Next Article in Journal
Quantitative Analysis of Anthocyanins in Grapes by UPLC-Q-TOF MS Combined with QAMS
Previous Article in Journal
Phytochemical Characterization and Heavy Metal and Thermal Analyses of Saussurea hypoleuca Root and Evaluation of Its Anthelmintic and Antioxidant Activity In Vitro and In Silico
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of the Removal of Dyestuffs from Effluents onto Biochar

1
Division of Sustainable Development, College of Science and Engineering, Hamad Bin Khalifa University, Qatar Foundation, Doha 5825, Qatar
2
Centre for Advanced Materials, Qatar University, Doha 2713, Qatar
*
Author to whom correspondence should be addressed.
Separations 2022, 9(6), 139; https://doi.org/10.3390/separations9060139
Submission received: 8 May 2022 / Revised: 26 May 2022 / Accepted: 27 May 2022 / Published: 30 May 2022

Abstract

:
The study provides a review of various applications of biomass-derived biochars, waste-derived biochars, and modified biochars as adsorbent materials for removing dyestuff from process effluents. Processing significant amounts of dye effluent discharges into receiving waters can supply major benefits to countries which are affected by the water crisis and anticipated future stress in many areas in the world. When compared to most conventional adsorbents, biochars can provide an economically attractive solution. In comparison to many other textile effluent treatment processes, adsorption technology provides an economic, easily managed, and highly effective treatment option. Several tabulated data values are provided that summarize the main characteristics of various biochar adsorbents according to their ability to remove dyestuffs from wastewaters.

1. Introduction

Dyestuffs color and pollute receiving waters, streams, and rivers as a result of inadequate processing of the industrial effluents by a variety of industrial applications including the food and beverage companies, paper and pulp processing, paint manufacturing, pharmaceutical processing, printing, textiles, dyeing, and printing [1]. Many dyes pose a grave danger to the water environmental ecosystem due to their chemical properties, with serious consequences for human health, animal, and plant ecosystems [2,3]. Aside from a limited number of studies indicating that specific dyes are toxic, the presence of dyestuffs into receiving waters reduces the photosynthetic process by inhibiting light from passing through [4]. During the degradation process, dyes consume the dissolved oxygen concentrations of the receiving water, therefore decreasing the water quality standards for aquatic species. This has detrimental visual aesthetic impacts which may result in health reproductive issues in fishes [5]. Specific dyes have a negative impact on the skin, kidneys, liver, reproductive system, heart, brain, and nervous system, and some may be carcinogenic or mutagenic.
Data on dyestuff effluent discharge volumes and production quantities are not readily available or recorded around the world. According to available data, 700,100 tons of dyestuffs are produced every year for 10,000 dyes. According to industry figures, the global dyestuffs produced yearly is 1.8 to 1.9 × 106 tons with more than 11,000 dye pigments applied primarily in the food, textile, cosmetics, leather, paper, and plastics industries [6]. Depending on the type of dyestuff and the process technology used, 1–10% of dye is not used in the dyeing process, indicating that significant amounts of dye are discharged to the water bodies via various means [7].
The majority of dyestuffs have specific characteristics such as chemically stable and light fastness [8]. Furthermore, the dye color reduces light penetration in streams and rivers, therefore reducing photosynthesis and dissolved oxygen content. They prevent a variety of chemical functions based on the material to which they are applied and the color they impart (Figure 1). All of these properties are advantageous to the dye user and are enhanced by dyestuff manufacturers. However, the huge volumes of effluent makes the treatment of these dyes to comply with environmental effluent discharge standards very problematic. In addition, water can be colored in certain cases with dye concentrations as little as 1 ppm. The majority of dyeing applications employ copious amount of water during the dyeing, washing, and rinsing stages [9,10].
The majority of recent review publications have focused on the biochar manufacturing process and dye removal by adsorption. However, there is a lack of data on the classification of dyes and their properties, as well as the removal of such dyes using biochar as an adsorbent material. This review examines the classification and qualities of dyestuffs in this setting. This article provides a rapid overview of various wastewater treatment systems, followed by a detailed explanation of the benefits of using adsorption technology. The key properties and applications of standard and modified (altered) biochars for color removal are also highlighted in the review. The investigations on the research reported in the directions of biochar modification and application marked a step toward the practical application of biochar.

2. Dye Classifications

Dyes are colored molecules or ions that can be applied to a wide range of materials including food, beverages, and textiles in solution or as a dispersion. Most dyes have a high-water solubility, and often contain a sulfonic acid group, usually in the form of a sodium salt, which is responsible for the solubility of many water-soluble dyestuffs [11]. Dye colors are created by chemical groups absorbing light of various wavelengths in the visible region of the spectrum. Different unsaturated chemical groups on chromophores promote this key distinguishing property. Figure 2 depicts the more common ones.
Auxochromes are groups that can also enhance the water solubility and improve the dye absorption potential for adsorbing material; examples include substituted sulfonic, hydroxyl, carbonyl, or amino groups. Dyes can be classified based on their chemistry or their types of application. As a result, the chemical structure and type of dye must be a primary consideration in determining which dye wastewater process treatment technology should be applied for effluent removal, as well as determining what adsorbent properties are required for the adsorption of the specific dyestuff type.

