds of anionic dyes, such as acid dyes [30,43], reactive dyes [27], and direct dyes [39]. Although this was seen as a common trend, some exceptions were also reported in which the pH showed a minor impact [38], or in which a moderate pH (7.5) was found to be more effective for better adsorption [88].
The bath temperature is also important because a higher temperature can often improve the mobility of dye molecules, and thus, the hindering force from the electronegative adsorbent surface diminishes [89]. As a result, a greater amount of dye can be transported into the adsorbent surface. Added to that, swelling can result in the adsorbent, which can promote intraparticle diffusion inside the structure [89,90], allowing more dyes be adsorbed onto the surface. However, the opposite phenomenon was also frequently reported (exothermic adsorption) [17,37,41], where a higher temperature was unfavorable for dye uptake. This was explained as a possible weakening of the related forces of attraction by a higher temperature [17,40].
The concentration of dye present in a solution largely affects the adsorption rate, and particularly at the initial stage. If the dye concentration is high, then the rate of adsorption will be higher, and vice versa. The content of cellulose in the biomass is also reported as an influencing factor in the adsorption. In a study, three different biomasses (i.e., coconut shells, cauliflower cores, and broccoli stalks) were used for the adsorption of the same anionic dyes, although the chemical compositions of the biomasses were different. The total cellulose, hemicellulose, starch, and pectinic sugar in the cauliflower cores were the highest among the three and resulted in a higher uptake of reactive and acid dyes [89]. Furthermore, the adsorption capacity was particularly significantly affected by the impact of cellulose as the external layer, rather than hemicellulose and lignin [89]. The chemical groups in lignocellulosic biomass (such as –OH, C=O, and C–O) are also good candidates to form hydrogen bonding with anionic dyes [30,43].

5. Adsorption Isotherms

Adsorption isotherms provide dye equilibrium relations that are widely used to assess the appropriateness of an adsorption mechanism. The most commonly used isotherms are the Langmuir and Freundlich models for any dye adsorption, which have also frequently been used to evaluate anionic dye attachment with plant-derived agricultural wastes.
The Langmuir isotherm model can be linearized as:
C e / q e = C e / q max + 1 / ( k L q max )
where qe is the amount of anionic dye absorbed by the adsorbent at equilibrium; Ce is the amount of dye left in the solution (mg/L) at the same point; qmax is the expression of the maximum dye-adsorption capacity (mg/g) by the adsorbent; kL is the binding affinity (L/mg), which is also known as the Langmuir constant. The Langmuir constant (kL) can be used to derive the suitability of the adsorption system. For example, the equilibrium parameter (RL) can be calculated using the value of the kL by the following equation [91]:
R L = 1 / ( 1 + k L C 0 )
where C0 is the initial dye concentration (mg/L) in the solution before the adsorption begins. The calculated RL indicates whether the adsorption system was favorable or not. For instance, if 0 < RL < 1, then the adsorption is favorable, if RL > 1, then the adsorption is unfavorable, if RL = 1, then the adsorption is linear, and if RL = 0, then the adsorption is irreversible.
The other frequently used isotherm is the Freundlich model, which can be linearized as follows [92]:
ln   q e = ln   k F ( 1 / n ) ln   C e
where kF is the adsorption capacity, and n is the adsorption intensity. The n value is also used to measure the deviation from the linearity (or heterogeneity). If 1/n < 1 (n ranges from 1 to 10), then the adsorption is counted as favorable, and if 1/n > 1, then the adsorption is considered unfavorable. If 1/n is close to zero, then the adsorbent is more heterogeneous, and vice versa [92]. Further indications were also noted from the value of the n, such as n = 1 indicates the linear nature of the adsorption, n < 1 indicates chemical adsorption, and n > 1 indicates physical adsorption [93].
While the Langmuir isotherm is considered more related to monolayer adsorption, the Freundlich isotherm is considered for multilayer adsorption. The coefficient-of-determination (R2) values of the Langmuir and Freundlich models listed in Table 4 across different kinds of dyes (acid, reactive, and direct), and after different sorts of modifications (mechanical, chemical, and thermal), show a sporadic result, and often closer values of the R2 between the Langmuir and Freundlich models. However, as a traditional pattern, often either chemisorption [38,88] or physisorption [7,31] has been claimed as the mechanism of interaction during the adsorption. A combination of multiple mechanisms has also been reported on many occasions. These include different combinations of physical and chemical adsorption [39,40,41,57], hydrogen bonding [89], ion exchange [34,94], valence forces [58], electrostatic and π–π interactions [75], ion exchange and chemical reaction [84], and so on. These interactions were only predicted and were not always confirmed through chemical analysis. Nevertheless, there are great possibilities for many of these interactions during the adsorption. For instance, the zeta potential (ζ) of chitosan-modified lignocellulose biomass transformed the ζ value from −10.8 mV to +5.1 mV, making it more appropriate for an electrostatic attraction towards anionic dye [7]. Furthermore, because lignocellulose is rich in phenolic groups from lignin, there are chances for π–π interactions between the rings present in the anionic dye structure. A schematic representation of different possible dye–adsorbent interactions during anionic dye separation is shown in Figure 7a. The mechanism of adsorption can also differ based on the concentration of dyes. For example, with lower and higher concentrations of dyes, monolayer and multilayer adsorption were observed, respectively [53], while simultaneous monolayer and multilayer adsorption was also reported [84]. Therefore, often the physical interaction between the dye and adsorbent is likely to be combined with chemical interactions, depending on whether there are enough chemical sites present for the reactions and favorable conditions.
Some other less commonly used isotherms for anionic dyes include the Redlich–Peterson model [84], Dubinin–Radushkevich model, Temkin model, and Flory–Huggins model [7].
The Redlich–Peterson model is a three-parameter model, and it is known as a combined form of the Langmuir and Freundlich models. The linear form can be expressed as [84]:
C e q e = 1 K RP + a RP K RP C e b RP
where KRP and aRP are the Redlich–Peterson constants, and bRP is an exponent from 0 and 1. The fitting results in a Redlich–Peterson model have occasionally been used for anionic dye adsorption to confirm and represent the adsorption mechanism as a combination of both the Langmuir and Freundlich models [31] (for instance, the simultaneous involvement of monolayer and multilayer adsorption [84]).
The linear form of the Dubinin–Radushkevich isotherm can be expressed as [95]:
ln   q e = ln   q m k DR ε 2
where qm represents the maximum adsorption capacity, kDR is the isotherm constant, and ε represents the Polanyi potential and is calculated by the following equation:
ε = RT   ln   ( 1 + 1 C e )
where T is the equilibrium temperature, and R represents the ideal gas constant (8.314 J·mol−1·K−1) [39]. The isotherm constant (kDR) is used to measure the free energy (E) of adsorption:
E = 1 / 2 k DR
If the E is found to be <8 kJ/mol, then this indicates physisorption, and when the E ranges 8–16 kJ/mol, then a chemisorption process is assumed [95]. This isotherm model has sometimes been used to determine between physisorption and chemisorption [39], and to justify the findings from other models [7].
The linear form of the Temkin model can be shown as:
q e = B 1 ln ( k T C e )
where kT is the Temkin constant, and B1 is derived from the below equation:
B 1 = RT b T
where bT is another Temkin constant related to the heat of adsorption [45]. From a good fitting with the Temkin model for anionic dye adsorption, the adsorption potential is predicted. The value of the Temkin constants can be used to evaluate the decrease and increase in the binding energy at different adsorption temperatures [45].
The linear equation of the Flory–Huggins model can be expressed as [96]:
ln θ C 0 = ln k FH + x ln ( 1 θ )
where kFH is the Flory–Huggins constant, x represents the remaining dye particles in the adsorption sites, and θ represents the degree of the adsorbent surface coverage by the dye particles and is derived from the following equation:
θ = 1 C e C 0
When the value of the kFH is positive, a suitable and spontaneous adsorption system is likely [96]. This model can be used to determine the surface coverage of the dye particles onto the adsorbents, although it is used infrequently for anionic dye adsorption [7].
However, as mentioned above, the use of these models was less frequent for anionic dye adsorption compared with the Langmuir and Freundlich isotherms, and they often showed an inferior fitting.