2.1. Reactive Dyes

Reactive dyes are used extensively in the dyeing of cellulosic textile fibers, namely, flax and cotton. Due to their high adhesion to a substrate, they can also be used to dye linen, viscose, and silk [12,13]. These reactive compounds in the dye can form chemical bonds with textile uptake of fibers. The uptake of the dichlorotriazine type of reactive dye which becomes attached to the cellulose fiber by displacing the chloride grouping is depicted in the mechanistic schemes below. One or both chlorides may be present. Figure 3 and Figure 4 show the typical dye uptake mechanisms for dyeing cellulosic materials.
Due to reactive dyes having a strong bonding affinity for cellulose, consequently, the hydroxyl group containing biosorbents have demonstrated a very strong dye uptake capacity to remove reactive dye compounds from textile dyehouse effluents [14].

2.2. Disperse Dyes

Disperse dyes are non-ionic substances that are commonly applied to polyesters but can be used in acetate or nylon fabrics. These dyes are water soluble and can be used for these fibers by diffusion into the fibers at increased temperatures. As there are no basic chemical groups, there are no attractive sites for acid dye groups, despite a weak attraction for basic dyes. The dye attachment mechanism is based on weak Van der Waals forces and dipole-dipole interactions, implying that like mechanisms may occur during the removal of disperse dyes onto biochar adsorbents [15]. Figure 5 depicts disperse blue 6 as an example of this class.
The dispersed dyestuff occurs typically as a fine suspension that can be filtered from the effluent discharge by biochar.

2.3. Vat Dyes

The majority of vat dyes have a ketonic style chromophore which may be applied to color cellulosic fibers and materials such as viscose, cotton, and linen. This is a broad category of dyes that includes indanthrones, anthraquinones, carbazoles, benzanthrones, polycyclic quinones, and acridones.
Figure 6 shows the structures of a typical vat dye [16].
The large anthraquinone groups suggest that the removal process may involve electron clouds of the anthraquinone dye by adsorption onto the positively charged surface groups and adequately sized pore diameters of biochars.

2.4. Direct Dyes

Direct dyes or substantive dyes, as there is no fixation phase necessary, may be applied to color cotton yarn, viscose, and loose cotton of fabrics [17]. Mordant chemicals, such as chromium compounds that can undergo complexation by attaching substrate to chromophore to form an insoluble color, are used in some direct dyes, but not all, to fix the dye and improve color fastness. In the case of dark color shades such as black or navy blue dyes, this technique has proven to be cost-effective in achieving high color fastness. These dyes are now being reviewed due to environment and safety concerns which have limited their use. The mechanism for the application of direct dyes involves establishing non-ionic forces to attach the dyestuff to the textile fiber material [18]. The structure of direct yellow 24 dye is depicted in Figure 7.
Direct dyestuffs produce negatively charged ions in solution and can be adsorbed onto positive sites on biochars.

2.5. Basic Dye

Congo red (CR) is a member of the very large group of basic dyes that are characterized by the color, high tinctorial strength, and brilliance. These basic dyes are most commonly used on acrylic fibers, but they can also be used on other fibrous textiles when mordants are used. Furthermore, basic dyes are soluble media but not soluble in alkaline solutions. These dyes are primarily made up of imino or amino groups that are linked to triarylmethane or xanthene; they are also used in typewriter ribbon, carbon paper, and inks [19]. Monoazo, methane, and oxazine are the three main subclasses. Figure 8 depicts the structure of a basic CR dye.
Since most basic dyes ionize in water, they have a positively charged colored ion or cationic species. These dyes are commonly classified as cationic dyes and adsorb most effectively onto negatively charged functional groups on the biochars [20].

2.6. Acid Dye

Acid dyestuffs possess a high color fastness and have a strong affinity for protein molecules, therefore, they are ideal for nylon, and wool, as well as silk. The majority of acidic dyestuffs are sodium salts based on sulfonic acid of organic species [21]. They are distinguished by various groupings of unsaturated aromatic rings known as the chromophore [22]. Concentrations of acidic dyes, even as low as 1 ppm, in the effluent raise the chemical oxygen demand (COD) level of the receiving waterbody and cause a visual disturbance to the environment [23]. Several dyes (i.e., arylamine-based types) are hazardous as they are carcinogenic or toxic [24]. The anthraquinone group is a major group of acid dyes. Figure 9 shows an example of the acid dye.
As the colored section is the anion of a sodium salt, acid dyes are frequently referred to as anionic dyes, possessing a colored ion based on sulfonic acid groups or, possibly, a carboxylic acid group which has the capability for adsorbing onto positive biochar sites.