6. Kinetics and Diffusion

The kinetics of anionic dye adsorption on biomass-derived adsorbents has mainly been investigated by pseudo-first-order and pseudo-second-order kinetic models, and often linear equations have been used to draw the plots and find the fitting of the data [18,40,54,57,75]. The first-order model is also known as the Lagergren equation, which relates to the adsorption rate associated with the number of unoccupied areas of the dye [97]. The linear equation of the pseudo-first-order model is represented as:
ln ( q e q t ) = ln q e k 1 t
where k1 is the first-order adsorption-rate constant, and qt is the weight of the dye adsorbed (mg/g) at time (t) [98,99]. When the data better fit with the first-order equation, the adsorption is more likely to be a physisorption process.
The pseudo-second-order model is also parallelly used with the first-order kinetic model to check the adsorption feasibility. There are four linear equations of this model, which can be derived from each other [92,100]:
t q t = 1 k 2 q e 2 + t q e
1 q t = 1 q e + 1 k 2 q e 2 ( 1 t )
q t = q e 1 k 2 q e ( q t t )
q t t = k 2 q e 2 k 2 ( q e ) q t
where k2 is the second-order adsorption-rate constant. In practice, Equation (13) is comprehensively used. However, after plotting the data, this equation often shows a very good fit (e.g., R2 > 0.99), which is related to the presence of the t value in both the x-axis and y-axis [101]. This can be justified by fitting the same data with another linear equation in which t is not present on both sides and can result in a poor fit (Figure 8). When the data better fit with the second-order equation, the adsorption is more related to a chemisorption process. However, fitting with second-order models for anionic dye adsorbents was often not justified in the literature by plotting with more than one equation when linear models were used. Furthermore, the conversion of the data necessary for the linearization of these kinetic models can also lead to an alteration of the error structure and variation in the weight in data points, and it can also introduce errors into the independent variable [102]. However, using a nonlinear model of second-order kinetics does not alter the kinetic parameters, and thus it is considered to be more accurate and is recommended for estimating the kinetic parameters [103,104].
The molecular-diffusion mechanism is another key factor in the adsorption rate that cannot be identified by the pseudo models. If the adsorption is a physisorption phenomenon, then there is more possibility that the diffusion is the rate-controlling step [101]. In an adsorbent/dye-solution adsorption process, the transfer of dye particles into the adsorbent can be controlled by one of the three known mechanisms: film diffusion, intraparticle diffusion, or a combination of both [105]. Film diffusion is the transfer of dye particles to the adsorbent surface, and intraparticle diffusion is the diffusion of dye particles into the adsorbent pores (Figure 7b). It was reported that both film diffusion (also known as external mass transfer) and intraparticle diffusion can play a significant role during anionic dye adsorption [35].
The linear form of the intraparticle-diffusion model can be expressed as [92]:
q t = k ID . t 1 / 2 + C
where kID represents the diffusion-rate constant, and C is the thickness of the boundary layer (Figure 7b). If C is considerably higher, then a superior influence of the surface adsorption on the adsorption rate is likely [106]. As per this model, if the fitted line goes through the origin (0, 0), then this is a sign of the sole control of intraparticle diffusion [107]. However, in most cases, more than one step of adsorption is observed from the plotting of the intraparticle-diffusion model. These include the initial film diffusion, followed by intraparticle diffusion, and possibly a third and slow adsorbing stage close to an equilibrium.

7. Thermodynamics of Adsorption

Thermodynamic studies of dye adsorption are performed when the adsorption is conducted at different temperatures. From the pseudo-second-order constant (k2), the activation energy (Ea) of the adsorption process can be determined by the Arrhenius equation:
lnk 2 = lnA E a RT
where A is the Arrhenius factor, Ea is the Arrhenius activation energy (kJ/mol), T represents the absolute temperature (K), and R represents the gas constant [99].
The entropy (ΔS°) and enthalpy (ΔH°) are derived using the Eyring formula, which is expressed as:
ln ( k 2 T ) = ln ( K b h ) + Δ S ° R Δ H ° RT
where Kb is the Boltzman constant, and h is the Planck constant. A positive value of the ΔH° indicates an endothermic reaction (adsorption increases with increasing temperature) [39], while a negative value of the ΔH° indicates an exothermic process (adsorption decreases with increasing temperature) [100]. A positive value of the ΔS° indicates a decrease in the randomness of solid/liquid interfaces, and vice versa [39].
Finally, the Gibbs energy of activation (ΔG°) is calculated from the following formula:
Δ G ° = Δ H ° T Δ S °
A positive value of the ΔG° indicates nonspontaneous reactions, which need an input of energy, while a negative value of the ΔG° indicates spontaneous reactions, which can continue without external energy [108].
Thermodynamic studies have been extensively conducted for anionic dye adsorption, and the values of the ΔS°, ΔH°, and ΔG° have been used to determine whether the process is spontaneous/nonspontaneous and exothermic/endothermic, as well as the randomness in the adsorbent/solution interfaces. As shown in Table 5, in most cases, the mechanism was found to be spontaneous (negative ΔG°) for either kind of dye or the modification methods. Although the endothermic process (negative ΔH°) was more observed, the exothermic process (positive ΔH°) was also seen. In most cases, positive ΔS° was also seen, which means more uniformity during the anionic dye adsorption.

8. Circularity and Sustainability

To prove the adsorbents as sustainable and economically viable, the circularity of the adsorbents has received significant attention. In this regard, the regeneration ability of the adsorbents is a key criterion. If the adsorbed dyes can be successfully desorbed from the adsorbent, then the reuse of both the adsorbent and dye could be possible, and the process becomes more economically feasible [34]. Regeneration studies have been widely considered for anionic dye adsorbents, and the results are promising. These are often performed by altering the pH of the solution. When anionic dyes are physically bound to the adsorbent, an alteration of the pH, often alkaline [33], could be more useful to transfer them into the solution [7]. For example, a desorption experiment using different components (i.e., water, HCl, acetic acid, NaOH, and ethanol) showed a maximum Methyl Orange desorption of 67.85% by a NaOH solution, while the second best was 34.5% by acetic acid [60]. This was also confirmed by different studies in which a 45 min treatment of the adsorbent with 0.1 M NaOH at 20 °C was enough to desorb most of the anionic dyes and reuse the adsorbents [7,109]. In many studies, the regeneration of anionic dyes and the reuse of the adsorbents were proven highly feasible. For example, 45 consecutive cycles from sawdust adsorbent were reported [45], while at least 3 cycles were possible by a cotton-gin-trash bioadsorbent [7] and activated carbon prepared from pine cone [62]. Although high desorption efficiencies, such as 98% [16] or 83.7–90.3% dye desorption, have been reported [77], a lower desorption rate, such as 35–43%, was also reported when the dyes were attached with adsorbent through a chemical reaction (chemisorption) [15]. However, more than half of the studies on anionic dye removal have considered desorption studies to justify the effectiveness of the proposed adsorbents, which shows the requirement for sustainability. From an economic evaluation, it was reported that the cost for the preparation and regeneration of adsorbents of biomass waste (USD 42.43/kg) was significantly lower than commercially available activated carbon (USD 111.37/kg) [60]. From both the economic and environmental points of view, the adsorbents prepared from natural plant-derived agricultural wastes seem to be more sustainable, as, on the one hand, they reduce the process cost, and, on the other hand, they diminish the amount of waste from the environment. The use of these adsorbents in a circular manner by successful regeneration is further beneficial. When the adsorbents reach their end of life, the dyes can be desorbed and, as the adsorbents are biodegradable, they can be discarded into the land without any hazard. Overall, the extent of the dye-desorption studies conducted along with the adsorption of anionic dyes indicates the urgency of making these bioadsorbents more sustainable and efficient, validating their more feasible use in the future.