2.7. Azo Dyes

Azo dyes account for over half the world’s yearly dye production. They are broadly utilized in different industries, including food processing, pharmaceutical, leather, cosmetic, and textile dyeing applications. As shown in Figure 10, the azo dyes possess one or more azo functional groups (-N=N-) as the chromophoric species and are very frequently occurring in synthetic dyes with the aromatic conjugate ring structure [25]. When azo dyes are absorbed, they are metabolized by the intestinal microflora of the liver, resulting in the formation of aromatic amines which are a health hazard; furthermore, the azo dyes themselves, if water soluble, may become carcinogens.
Consequently, a number of countries have enacted legislation governing the manufacture, handling, and use of azo dyestuffs. The implemented legislation for azo dyes has been divided into two categories (i) those capable of producing metabolic carcinogens, and (ii) those not capable of producing carcinogenic species [26].

2.8. Sulfur Dyes

Sulfur dyes are widely produced and used in cotton-related industries because they are simple to apply, possess high color fastness, and are inexpensive. At room temperature, the dyes are insoluble in water; however, in an alkaline pH in the presence of a reducing agent and high temperature, these dyes become water soluble, allowing them to be adsorbed onto and thus dyeing the fabric. Environmentally, sulfur containing dyes are very polluting because of dye effluents and, as a result, are steadily losing favor [27]. Figure 11 depicts a typical sulfur dye.
Due to there being an abundance of lone pairs of electrons due to oxygen, sulfur, and nitrogen, sulfur dyes can be adsorbed onto the positive surface sites, especially in the case of chelating groups acquiring lone electron pairs.

2.9. Aniline Dyes

The aniline class of synthetic dyes are predominantly made from coal tar that contains one or several phenyl groups. Around them, a new dye production industry was developed many years ago, manufacturing several of the most widely applied and well-known dyestuffs, such as malachite green (MG) and CV. The dye is found in the form of fine powders. Aniline dyestuffs are soluble in water and are produced in a variety of colors suitable for wood, leather, and fabric [28]. Some aniline dyestuffs have recently been linked to an increased cancer risk factor, according to research. Figure 12 depicts a typical aniline dye.
As these dyestuffs are a cationic species [20], they can adsorb well on negative biochar groups. Additionally, electron clouds on aromatic ring species and nitrogen lone electron pairs are drawn to the positive sites on biochars.

2.10. Metal Complex Dyes

Metal complex dyes, also known as premetallized dyes, are insoluble in water. Such dyes are primarily monoazo compounds that are used to color wool, polyamides, silk, and nylon. The typical structure contains monoazo basic functional groups with some attached hydroxyl, amino, and carboxyl groups, giving these dye compounds a strong bond-forming capacity as well as the ability to produce coordination compounds with various transition metals such as cobalt, chromium, nickel, and copper. However, because their effluent discharges contain unfixed dye and their associated metal species, these metal complex dyes are now the cause of health, environmental, and safety concerns [15]. Figure 13 depicts the structure of a dithiolene metal complex dye.
These dyes can ionize in water, forming colored anionic species that may be adsorbed onto positive biochar sites; however, removing the toxic heavy metal counter ion, which requires negative sites on the biochar adsorbent is of greater concern.

2.11. Mordant Dyes

Most naturally occurring dyes cannot attach and fix strongly to many textile fibers and must be combined with a mordant compound that binds to both dye and the fiber. This can be used on both the fiber and the dye itself and it can be applied to both wood and nylon. This mordant substance creates a bond between the fiber and the dye molecule [29], however, other dyes have begun to replace mordants in recent years, in particular, direct dyes. The structure of mordant red 11 is depicted in Figure 14.
Mordant dyes ionize in solution, producing colored anionic species, necessitating adsorbing treatment with positive functional sites on the biochar.

3. Dye Removal Technologies

A variety of treatment processes are used in typical dyehouse effluent treatment technologies. These are generally classified into three types: chemical, biological, and physical processes. These technologies can be employed separately or collectively to achieve the best balance of multiple criteria including economic feasibility and process efficiency.