9. Conclusions and Future Scopes

Anionic dyes are widely used in different applications and are one of the major contaminants in dye wastewater. Over the years, lignocellulosic agricultural wastes have been widely used to remove these dyes from wastewater. Adsorbing anionic dyes by these adsorbents is challenging, as they inherently possess some repulsing effect on anionic dyes due to their electronegative surfaces. Nevertheless, given the low cost and abundance of these natural resources, researchers have frequently made them capable of anionic dye adsorption by mechanical, chemical, and thermal modifications. In cases in which only mechanical modifications are performed, adjusting the bath solution to a lower pH can enhance the dye adsorption. Reducing the particle size and tuning the temperature to a favorable condition can also promote dye adsorption. Although thermal modification, such as transformation into activated carbon, showed an improvement in the adsorption, chemically modifying the surface by inducing new cationic groups was found to be the most effective method of anionic dye removal when using these resources. The chemical modification of these biomasses was most successful when using amine-based and quaternary ammonium compounds. It is worth mentioning that quaternary ammonium compounds often possess environmental hazards, while amine-based compounds can also be hazardous, and thus require careful attention. Even though the use of amine-rich nonhazardous biopolymers, such as chitosan and its derivatives, have been well explored for the removal of anionic dyes, they are not widely considered for the modifications of lignocellulosic biomass (with a few exceptions), which could be a future research area. Furthermore, when the efficiency of a chemically modified biomass adsorbent was reported, often it was not parallelly compared with the chemically unmodified biomass. Thus, the degree of improvement in the dye adsorption was often not identified. However, the extent of regeneration studies shows the keen attention of researchers to not only just proposing an adsorbent to separate anionic dyes, but also to proving its sustainability in terms of cost reduction and effectiveness. The success trend portrays the promising future of using these abundant wastes as the key resources for the essential separation of anionic dyes.

Author Contributions

Conceptualization, A.N.M.A.H.; writing—original draft preparation, A.N.M.A.H., N.S., A.S.M.S. and S.A.S.; writing—review and editing, A.N.M.A.H. and N.S.; visualization, A.N.M.A.H., A.S.M.S. and S.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Haque, A.N.M.A.; Remadevi, R.; Wang, X.; Naebe, M. Sorption properties of fabricated film from cotton gin trash. Mater. Today Proc. 2020, 31, S221–S226. [Google Scholar] [CrossRef]
  2. Hassan, M.M.; Carr, C.M. A critical review on recent advancements of the removal of reactive dyes from dyehouse effluent by ion-exchange adsorbents. Chemosphere 2018, 209, 201–219. [Google Scholar] [CrossRef]
  3. Lellis, B.; Fávaro-Polonio, C.Z.; Pamphile, J.A.; Polonio, J.C. Effects of textile dyes on health and the environment and bioremediation potential of living organisms. Biotechnol. Res. Innov. 2019, 3, 275–290. [Google Scholar] [CrossRef]
  4. Salleh, M.A.M.; Mahmoud, D.K.; Karim, W.A.W.A.; Idris, A. Cationic and anionic dye adsorption by agricultural solid wastes: A comprehensive review. Desalination 2011, 280, 1–13. [Google Scholar] [CrossRef]
  5. Haque, A.N.M.A.; Remadevi, R.; Rojas, O.J.; Wang, X.; Naebe, M. Kinetics and equilibrium adsorption of methylene blue onto cotton gin trash bioadsorbents. Cellulose 2020, 27, 6485–6504. [Google Scholar] [CrossRef]
  6. Ray, R.C.; Behera, S.S. Solid state fermentation for production of microbial cellulases. In Biotechnology of Microbial Enzymes; Elsevier: Amsterdam, The Netherlands, 2017; pp. 43–79. [Google Scholar]
  7. Haque, A.N.M.A.; Remadevi, R.; Wang, X.; Naebe, M. Adsorption of anionic Acid Blue 25 on chitosan-modified cotton gin trash film. Cellulose 2020, 27, 9437–9456. [Google Scholar] [CrossRef]
  8. Roa, K.; Oyarce, E.; Boulett, A.; ALSamman, M.; Oyarzún, D.; Pizarro, G.D.C.; Sánchez, J. Lignocellulose-based materials and their application in the removal of dyes from water: A review. Sustain. Mater. Technol. 2021, 29, e00320. [Google Scholar] [CrossRef]
  9. Mishra, S.; Cheng, L.; Maiti, A. The utilization of agro-biomass/byproducts for effective bio-removal of dyes from dyeing wastewater: A comprehensive review. J. Environ. Chem. Eng. 2021, 9, 104901. [Google Scholar] [CrossRef]
  10. Mo, J.; Yang, Q.; Zhang, N.; Zhang, W.; Zheng, Y.; Zhang, Z. A review on agro-industrial waste (AIW) derived adsorbents for water and wastewater treatment. J. Environ. Manag. 2018, 227, 395–405. [Google Scholar] [CrossRef]
  11. Rachna, K.; Agarwal, A.; Singh, N. Rice husk and Sodium hydroxide activated Rice husk for removal of Reactive yellow dye from water. Mater. Today Proc. 2019, 12, 573–580. [Google Scholar] [CrossRef]
  12. Xu, X.; Gao, B.-Y.; Yue, Q.-Y.; Zhong, Q.-Q. Preparation and utilization of wheat straw bearing amine groups for the sorption of acid and reactive dyes from aqueous solutions. J. Hazard. Mater. 2010, 182, 1–9. [Google Scholar] [CrossRef] [PubMed]
  13. Doltabadi, M.; Alidadi, H.; Davoudi, M. Comparative study of cationic and anionic dye removal from aqueous solutions using sawdust-based adsorbent. Environ. Prog. Sustain. Energy 2016, 35, 1078–1090. [Google Scholar] [CrossRef]
  14. Abdelghaffar, F. Biosorption of anionic dye using nanocomposite derived from chitosan and silver Nanoparticles synthesized via cellulosic banana peel bio-waste. Environ. Technol. Innov. 2021, 24, 101852. [Google Scholar] [CrossRef]
  15. Soldatkina, L.; Zavrichko, M. Equilibrium, kinetic, and thermodynamic studies of anionic dyes adsorption on corn stalks modified by cetylpyridinium bromide. Colloids Interfaces 2018, 3, 4. [Google Scholar] [CrossRef]
  16. Munagapati, V.S.; Kim, D.-S. Adsorption of anionic azo dye Congo Red from aqueous solution by Cationic Modified Orange Peel Powder. J. Mol. Liq. 2016, 220, 540–548. [Google Scholar] [CrossRef]
  17. Tahir, M.A.; Bhatti, H.N.; Iqbal, M. Solar Red and Brittle Blue direct dyes adsorption onto Eucalyptus angophoroides bark: Equilibrium, kinetics and thermodynamic studies. J. Environ. Chem. Eng. 2016, 4, 2431–2439. [Google Scholar] [CrossRef]
  18. Wong, S.; Ghafar, N.A.; Ngadi, N.; Razmi, F.A.; Inuwa, I.M.; Mat, R.; Amin, N.A.S. Effective removal of anionic textile dyes using adsorbent synthesized from coffee waste. Sci. Rep. 2020, 10, 2928. [Google Scholar] [CrossRef]
  19. Broadbent, A.D. Basic Principles of Textile Coloration; Society of Dyers and Colorists Bradford: Bradford, UK, 2001; Volume 132. [Google Scholar]
  20. Khatri, A.; Peerzada, M.H.; Mohsin, M.; White, M. A review on developments in dyeing cotton fabrics with reactive dyes for reducing effluent pollution. J. Clean. Prod. 2015, 87, 50–57. [Google Scholar] [CrossRef]
  21. Haque, A.N.M.A.; Islam, M.A. The contribution of different vinyl sulphone-reactive dyes to an effluent. J. Taibah Univ. Sci. 2015, 9, 594–600. [Google Scholar] [CrossRef]
  22. Sekar, N. Acid dyes. In Handbook of Textile and Industrial Dyeing: Principles, Processes and Types of Dyes; Clark, M., Ed.; Woodhead Publishing Limited: Cambridge, UK, 2011; pp. 486–514. [Google Scholar]
  23. Chakraborty, J. Fundamentals and Practices in Colouration of Textiles; CRC Press: Boca Raton, FL, USA, 2015. [Google Scholar]
  24. Haque, A.N.M.A.; Zhang, Y.; Naebe, M. A review on lignocellulose/poly (vinyl alcohol) composites: Cleaner approaches for greener materials. Cellulose 2021, 28, 10741–10764. [Google Scholar] [CrossRef]
  25. Haque, A.N.M.A.; Remadevi, R.; Naebe, M. Lemongrass (Cymbopogon): A review on its structure, properties, applications and recent developments. Cellulose 2018, 25, 5455–5477. [Google Scholar] [CrossRef]
  26. Sanderson, K. Lignocellulose: A chewy problem. Nature 2011, 474, S12–S14. [Google Scholar] [CrossRef] [PubMed]
  27. Değermenci, G.D.; Değermenci, N.; Ayvaoğlu, V.; Durmaz, E.; Çakır, D.; Akan, E. Adsorption of reactive dyes on lignocellulosic waste; characterization, equilibrium, kinetic and thermodynamic studies. J. Clean. Prod. 2019, 225, 1220–1229. [Google Scholar] [CrossRef]
  28. Aziz, E.K.; Abdelmajid, R.; Rachid, L.M.; Mohammadine, E.H. Adsorptive removal of anionic dye from aqueous solutions using powdered and calcined vegetables wastes as low-cost adsorbent. Arab. J. Basic Appl. Sci. 2018, 25, 93–102. [Google Scholar] [CrossRef]
  29. Rodríguez, A.; García, J.; Ovejero, G.; Mestanza, M. Adsorption of anionic and cationic dyes on activated carbon from aqueous solutions: Equilibrium and kinetics. J. Hazard. Mater. 2009, 172, 1311–1320. [Google Scholar] [CrossRef]
  30. Stavrinou, A.; Aggelopoulos, C.; Tsakiroglou, C. Exploring the adsorption mechanisms of cationic and anionic dyes onto agricultural waste peels of banana, cucumber and potato: Adsorption kinetics and equilibrium isotherms as a tool. J. Environ. Chem. Eng. 2018, 6, 6958–6970. [Google Scholar] [CrossRef]
  31. Jain, S.N.; Tamboli, S.R.; Sutar, D.S.; Jadhav, S.R.; Marathe, J.V.; Shaikh, A.A.; Prajapati, A.A. Batch and continuous studies for adsorption of anionic dye onto waste tea residue: Kinetic, equilibrium, breakthrough and reusability studies. J. Clean. Prod. 2020, 252, 119778. [Google Scholar] [CrossRef]