3.1. Physical Process Treatments

Adsorption [30] and biosorption [31] mechanisms describe the method of attraction of dyes onto the periphery of adsorbents in the form of particulates, granules or powders by physical bonding, strong chemical affinities, Van der Waals forces, or exchange of ions with surface functional sites on the adsorbent material surfaces via chemisorption or high energy chemical affinities.
To remove pollutants and colors from wastewater, flocculation and coagulation have long been utilized. This involves adding chemicals to dirty wastewater, often with agitation, and then permitting the pollutant-laden coagulants/flocculants to settle at the vessel bottom. Additional tertiary treatment or recycling may be applied to the refined water overflow that is discharged into water streams. The dye is adsorbed on the periphery/surface of the flocculant/coagulant without undergoing any chemical decomposition, similar to adsorption. Two typically utilized materials are ferrous sulfate and alum [32]. A high settling time for cleared water, the large amount of dye-laden flocculant/coagulant produced and requiring dumping, and the economics involved with the drying of highly wet slurry preceding landfill dumping are all disadvantages.
Electrokinetic coagulation and irradiation [33] are two other alternatives. The use of aluminum electrodes in electrocoagulation has been extensively studied. They unleash charged aluminum ions in solution [34], resulting in additional active or charged adsorption sites. However, they also have the drawbacks of coagulation. Catalytic irradiation/irradiation for removal of color has been researched in some circumstances. The efficacy of methylene blue (MB) removal was examined. The removal effectiveness for the maximum involvement of Pd catalyst reached around 86.4 percent after 120 min of irradiation [35], showing a five-fold improvement in removal rate. The two procedures were highly effective in removing some dyes but less so at removing others [36]. A more detailed examination into these procedures is required in order to verify them.
Membrane filtration is the process of physically separating soluble colors in water effluents using permeable membranes under pressure [37]. The techniques such as ultrafiltration and nanofiltration are also gaining popularity [38,39]. The aforementioned traditional pathways are analogous to membrane technology. Osmosis (forward and reverse) [39,40,41] and filtration (ultra/nano) [40,41] may both separate molecules larger than 1 nm. The high cost of pumping energy and the likelihood of membrane pore obstruction are two of membranes’ major flaws. In case the dye effluents were previously employed for dyeing clothing/textiles, textile threads and fine particles are the ones that are commonly cleared along with the spent liquid. A pre-filtration stage probably is advantageous in these situations to prevent blockage and early membrane fouling [42]. If a condensed dye coat formed, the net process efficiency drops as the flow rate is reduced. Furthermore, the expense of filter media becomes increasingly significant as the unit’s capacity increases. Several recent researches [43,44] have employed photocatalysis and ultrasonics to increase dye adsorption on solid particulates, particularly when magnetic biochars and nanoparticles were used. One drawback of utilizing photocatalytic processes to remove dye at full scale is the preparation of the photocatalysts, which can significantly change the cost of the process, whereas ultrasound treatment has the disadvantage of consuming very high amounts of energy.

3.2. Chemical Process Treatments

Ultraviolet light degradation has little effect on dyes. Most chemical techniques for the treatment of dye-polluted streams seek to knock down the intricate dye particles into simpler and less hazardous chemicals such as CO2 or H2O through a sequence of oxidative reactions. For example, chlorine, wet air oxidation (WAO), ozonation, and other advanced oxidation processes (AOP).
Although WAO [45] is found to be efficient in the treatment of dye effluents, catalysts have been suggested as a way to improve process effectiveness [46,47]. Ozone is commonly employed in commercial AOP’s for the elimination of organics and color [48]. Ozonation catalysts are being investigated [49]. Other powerful oxidizing reagent chemicals viz. sodium hypochlorite, chlorine, and hydrogen peroxide are being utilized commercially in the oxidation of dye-polluted streams [50,51]. These reagents break down dyes into simpler compounds [52]. Nonetheless, there are evidences that the benefits of utilizing chlorinated reagents may be exceeded by the production of harmful byproducts (such as chloroanilines) [53]. Oxidized byproducts and bromoforms are also of concern and are extensively used in the area of ozonation [54], and several studies have investigated their toxic characteristics [55].
For dye removal from wastewater effluents, several catalysts such as TiO2 or Fenton’s reagents have been used as an alternate to typical oxidizers in AOP [56]. Through the application of biochar as a support [57,58] and, more recently, co-catalytic nanoparticles/carbon dots [59,60] as co-catalysts, TiO2-based photocatalysts can improve dye removal performances by enabling synergistic adsorption and degradation. The AOP technique has been shown to have a high dye removal efficiency [61]. However, one significant disadvantage is the intricacy of the resulting ferrous sludge [62]. Many AOP procedures are in place. It is possible to remove up to 90% of the dye; however, a suitable pH level balance and the sludge treatment process must be considered. When using typical titanium dioxide for removing colors, photochemical deterioration [63] is a continuous subject of attention, but it is limited by the technique. Electrochemical treatment was applied to commercial azo dye effluents [64].

3.3. Biological-Based Processes

Traditional aerobic techniques viz. activated sludge has proved successful in treating polluted streams, nonetheless, they are unsuccessful at removing colors. Dye compounds are frequently hydrophilic and have a low affinity for biomass, resulting in early application [65]. Dyes are frequently resistant to microbial breakdown due to their poisonous and stable molecular structures. Dye molecules are broken down into soluble organics in the absence of oxygen, then to CH4 and H2S by a sequence of microbial processes. Although reasonable results for reactive, azo, and diazo dye decolorization were documented [66], the procedure had no effect on hazardous aromatic amines in the treated stream [67].

4. Adsorption Process Technologies

Adsorption is a popular process for purification purposes at large scales. Various contaminants in solution and in gaseous emission streams may be removed by being attracted to the surfaces of several solid phase materials, referred to as adsorbents. Adsorbents have been made from a variety of materials for a variety of uses, including water treatment, dye separation, indicators, desiccants, and catalysis. An effective adsorption system may be capable of removing all the pollutants, releasing a contaminant-free fluid.
Due to its simplicity, ease of operation, simple design, and ease of scaling up, the adsorption process is considered a better alternative in the treatment of dye from industrial effluents. It also has a high capacity and a favorable rate, and it is insensitive to harmful chemicals [68]. It can also help to resolve the challenges of high energy input (used in reverse osmosis and UV sterilization), which is a problem in many developing countries. Adsorption is preferable to photocatalytic and ultrasound treatment processes because photocatalyst preparation is expensive and ultrasound treatment consumes a lot of energy.

4.1. Adsorbent Properties and Adsorption Mechanism

Adsorbent materials with high selectivity and porosity are required. Activated carbon materials have demonstrated tremendous removal of organics such as dyes [30,69,70]. The diffusion of dye on the surface of adsorbents is either by physical phenomena or chemical means. The dye materials bind to the adsorbent periphery by hydrogen bonding or Van der Waals forces in physisorption mechanism [71]. In chemisorption, the dye ions or molecules form chemical bonds with certain surface functional sites or groups. Section 2 describes various dye functional sites or groups that can involve in multiple adsorptive bonds on different adsorbents. Section 5 of this article explains dye removal via adsorption mechanism using biochar materials.

4.2. Adsorbents

The choice of adsorbent is largely determined by the uptake potential of the adsorbent material for a particular adsorbate. Preferably, the adsorbent material shall meet the following criteria [70,72]:
i.
Reasonably a decent surface area and pore volume;
ii.
Appropriate pore size dissemination and pore network;
iii.
Adsorbent functional groups of the surface charge and appropriate type;
iv.
Surface functional group types and charge on colored dye ions/group;
v.
pH of the solution that is appropriate for uptake.
As many dyestuffs are large molecules, it is critical to compare the dye particle dimensions to the pore (aperture) size distribution of biochar in the first two categories to make certain that the dye particles can penetrate across the pores easily. Furthermore, dye molecules frequently form larger groups or micelles by interacting with one another. As a result, mesoporous adsorbents may be preferable to those with a microporous design and a decent surface area. A high surface area, on the other hand, are usually beneficial. A broad mesoporous network benefits adsorption kinetics and process design by allowing for quicker diffusion.
Clay, silica, kraft lignin, fly ash, sludge, slag, and red mud are some of the most often utilized natural adsorbents. Activated carbon [73], activated alumina [74,75], silica gel, and zeolites are commercially available adsorbents [76]. Agricultural residues/wastes viz. sugarcane bagasse, rice brans, lignocellulosic biomasses, fruit stones, and nut hulls [77,78,79,80], inorganic materials viz. clay, Fuller’s earth, bentonite, lignite coal, peat, chitosan, and io-exchange resin materials [81] are also used as adsorbents. New adsorbent materials such as CNTs, MCM-41 [82,83], and molecular sieves [82] have been discovered in recent literature.
As long as the appropriate adsorbent materials with decent adsorption potential are accessible at a reasonable price, adsorption is an attractive color removal method because it uses large amounts of water and has relatively low dye concentrations. Consequently, biochars have a good chance of being used to remove dyes from effluent. As opposed to activated carbons, biochar is generated through pyrolysis technique at temperatures between 623 and 873 K, and by employing inexpensive biomass materials. Numerous operational factors, including feedstock type and pyrolysis temperature, influence biochar attributes, resulting in products with such a broad range of specific surface area, pore volume, carbon content, volatile content, ash, pH, and cation exchange capacity (CEC). The formation of biochar with a highly evolved high porosity, specific surface area, pH, carbon, and ash content, but low CEC and volatile content, is aided by a high pyrolysis temperature, which is most closely attributable to a high level of breakdown of organic materials. Even at greater pyrolysis temperatures, biochars made from animal and solid wastes have lower carbon content, volatile content, CEC, and surface areas than biochars made from woody biomass and agricultural residues [84]. The reason for this discrepancy is because the content of lignin and cellulose, as well as the moisture content of biomass, varies greatly among plant biomass wastes [85]. During the preparation phase of biochars, modifying agents are also added to create highly charged surface properties. Several instances of altered and unmodified biochar materials are included in Section 5.