  32. El Messaoudi, N.; Dbik, A.; El Khomri, M.; Sabour, A.; Bentahar, S.; Lacherai, A. Date stones of Phoenix dactylifera and jujube shells of Ziziphus lotus as potential biosorbents for anionic dye removal. Int. J. Phytoremediat. 2017, 19, 1047–1052. [Google Scholar] [CrossRef]
  33. Namasivayam, C.; Prabha, D.; Kumutha, M. Removal of direct red and acid brilliant blue by adsorption on to banana pith. Bioresour. Technol. 1998, 64, 77–79. [Google Scholar] [CrossRef]
  34. Reddy, M.S.; Sivaramakrishna, L.; Reddy, A.V. The use of an agricultural waste material, Jujuba seeds for the removal of anionic dye (Congo red) from aqueous medium. J. Hazard. Mater. 2012, 203, 118–127. [Google Scholar] [CrossRef]
  35. Tunç, Ö.; Tanacı, H.; Aksu, Z. Potential use of cotton plant wastes for the removal of Remazol Black B reactive dye. J. Hazard. Mater. 2009, 163, 187–198. [Google Scholar] [CrossRef] [PubMed]
  36. Parimelazhagan, V.; Jeppu, G.; Rampal, N. Continuous Fixed-Bed Column Studies on Congo Red Dye Adsorption-Desorption Using Free and Immobilized Nelumbo nucifera Leaf Adsorbent. Polymers 2021, 14, 54. [Google Scholar] [CrossRef] [PubMed]
  37. Ben Arfi, R.; Karoui, S.; Mougin, K.; Ghorbal, A. Adsorptive removal of cationic and anionic dyes from aqueous solution by utilizing almond shell as bioadsorbent. Euro-Mediterr. J. Environ. Integr. 2017, 2, 20. [Google Scholar] [CrossRef]
  38. Farah, J.Y.; Elgendy, N. Performance, kinetics and equilibrium in biosorption of anionic dye Acid Red 14 by the waste biomass of Saccharomyces cerevisiae as a low-cost biosorbent. Turk. J. Eng. Environ. Sci. 2013, 37, 146–161. [Google Scholar]
  39. Alhujaily, A.; Yu, H.; Zhang, X.; Ma, F. Adsorptive removal of anionic dyes from aqueous solutions using spent mushroom waste. Appl. Water Sci. 2020, 10, 183. [Google Scholar] [CrossRef]
  40. Grabi, H.; Lemlikchi, W.; Derridj, F.; Lemlikchi, S.; Trari, M. Efficient native biosorbent derived from agricultural waste precursor for anionic dye adsorption in synthetic wastewater. Biomass Convers. Biorefinery 2021, 1–18. [Google Scholar] [CrossRef]
  41. Grabi, H.; Derridj, F.; Lemlikchi, W.; Guénin, E. Studies of the potential of a native natural biosorbent for the elimination of an anionic textile dye Cibacron Blue in aqueous solution. Sci. Rep. 2021, 11, 9705. [Google Scholar] [CrossRef]
  42. Banerjee, S.; Dastidar, M. Use of jute processing wastes for treatment of wastewater contaminated with dye and other organics. Bioresour. Technol. 2005, 96, 1919–1928. [Google Scholar] [CrossRef]
  43. Mbarki, F.; Kesraoui, A.; Seffen, M.; Ayrault, P. Kinetic, thermodynamic, and adsorption behavior of cationic and anionic dyes onto corn stigmata: Nonlinear and stochastic analyses. Water Air Soil Pollut. 2018, 229, 95. [Google Scholar] [CrossRef]
  44. Jin, L.; Li, W.; Xu, Q.; Sun, Q. Amino-functionalized nanocrystalline cellulose as an adsorbent for anionic dyes. Cellulose 2015, 22, 2443–2456. [Google Scholar] [CrossRef]
  45. Lyu, W.; Yu, M.; Feng, J.; Yan, W. Highly crystalline polyaniline nanofibers coating with low-cost biomass for easy separation and high efficient removal of anionic dye ARG from aqueous solution. Appl. Surf. Sci. 2018, 458, 413–424. [Google Scholar] [CrossRef]
  46. Zhou, T.; Lu, W.; Liu, L.; Zhu, H.; Jiao, Y.; Zhang, S.; Han, R. Effective adsorption of light green anionic dye from solution by CPB modified peanut in column mode. J. Mol. Liq. 2015, 211, 909–914. [Google Scholar] [CrossRef]
  47. Su, Y.; Zhao, B.; Xiao, W.; Han, R. Adsorption behavior of light green anionic dye using cationic surfactant-modified wheat straw in batch and column mode. Environ. Sci. Pollut. Res. 2013, 20, 5558–5568. [Google Scholar] [CrossRef] [PubMed]
  48. Oei, B.C.; Ibrahim, S.; Wang, S.; Ang, H.M. Surfactant modified barley straw for removal of acid and reactive dyes from aqueous solution. Bioresour. Technol. 2009, 100, 4292–4295. [Google Scholar] [CrossRef] [PubMed]
  49. Ibrahim, S.; Fatimah, I.; Ang, H.-M.; Wang, S. Adsorption of anionic dyes in aqueous solution using chemically modified barley straw. Water Sci. Technol. 2010, 62, 1177–1182. [Google Scholar] [CrossRef] [PubMed]
  50. Ansari, R.; Seyghali, B.; Mohammad-Khah, A.; Zanjanchi, M.A. Highly efficient adsorption of anionic dyes from aqueous solutions using sawdust modified by cationic surfactant of cetyltrimethylammonium bromide. J. Surfactants Deterg. 2012, 15, 557–565. [Google Scholar] [CrossRef]
  51. Namasivayam, C.; Sureshkumar, M. Anionic dye adsorption characteristics of surfactant-modified coir pith, a ‘waste’lignocellulosic polymer. J. Appl. Polym. Sci. 2006, 100, 1538–1546. [Google Scholar] [CrossRef]
  52. Jiang, Z.; Hu, D. Molecular mechanism of anionic dyes adsorption on cationized rice husk cellulose from agricultural wastes. J. Mol. Liq. 2019, 276, 105–114. [Google Scholar] [CrossRef]
  53. Koay, Y.S.; Ahamad, I.S.; Nourouzi, M.M.; Abdullah, L.C.; Choong, T.S.Y. Development of novel low-cost quaternized adsorbent from palm oil agriculture waste for reactive dye removal. BioResources 2014, 9, 66–85. [Google Scholar] [CrossRef]
  54. Benabbas, K.; Zabat, N.; Hocini, I. Study of the chemical pretreatment of a nonconventional low-cost biosorbent (Callitriche obtusangula) for removing an anionic dye from aqueous solution. Euro-Mediterr. J. Environ. Integr. 2021, 6, 54. [Google Scholar] [CrossRef]
  55. Noreen, S.; Bhatti, H.N.; Nausheen, S.; Sadaf, S.; Ashfaq, M. Batch and fixed bed adsorption study for the removal of Drimarine Black CL-B dye from aqueous solution using a lignocellulosic waste: A cost affective adsorbent. Ind. Crops Prod. 2013, 50, 568–579. [Google Scholar] [CrossRef]
  56. Won, S.W.; Kim, H.-J.; Choi, S.-H.; Chung, B.-W.; Kim, K.-J.; Yun, Y.-S. Performance, kinetics and equilibrium in biosorption of anionic dye Reactive Black 5 by the waste biomass of Corynebacterium glutamicum as a low-cost biosorbent. Chem. Eng. J. 2006, 121, 37–43. [Google Scholar] [CrossRef]
  57. Saputra, O.A.; Nauqinida, M.; Pujiasih, S.; Kusumaningsih, T.; Pramono, E. Improvement of anionic and cationic dyes removal in aqueous solution by Indonesian agro-waste oil palm empty fruit bunches through silylation approach. Groundw. Sustain. Dev. 2021, 13, 100570. [Google Scholar] [CrossRef]
  58. Lin, Q.; Wang, K.; Gao, M.; Bai, Y.; Chen, L.; Ma, H. Effectively removal of cationic and anionic dyes by pH-sensitive amphoteric adsorbent derived from agricultural waste-wheat straw. J. Taiwan Inst. Chem. Eng. 2017, 76, 65–72. [Google Scholar] [CrossRef]
  59. Velinov, N.; Radović Vučić, M.; Petrović, M.; Najdanović, S.; Kostić, M.; Mitrović, J.; Bojić, A. The influence of various solvents’ polarity in the synthesis of wood biowaste sorbent: Evaluation of dye sorption. Biomass Convers. Biorefinery 2021, 1–12. [Google Scholar] [CrossRef]
  60. Ahmaruzzaman, M.; Reza, R.A. Decontamination of cationic and anionic dyes in single and binary mode from aqueous phase by mesoporous pulp waste. Environ. Prog. Sustain. Energy 2015, 34, 724–735. [Google Scholar] [CrossRef]
  61. Demiral, H.; Demiral, I.; Karabacakoğlu, B.; Tümsek, F. Adsorption of textile dye onto activated carbon prepared from industrial waste by ZnCl2 activation. J. Int. Environ. Appl. Sci. 2008, 3, 381–389. [Google Scholar]