5. Adsorption of Dye onto Biochar Materials for Dye Removal

5.1. Biochar as a Dye Removal Adsorbent

Activated carbons are amorphous carbon-based substances with decent porosity and large internal surface area. Its feedstock can be almost any organic substances with a relatively decent carbon composition, covering traditional materials such as hard and softwood, coconut hull, peat and lignite coal, natural and artificial polymers. Surface areas for marketable carbons typically fall between the 500–1500 m2/g range [86] and can even reach 3000 m2/g. Carbon materials are again classified into two sub-categories based on whether they are used to remove pollutant particles from fluids.
The first ones are typically microporous with a pore size of 2 nm diameter and are usually granular, whilst the latter are mesoporous materials with a pore size ranging between 2–50 nm diameter and are usually in powder form [71]. Both types are useful in wastewater treatment, where they aid in decolorization, odor removal, metal recovery, and organics adsorption. The pore volume, internal surface area, and size distribution are all proportional to its adsorption capacity.
Organics have been reported to adsorb in pores just fitting the adsorbate molecule [45]. Humic acids and dyes with size ranging between 1.5 and 3.0 nm that support adsorption phenomenon in mesopores are examples [87]. As a result, the biochar’s pore size dissemination influences its adsorption potential for ions of varying size and shape. The electric strength between the adsorbate and the adsorbent (carbon surface) has proven to improve dye removal efficacy greatly.
The dissociation equilibria and functionality of specific functional sites on biochar material surfaces, such as carboxylic-lactonic groups, phenolic-alcolohic hydroxyl groups, aromatic-heterocyclic carbons, ketone-carbonyl groups, pyridinic-N, pyrollic-N, and quaternary-N nitrogen species, can influence adsorption. These potential biochar surface sites are influenced by several factors, including biomass source, pyrolysis parameters such as heating rate, temperature, residence time, nature of pyrolysis, etc.
Despite that activated and modified carbons are broadly used as adsorbent materials, they are relatively expensive due to the high costs of raw materials, energy, and chemical production. As a result, many researchers have focused on developing novel, high-capacity, low-cost adsorbents obtained from biomass residues. Metal organic frameworks and nano-adsorbent substances have lately been used to create highly efficient adsorbents [88]; however, the cost of treatment renders these materials prohibitively expensive.
As a result, in recent years, several low-cost adsorbents known as biochars have been produced by biomass pyrolysis and used in polluted water treatment applications. The technique is affordable and cost-effective only when the adsorbent is inexpensive and copious [89]. Pyrolysis of biomass leftovers into value-added biochar materials is a cost-effective process that produces high-value-added products: syngas and bio-oil. The pyrolysis process requires energy to run, but the process is driven by the by-products of the side reactions, and biochars possess a larger surface area and pore volume, as well as chemically functional moiety content, making them a much more potent adsorbent material than the biomass feedstock [90,91,92,93].
There are over a thousand papers on color removal in the literature, with over a hundred of these based on dye elimination employing biochar substances, including biochar products derived from vermicompost, cabbage residues, algae, and animal litters [94,95,96,97,98]. There have also been numerous publications on the synthesis and usage of altered biochar materials for dye color elimination. The dye potentials of unmodified/unaltered biochars and others are shown in Table 1 and Table 2.
Table 1 shows the adsorption properties for cationic dye uptake onto unmodified biochars. MG [99,100,101,102,103], MB [104,105,106,107,108,109], rhodamine B (Rh B) [99], basic red 9 (BR 9), and CV [103,110] values are included in the data. Many citations only present the quantity of dye removed (in %) [99,100,104,106,110,111], which is valuable, however, this value varies with adsorbent quantity, dye concentration, and adsorbate volume.
Table 1. Basic dyes (cationic) and their adsorption onto biochars.
Table 1. Basic dyes (cationic) and their adsorption onto biochars.
DyeBiochar FeedstockPyrolysis ConditionsPore Volume (cm3/g)BET Surface Area (m2/g)Adsorption Capacity (mg/g) or Dye Removal (%)Isotherm TypeKinetic ModelParametersMechanismReference
Temperature (K)Heating Rate (K/min)Time (min)pHEquilibrium Time (min)