  62. Bhomick, P.C.; Supong, A.; Baruah, M.; Pongener, C.; Sinha, D. Pine Cone biomass as an efficient precursor for the synthesis of activated biocarbon for adsorption of anionic dye from aqueous solution: Isotherm, kinetic, thermodynamic and regeneration studies. Sustain. Chem. Pharm. 2018, 10, 41–49. [Google Scholar] [CrossRef]
  63. Zhao, B.; Xiao, W.; Shang, Y.; Zhu, H.; Han, R. Adsorption of light green anionic dye using cationic surfactant-modified peanut husk in batch mode. Arab. J. Chem. 2017, 10, S3595–S3602. [Google Scholar] [CrossRef]
  64. Velinov, N.; Mitrović, J.; Kostić, M.; Radović, M.; Petrović, M.; Bojić, D.; Bojić, A. Wood residue reuse for a synthesis of lignocellulosic biosorbent: Characterization and application for simultaneous removal of copper (II), Reactive Blue 19 and cyprodinil from water. Wood Sci. Technol. 2019, 53, 619–647. [Google Scholar] [CrossRef]
  65. Ma, H.; Li, J.-B.; Liu, W.-W.; Miao, M.; Cheng, B.-J.; Zhu, S.-W. Novel synthesis of a versatile magnetic adsorbent derived from corncob for dye removal. Bioresour. Technol. 2015, 190, 13–20. [Google Scholar] [CrossRef] [PubMed]
  66. Markets and Markets, Activated Carbon Market by Type, Application (Liquid Phase (Water Treatment, Foods & Beverages, Pharmaceutical & Medical), Gas Phase (Industrial, Automotive), and Region (APAC, North America, Europe, Middle East, South America)—Global Forecast to 2026. 2022. Available online: https://www.marketsandmarkets.com/Market-Reports/activated-carbon-362.html (accessed on 15 August 2022).
  67. Ao, W.; Fu, J.; Mao, X.; Kang, Q.; Ran, C.; Liu, Y.; Zhang, H.; Gao, Z.; Li, J.; Liu, G. Microwave assisted preparation of activated carbon from biomass: A review. Renew. Sustain. Energy Rev. 2018, 92, 958–979. [Google Scholar] [CrossRef]
  68. Büchel, K.H.; Moretto, H.-H.; Werner, D. Industrial Inorganic Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 2008. [Google Scholar]
  69. 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]
  70. Belyaeva, O.; Krasnova, T.; Semenova, S. Effect of modification of granulated activated carbons with ozone on their properties. Russ. J. Appl. Chem. 2011, 84, 597–601. [Google Scholar] [CrossRef]
  71. Delamar, M.; Desarmot, G.; Fagebaume, O.; Hitmi, R.; Pinsonc, J.; Savéant, J.-M. Modification of carbon fiber surfaces by electrochemical reduction of aryl diazonium salts: Application to carbon epoxy composites. Carbon 1997, 35, 801–807. [Google Scholar] [CrossRef]
  72. García, A.B.; Martínez-Alonso, A.; y Leon, C.A.L.; Tascón, J.M. Modification of the surface properties of an activated carbon by oxygen plasma treatment. Fuel 1998, 77, 613–624. [Google Scholar] [CrossRef]
  73. Otake, Y.; Jenkins, R.G. Characterization of oxygen-containing surface complexes created on a microporous carbon by air and nitric acid treatment. Carbon 1993, 31, 109–121. [Google Scholar] [CrossRef]
  74. Silva, A.R.; Freire, C.; De Castro, B.; Freitas, M.; Figueiredo, J. Anchoring of a nickel (II) Schiff base complex onto activated carbon mediated by cyanuric chloride. Microporous Mesoporous Mater. 2001, 46, 211–221. [Google Scholar] [CrossRef]
  75. Theamwong, N.; Intarabumrung, W.; Sangon, S.; Aintharabunya, S.; Ngernyen, Y.; Hunt, A.J.; Supanchaiyamat, N. Activated carbons from waste Cassia bakeriana seed pods as high-performance adsorbents for toxic anionic dye and ciprofloxacin antibiotic remediation. Bioresour. Technol. 2021, 341, 125832. [Google Scholar] [CrossRef]
  76. Nayak, A.; Bhushan, B.; Gupta, V.; Sharma, P. Chemically activated carbon from lignocellulosic wastes for heavy metal wastewater remediation: Effect of activation conditions. J. Colloid Interface Sci. 2017, 493, 228–240. [Google Scholar] [CrossRef]
  77. Saha, A.; Basak, B.B.; Ponnuchamy, M. Performance of activated carbon derived from Cymbopogon winterianus distillation waste for scavenging of aqueous toxic anionic dye Congo red: Comparison with commercial activated carbon. Sep. Sci. Technol. 2020, 55, 1970–1983. [Google Scholar] [CrossRef]
  78. Alhogbi, B.G.; Altayeb, S.; Bahaidarah, E.; Zawrah, M.F. Removal of anionic and cationic dyes from wastewater using activated carbon from palm tree fiber waste. Processes 2021, 9, 416. [Google Scholar] [CrossRef]
  79. Extross, A.; Waknis, A.; Tagad, C.; Gedam, V.; Pathak, P. Adsorption of congo red using carbon from leaves and stem of water hyacinth: Equilibrium, kinetics, thermodynamic studies. Int. J. Environ. Sci. Technol. 2022, 1–38. [Google Scholar] [CrossRef]
  80. Ravenni, G.; Cafaggi, G.; Sárossy, Z.; Nielsen, K.R.; Ahrenfeldt, J.; Henriksen, U. Waste chars from wood gasification and wastewater sludge pyrolysis compared to commercial activated carbon for the removal of cationic and anionic dyes from aqueous solution. Bioresour. Technol. Rep. 2020, 10, 100421. [Google Scholar] [CrossRef]
  81. Paredes-Laverde, M.; Salamanca, M.; Diaz-Corrales, J.D.; Flórez, E.; Silva-Agredo, J.; Torres-Palma, R.A. Understanding the removal of an anionic dye in textile wastewaters by adsorption on ZnCl2 activated carbons from rice and coffee husk wastes: A combined experimental and theoretical study. J. Environ. Chem. Eng. 2021, 9, 105685. [Google Scholar] [CrossRef]
  82. Güzel, F.; Sayğılı, H.; Sayğılı, G.A.; Koyuncu, F. New low-cost nanoporous carbonaceous adsorbent developed from carob (Ceratonia siliqua) processing industry waste for the adsorption of anionic textile dye: Characterization, equilibrium and kinetic modeling. J. Mol. Liq. 2015, 206, 244–255. [Google Scholar] [CrossRef]
  83. Malik, P.K. Use of activated carbons prepared from sawdust and rice-husk for adsorption of acid dyes: A case study of Acid Yellow 36. Dye. Pigment. 2003, 56, 239–249. [Google Scholar] [CrossRef]
  84. Zubair, M.; Mu’azu, N.D.; Jarrah, N.; Blaisi, N.I.; Aziz, H.A.; A Al-Harthi, M. Adsorption behavior and mechanism of methylene blue, crystal violet, eriochrome black T, and methyl orange dyes onto biochar-derived date palm fronds waste produced at different pyrolysis conditions. Water Air Soil Pollut. 2020, 231, 240. [Google Scholar] [CrossRef]
  85. Periyaraman, P.M.; Karan, S.; Ponnusamy, S.K.; Vaidyanathan, V.; Vasanthakumar, S.; Dhanasekaran, A.; Subramanian, S. Adsorption of an anionic dye onto native and chemically modified agricultural waste. Environ. Eng. Manag. J. 2019, 18, 257–270. [Google Scholar]
  86. Bouhadjra, K.; Lemlikchi, W.; Ferhati, A.; Mignard, S. Enhancing removal efficiency of anionic dye (Cibacron blue) using waste potato peels powder. Sci. Rep. 2021, 11, 2090. [Google Scholar] [CrossRef]
  87. Sayğılı, H.; Güzel, F.; Önal, Y. Conversion of grape industrial processing waste to activated carbon sorbent and its performance in cationic and anionic dyes adsorption. J. Clean. Prod. 2015, 93, 84–93. [Google Scholar] [CrossRef]
  88. Mittal, A.; Mittal, J.; Malviya, A.; Gupta, V. Adsorptive removal of hazardous anionic dye “Congo red” from wastewater using waste materials and recovery by desorption. J. Colloid Interface Sci. 2009, 340, 16–26. [Google Scholar] [CrossRef]
  89. Landin-Sandoval, V.; Mendoza-Castillo, D.; Seliem, M.; Mobarak, M.; Villanueva-Mejia, F.; Bonilla-Petriciolet, A.; Navarro-Santos, P.; Reynel-Ávila, H. Physicochemical analysis of multilayer adsorption mechanism of anionic dyes on lignocellulosic biomasses via statistical physics and density functional theory. J. Mol. Liq. 2021, 322, 114511. [Google Scholar] [CrossRef]
  90. Demirbas, E.; Nas, M. Batch kinetic and equilibrium studies of adsorption of Reactive Blue 21 by fly ash and sepiolite. Desalination 2009, 243, 8–21. [Google Scholar] [CrossRef]
  91. Vadivelan, V.; Kumar, K.V. Equilibrium, kinetics, mechanism, and process design for the sorption of methylene blue onto rice husk. J. Colloid Interface Sci. 2005, 286, 90–100. [Google Scholar] [CrossRef]
  92. Tanzifi, M.; Yaraki, M.T.; Kiadehi, A.D.; Hosseini, S.H.; Olazar, M.; Bharti, A.K.; Agarwal, S.; Gupta, V.K.; Kazemi, A. Adsorption of Amido Black 10B from aqueous solution using polyaniline/SiO2 nanocomposite: Experimental investigation and artificial neural network modeling. J. Colloid Interface Sci. 2018, 510, 246–261. [Google Scholar] [CrossRef] [PubMed]
  93. Senthil Kumar, P.; Palaniyappan, M.; Priyadharshini, M.; Vignesh, A.; Thanjiappan, A.; Sebastina Anne Fernando, P.; Tanvir Ahmed, R.; Srinath, R. Adsorption of basic dye onto raw and surface-modified agricultural waste. Environ. Prog. Sustain. Energy 2014, 33, 87–98. [Google Scholar] [CrossRef]
  94. Iakovleva, E.; Sillanpää, M.; Maydannik, P.; Liu, J.T.; Allen, S.; Albadarin, A.B.; Mangwandi, C. Manufacturing of novel low-cost adsorbent: Co-granulation of limestone and coffee waste. J. Environ. Manag. 2017, 203, 853–860. [Google Scholar] [CrossRef] [PubMed]
  95. Batool, F.; Akbar, J.; Iqbal, S.; Noreen, S.; Bukhari, S.N.A. Study of isothermal, kinetic, and thermodynamic parameters for adsorption of cadmium: An overview of linear and nonlinear approach and error analysis. Bioinorg. Chem. Appl. 2018, 2018, 3463724. [Google Scholar] [CrossRef]
  96. Dada, A.O.; Latona, D.F.; Ojediran, O.J.; Nath, O.O. Adsorption of Cu (II) onto bamboo supported manganese (BS-Mn) nanocomposite: Effect of operational parameters, kinetic, isotherms, and thermodynamic studies. J. Appl. Sci. Environ. Manag. 2016, 20, 409–422. [Google Scholar] [CrossRef]
  97. Kezerle, A.; Velic, N.; Hasenay, D.; Kovacevic, D. Lignocellulosic materials as dye adsorbents: Adsorption of methylene blue and Congo red on Brewers’ spent grain. Croat. Chem. Acta 2018, 91, 53–65. [Google Scholar] [CrossRef]
  98. Farzana, N.; Uddin, M.Z.; Haque, M.M.; Haque, A.N.M.A. Dyeability, kinetics and physico-chemical aspects of Bombyx mori muslin silk fabric with bi-functional reactive dyes. J. Nat. Fibers 2018, 17, 986–1000. [Google Scholar] [CrossRef]
  99. Haque, A.N.M.A.; Hussain, M.; Smriti, S.A.; Siddiqa, F.; Farzana, N. Kinetic Study of Curcumin on Modal Fabric. Tekstilec 2018, 61, 27–32. [Google Scholar] [CrossRef]
  100. Haque, A.N.M.A.; Hussain, M.; Siddiqa, F.; Haque, M.; Islam, G. Adsorption Kinetics of Curcumin on Cotton Fabric. Tekstilec 2018, 61, 76–81. [Google Scholar] [CrossRef]
  101. Hubbe, M.A.; Azizian, S.; Douven, S. Implications of apparent pseudo-second-order adsorption kinetics onto cellulosic materials: A review. BioResources 2019, 14, 7582–7626. [Google Scholar] [CrossRef]
  102. El-Khaiary, M.I.; Malash, G.F.; Ho, Y.-S. On the use of linearized pseudo-second-order kinetic equations for modeling adsorption systems. Desalination 2010, 257, 93–101. [Google Scholar] [CrossRef]
  103. Ho, Y.-S. Second-order kinetic model for the sorption of cadmium onto tree fern: A comparison of linear and non-linear methods. Water Res. 2006, 40, 119–125. [Google Scholar] [CrossRef]
  104. Ho, Y.-S. Isotherms for the sorption of lead onto peat: Comparison of linear and non-linear methods. Pol. J. Environ. Stud. 2006, 15, 81–86. [Google Scholar]
  105. Perez-Ameneiro, M.; Vecino, X.; Cruz, J.; Moldes, A. Wastewater treatment enhancement by applying a lipopeptide biosurfactant to a lignocellulosic biocomposite. Carbohydr. Polym. 2015, 131, 186–196. [Google Scholar] [CrossRef]
  106. Islam, G.N.; Ke, G.; Haque, A.N.M.A.; Islam, A. Effect of ultrasound on dyeing of wool fabric with acid dye. Int. J. Ind. Chem. 2017, 8, 425–431. [Google Scholar] [CrossRef]
  107. Roy, A.; Adhikari, B.; Majumder, S. Equilibrium, kinetic, and thermodynamic studies of azo dye adsorption from aqueous solution by chemically modified lignocellulosic jute fiber. Ind. Eng. Chem. Res. 2013, 52, 6502–6512. [Google Scholar] [CrossRef]
  108. Haque, A.N.M.A.; Farzana, N.; Smriti, S.; Islam, M. Kinetics and thermodynamics of silk dyeing with turmeric extract. AATCC J. Res. 2018, 5, 8–14. [Google Scholar] [CrossRef]
  109. Ibrahim, S.M.; Badawy, A.A.; Essawy, H.A. Improvement of dyes removal from aqueous solution by Nanosized cobalt ferrite treated with humic acid during coprecipitation. J. Nanostructure Chem. 2019, 9, 281–298. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The number of studies retrieved from the Web of Science and Scopus databases: (a) on ‘anionic dye removal’ since 2015, and (b) their distribution across ten key subject categories.
Figure 1. The number of studies retrieved from the Web of Science and Scopus databases: (a) on ‘anionic dye removal’ since 2015, and (b) their distribution across ten key subject categories.
Sustainability 14 11098 g001
Figure 2. Ion-based-dye classification with examples.
Figure 2. Ion-based-dye classification with examples.
Sustainability 14 11098 g002
Figure 3. Common chemical groups in plant-derived agricultural waste.
Figure 3. Common chemical groups in plant-derived agricultural waste.
Sustainability 14 11098 g003
Figure 4. Schematic of the key concept of the mechanical modification of plant-derived agricultural waste for anionic dye removal.
Figure 4. Schematic of the key concept of the mechanical modification of plant-derived agricultural waste for anionic dye removal.
Sustainability 14 11098 g004
Figure 5. Different chemical-treatment methods are proposed for modifying plant-derived agricultural waste.
Figure 5. Different chemical-treatment methods are proposed for modifying plant-derived agricultural waste.
Sustainability 14 11098 g005
Figure 6. Schematic of thermal-modification steps for plant-derived agricultural waste towards the preparation of activated carbon for anionic dye removal.
Figure 6. Schematic of thermal-modification steps for plant-derived agricultural waste towards the preparation of activated carbon for anionic dye removal.
Sustainability 14 11098 g006
Figure 7. Schematic representations of: (a) different interactions between dye particles and lignocellulosic biomass; (b) diffusion mechanism of dye particles during an adsorption process by a biomass-derived adsorbent.
Figure 7. Schematic representations of: (a) different interactions between dye particles and lignocellulosic biomass; (b) diffusion mechanism of dye particles during an adsorption process by a biomass-derived adsorbent.
Sustainability 14 11098 g007
Figure 8. An example of the alteration of the fitting value when graphs were plotted using two different pseudo-second-order equations, showing: (a) good fit, and (b) poor fit from the same data.