MBDate palm fronds973-2400.134430205--636-[104]
MGTapioca peel107310180--32%Langmuir, FreundlichPseudo I-order, Pseudo-II order2–100–180-[99]
Rh BTapioca peel107310180--66%Langmuir, FreundlichPseudo I-order, Pseudo-II order2–100–180-[99]
MBChlorella sp. microalgaeMW heating (2450 MHz, 800 W)---3110Freundlich, TemkinPseudo I-order, Pseudo-II order, Elovich2–107200Boyd, Intraparticle diffusion[105]
MGRice husk673–873-60--65Langmuir, FreundlichPseudo I-order, Pseudo-II order, Elovich2, 4, 6, 81440-[100]
MGCrab shell1073-1200.0868212,500LangmuirPseudo-II order72Electrostatic attraction, Hydrogen bonding, π-π interactions[101]
MBAreca leaf473560-21120Langmuir, FreundlichPseudo I-order, Pseudo-II order7720Electrostatic attraction[106]
MBWodyetia Bifurcate9731030--150SipsPseudo I-order, Pseudo-II order-30-[107]
MGWaste wheat straw/wheat bran10731590--1740LangmuirPseudo-II order2, 4, 6, 8, 10-Electrostatic interaction, Chemisorption[102]
CVWaste wheat straw/wheat bran10731590--175LangmuirPseudo-II order2, 4, 6, 8, 10-Electrostatic interaction,
Chemisorption
[102]
MBSwitchgrass873-600.02925540LangmuirPseudo-II order6-Intraparticle diffusion[108]
MBSwitchgrass-1173-600.058640200LangmuirPseudo-II order6-Intraparticle diffusion[108]
CVMango leaves1073-60-170180 -848-[110]
MGUlothrix zonata algae10731590-1305300FreundlichPseudo-II order2, 4, 6, 10840Chemisorption[103]
CVUlothrix zonata algae10731590-1301220FreundlichPseudo-II order2, 4, 6, 10840Chemisorption[103]
BR 9Bovine bones107310600.2719050Langmuir, FreundlichPseudo-II order7180-[111]
BR 9Bovine bones1073101800.1939550 Pseudo I-order7180-[111]
MBSugarcane bagasse7731090-26070Langmuir, FreundlichPseudo I-order, Pseudo-II order7.4180Intraparticle diffusion[109]
MB adsorption capacities (Table 1) are 110, 150, 38, and 195 mg/g for biochar materials derived from microalgae [105], Wodyetia [107], and switchgrass at 873 and 1173 K [45,108], respectively. Pyrolysis temperature significantly influences the adsorption potential of switchgrass biochar. MB adsorption capacity values in the literature for sugarcane bagasse [109], phosphoric acid-treated olive seed carbon [112], and bamboo cane active carbon [113] are 110, 135, and 455 mg/g, respectively. Biochar adsorption capacities for MG dye are remarkably decent, with values of 12,500 mg/g on crab shell [101], 1740 mg/g on wheat/bran straw-fed larvae biochar [102], and 5300 mg/g on biochar synthesized from Ulothrix algae [103].
The adsorbent potential values reported in the literature for activated carbon made from grape processing residue [114], shrimp shell [115], and plastic waste [116] are 665, 320, and 1430 mg/g, respectively. The amount of Rh B adsorbed on the surface of biochar sourced from tapioca shell is 33 mg/g [99] compared to reported values of 77 using Acacia mangium wood-derived carbon [117] and 30–40 mg/g on activated carbons from carnauba, macauba, and pine nut wastes activated using calcium chloride and phosphoric acid [118].
On biochar from mango leaves [102] and Ulothrix zonata algae [103], two capacities for the removal of CV are listed: 175 mg/g and a quite high value of 1220 mg/g. The capacity values reported in the literature range from 600 mg/g for a bentonite-alginate composite [119] to 75 mg/g for chitosan hydrogel beads [120]. Table 1 shows the final values for the adsorption of BR9 onto biochar materials from animal residue [111] after 1 and 3 h of heat treatment. At one (90 m2/g) and three hours (95 m2/g), the BR9 capacities were 50 and 52 mg/g, respectively. The general trend in surface areas and pyrolysis times was followed by these capacities. Literature values are slightly lower but of relative magnitude, for example, 29 and 15 mg/g for sepiolite [121] and fish bone [122], respectively.
Table 2 shows the anionic dye adsorption properties on unmodified biochars. Acid orange 7 (AO 7) [104], CR [96,97,101,102,103,105,108,123,124,125,126], reactive red RR 120 [127], Remazol violet 5R (RV5R), Remazol orange 3R (RO 3R), Remazol blue R (RBR) [128], orange G (OG) [108], and methyl orange (MO) [129] are some of them.
Table 2 shows only the dye removal composition (%) for the adsorption of AO7 using biochar derived from groundnut shell. Only a few instances of AO7 adsorption capacity potential are documented in articles, and they range from 50 to 180 mg/g on fly ash [130], oxihumolite [131], and chemically reactivated sawdust [132].
Table 2. Acid dyes (anionic) and their adsorption onto biochars.
Table 2. Acid dyes (anionic) and their adsorption onto biochars.
DyeBiochar FeedstockPyrolysis ConditionsPore Volume (cm3/g)BET Surface Area (m2/g)Adsorption Capacity (mg/g) or Dye Removal (%)Isotherm TypeKinetic ModelAdsorbent ParametersMechanismReference