Figure 8. An example of the alteration of the fitting value when graphs were plotted using two different pseudo-second-order equations, showing: (a) good fit, and (b) poor fit from the same data.
Sustainability 14 11098 g008
Table 1. Maximum adsorption capacities (qmax) of anionic dye adsorbents derived from mechanically modified agricultural wastes.
Table 1. Maximum adsorption capacities (qmax) of anionic dye adsorbents derived from mechanically modified agricultural wastes.
ResourcesParticle Size (µm)Dyeqmax (mg/g)Reference
Waste
tea residue
0.3475Acid Blue 25127.14[31]
Date stones and jujube shells 50–100Congo Red45.08–59.55[32]
Waste banana pith>53Direct Red5.92[33]
Waste banana pith>53Acid
Brilliant Blue
4.42[33]
Jujuba seeds53–150Congo Red55.56[34]
Cotton plant wastes75–500Remazol Black
B
35.7–50.9[35]
Lotus<100Congo Red0.783–1.179[36]
Almond shell100–500Eriochrome Black T123.92[37]
Waste of corn silk250–500Reactive Blue 1971.6[27]
Waste of corn silk250–500Reactive Red 21863.3[27]
Banana peel, cucumber peel, and potato peel250–500Orange G20.9–40.5[30]
Eucalyptus bark250–700Solar Red BA43.5[17]
Eucalyptus bark250–700Solar Brittle Blue A49[17]
Saccharomyces cerevisiae (yeast) 315–400Acid Red 1418–23[38]
Mushroom waste<400Direct Red 5B18[39]
Mushroom waste<400Direct Black 2215.46[39]
Mushroom waste<400Direct Black 7120.19[39]
Mushroom waste<400Reactive Black 514.62[39]
Ash seed≤1000Cibacron Blue67.114[40]
Bean peel≤1000Cibacron Blue28.490[41]
Jute processing waste10,000Congo Red13.18[39,42]
Corn stigmataGround (size not mentioned)Indigo Carmine63.7[39,43]
Cotton gin trashFilm formAcid Blue 2535.46[7]
Table 3. Maximum anionic dye adsorptions (qmax) reported in different studies onto activated carbon prepared from agricultural waste.
Table 3. Maximum anionic dye adsorptions (qmax) reported in different studies onto activated carbon prepared from agricultural waste.
ResourcesSurface Area (m2/g)Anionic Dyeqmax (mg/g)Reference
Java citronella Not reportedCongo Red4.29[77]
Palm tree waste648.90Congo Red10.4[78]
Water hyacinthNot reportedCongo Red14.367[79]
Wastewater sludge98.8Amaranth19.6[80]
Coffee husk613 ± 14Indigo Carmine36.63[81]
Carob waste921.07Reactive Black 536.90[82]
Rick husk272Acid Yellow 3686.9[83]
Pinecone878.07Alizarin Red S118.06[62]
Vegetable wasteNot reportedEriochrome Black T120.50[28]
Date palm fronds431.82Methyl Orange163.132[84]
Sawdust516.3Acid Yellow 36183.8[83]
Hazelnut bagasse1489Acid Blue 350450[61]
Psyllium stalksNot reportedCoomassie Brilliant Blue237.2[83,85]
Waste potato peels Not reportedCibacron Blue270.3[83,86]
Pulp waste1022.46Methyl Orange285.71[60,85]
Date palm fronds431.82Eriochrome Black T309.59[84]
Grape waste1455Acid yellow 36386[84,87]
Pink shower 283.4Congo Red970[75]
Table 4. Reported R2 values of Langmuir and Freundlich isotherm models for anionic dye adsorption.
Table 4. Reported R2 values of Langmuir and Freundlich isotherm models for anionic dye adsorption.
Modification TechniqueDyeAdsorbentR2Reference
Langmuir ModelFreundlich Model
MechanicalCongo RedJujuba seeds0.9990.987–0.989[34]
MechanicalReactive
Black 5
Spent
mushroom
waste
0.9970.907[39]
MechanicalCibacron
Blue
Ash seed0.93570.98[40]
MechanicalCibacron
Blue
Bean peel0.98220.9414[41]
MechanicalDirect Red
5B
Spent
mushroom
waste
0.9980.918[39]
MechanicalDirect
Black 22
Spent
mushroom
waste
0.9960.905[39]
MechanicalAcid Blue
25
Waste
tea residue
0.9197–0.9309Not reported[31]
MechanicalCongo RedBottom ash and
deoiled soya
0.79–0.990.85–0.97[88]
MechanicalCongo RedJute processing
Waste
0.99160.9912[42]
ChemicalCongo RedSawdust0.9418–0.98670.9040–0.9876[50]
ChemicalCongo RedRice husk0.9939–0.99880.9672–0.9861[52]
ChemicalCongo RedPalm tree fiber waste0.9960.857[78]
ChemicalCongo RedCoffee
Waste
0.990.67[18]
ChemicalDiamine Green BRice husk0.9209–0.97310.8085–0.8988[52]
ChemicalReactive
Black 5
Palm kernel
Shell
0.9010.854[53]
ChemicalReactive
Black 5
Coffee
Waste
0.960.93[18]
ChemicalReactive Black
5
Fermentation
Wastes
0.994–0.999Not reported[56]
ChemicalAcid Blue
25
Cotton gin trash0.967–0.9960.875–0.947[7]
ChemicalDirect
Red 89
Aquatic
Plant
0.9950.999[54]
ChemicalAcid
Black
24
Rice husk0.9887–0.99750.9715–0.9823[52]
ThermalEriochrome Black TDate palm
Fronds
0.694–0.9530.711–0.975[84]
ThermalCongo RedJava
Citronella
0.904–0.9570.994–0.999[77]
ThermalEriochrome Black TVegetable
Waste
0.853–0.9780.965–0.996[28]
Table 5. The trends of the enthalpy (ΔH°), Gibbs energy of activation (ΔG°), and entropy (ΔS°) of anionic dye adsorption onto plant-derived agricultural waste.
Table 5. The trends of the enthalpy (ΔH°), Gibbs energy of activation (ΔG°), and entropy (ΔS°) of anionic dye adsorption onto plant-derived agricultural waste.
AdsorbentModification TechniqueDyeΔH° (kJ/mol)ΔG° (kJ/mol) ΔS° (J/mol K)References
Bean peelMechanicalCibacron Blue−32.36−3.97 to −4.89 −92.21[41]
Ash seedMechanicalCibacron Blue−28.95−18.01 to −18.37−35.36[40]
Spent
mushroom
waste
MechanicalDirect Red 5B0.99−0.79 to −1.831.63[39]
Spent
mushroom
waste
MechanicalDirect Black 220.59−0.69 to −1.331.65[39]
Spent
mushroom
waste
MechanicalDirect Black 713.79−3.32 to −7.477.02[39]
Spent
mushroom
waste
MechanicalReactive Black 50.62−0.15 to −0.770.21[39]
Vegetable
waste
MechanicalEriochrome Black T44.51−2.72 to −5.89158.51
Jujuba seedsMechanicalCongo Red12.94−3.49 to −6.4357.9[34]
SawdustChemicalAcid Red G28.7−1.07 to −3.14103.4[45]
Orange peelChemicalCongo Red19−1.88 to −3.2370[16]
Peanut huskChemicalDrimarine Black CL-B−23.74−0.22 to −2.972[55]
Coffee wasteChemicalReactive Black 58.28−9.66 to −11.0859.99[18]
Coffee wasteChemicalCongo Red35.05−3.76 to −8.17130.59[18]
Rice huskChemicalDiamine Green B15.06−3.46 to −4.3935.11[52]
Rice huskChemicalAcid Black 2416.32−2.65 to −3.1043.52[52]
Rice huskChemicalCongo Red2.96−2.98 to −3.3719.57[52]
Wheat strawChemicalReactive Red 244.53−29.05 to −32.93135.1[12]
Waste coir
pith
ChemicalProcion Orange25.69−8.43 to −11.15111.52[51]
Waste coir
pith
ChemicalAcid Brilliant Blue70.09−5.65 to −11.37245.64[51]
Water
hyacinth
ThermalCongo Red−189.01 to −918.94−3001.04 to −12,254.116.39 to 53.97[79]
Coffee huskThermalIndigo Carmine21.72 to 35.83−46.71 to −61.05229.9 to 287.6[81]
Java
citronella
ThermalCongo Red−0.007 to −0.0297.9 to 30.2−6 to −92[77]
PineconeThermalAlizarin Red S29.13−2.75 to −4.3982[62]
Vegetable
waste
ThermalEriochrome Black T−10.08−2.76 to 3.43−22.32[28]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Haque, A.N.M.A.; Sultana, N.; Sayem, A.S.M.; Smriti, S.A. Sustainable Adsorbents from Plant-Derived Agricultural Wastes for Anionic Dye Removal: A Review. Sustainability 2022, 14, 11098. https://doi.org/10.3390/su141711098

AMA Style

Haque ANMA, Sultana N, Sayem ASM, Smriti SA. Sustainable Adsorbents from Plant-Derived Agricultural Wastes for Anionic Dye Removal: A Review. Sustainability. 2022; 14(17):11098. https://doi.org/10.3390/su141711098

Chicago/Turabian Style

Haque, Abu Naser Md Ahsanul, Nigar Sultana, Abu Sadat Muhammad Sayem, and Shamima Akter Smriti. 2022. "Sustainable Adsorbents from Plant-Derived Agricultural Wastes for Anionic Dye Removal: A Review" Sustainability 14, no. 17: 11098. https://doi.org/10.3390/su141711098

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