Temperature (K)Heating Rate (K/min)Time (min)pHEquilibrium Time (min)
CRChlorella sp. microalgalMW heating (2450 MHz, 800 W)---3160Langmuir, Freundlich, TemkinPseudo I-order, Pseudo II-order, Elovich2–10240Boyd, Intraparticle diffusion[105]
CRRice husk7735180--66–97%Langmuir, Freundlich-2, 4, 6, 7, 9, 115760-[97]
RR 120Eucheuma spinosum573–87310120--330Langmuir, Freundlich, TemkinPseudo I-order, Pseudo II-order, Elovich3–920Electrostatic interaction, Ion exchange, Metal complexation, Hydrogen bonding[127]
CRPhoenix dactylifera leaves673---125Langmuir, FreundlichPseudo I-order, Pseudo II-order5.8120-[125]
CRCotton stalks673890--250Langmuir, Freundlich, Temkin, Dubinin-RadushkevichPseudo I-order, Pseudo II-order2–10180Electrostatic attraction[126]
CROrange peel10731515- 20 ----[123]
Remazol BV 5RGreen marine algae (Caulerpa scalpelliformis)573–7735120--70%Langmuir, Freundlich, Sips, TPseudo I-order, Pseudo II-order2–5--[128]
Remazol BO 3RGreen marine algae (Caulerpa scalpelliformis)573–7735120--77%Langmuir, Freundlich, Sips, TemkinPseudo I-order, Pseudo II-order2–5--[128]
Remazol BO 3RGreen marine algae (Caulerpa scalpelliformis)573–7735120--75%Langmuir, Freundlich, Sips, TemkinPseudo I-order, Pseudo II-order2–5--[128]
Remazol BO 3RCrab shell1073-1200.0868220,315LangmuirPseudo I-order, Pseudo II-order42Electrostatic attraction, Hydrogen bonding, π-π interactions[101]
CRActivated Carbon72320120--230Freundlich-2–10120-[124]
CRSpirulina platensis algae72320120-- Freundlich-2–10120-[96]
CRWaste wheat straw/wheat bran10731590--90LangmuirPseudo II-order2, 4, 6, 8, 10-Chemisorption, Electrostatic interaction[102]
OGSwitchgrass873-600.0292558LangmuirPseudo II-order6-Outer boundary[108]
CRSwitchgrass873-600.0292558LangmuirPseudo II-order6-Outer boundary[108]
CRSwitchgrass1173 600.05864020LangmuirPseudo II-order6-Outer boundary[108]
CRUlothrix zonata algae10731590-130345FreundlichPseudo II-order2, 4, 6, 10840Chemisorption[103]
MOCorn cob87315120-47090FreundlichPseudo II-order5.6-Physiochemical[129]
As CR is one of the most researched anionic dyestuffs, the citations in Table 2 are merely illustrative. Adsorption capacity potential of biochars derived from chlorella microalgae species [105], phoenix dactylifera [125], cotton stalk [126], orange skin [123], carapace (crab shell) [101], activated carbon [124], spirulina algae species [96], wheat bran larvae [102], switchgrass (charred at 873 K and 1173 K) [108], and Ulothrix algae species [103] are 160, 25, 250, 90, 20,315, 230, 85, 8, 23, and 345 mg/g. The study on switchgrass indicated that at elevated pyrolysis temperatures, a high-quality biochar is generated.
Most investigations have found that the CR dye adsorption capacity potential is below 100 mg/g. The maximum value was observed from pyrolyzed crab shell with 80 m2/g surface area. This result was achieved at a pH of 4 and volume to mass ratio of 2; nonetheless, at a CR concentration above 20 g/L. The activated carbon obtained from date stone exhibited a low adsorption capacity potential of 35 mg/g. The huge dye molecular size (695 g/mol) and the total pore volume of 0.086 are responsible for the poor capacities. CR dye, unlike other anionic acid dyes, is a direct dye with no anionic bonding characteristics [133,134]. The significance of examining adsorbent pore size distribution is shown by this phenomenon [135]. Microwave treatment was used following phosphoric acid activation to synthesize mesoporous activated carbon. The carbon showed a higher surface area and a 350 mg/g adsorption capability.
At temperatures between 513 to 553 K and reaction durations varying from 0.5 to 6.0 h, many bamboo biochars were generated [136]. Their CR absorption capabilities ranged from 30–100 mg/g, with the maximum values seen in biochars generated at high temperatures of 513 K and 553 K, as well as the longer treatment times of 5 and 6 h. After only 4 h at 523 K, an activated carbon obtained from apricot seeds [134] had a low CR capacity of 33 mg/g. This was ascribed to the small pyrolysis temperature and specific surface area. The latter two investigations’ adsorption capacity potential was similar to that of date seed carbon. More research into the production of biochars through microwave and plasma pyrolysis techniques should be conducted.
Anionic reactive red (RR) 120 was adsorbed at a high (330 mg/g) capacity on biochar synthesized from Eucheuma spinosum [127]. Fe3O4-activated magnetic nanoparticles [137] and activated carbon [138] have high capacity values in the literature, exhibiting adsorption capacity potentials of 165 and 255 mg/g, respectively. Biochar made from green sea algae [128] has been reported to remove brilliant violet 5R, Remazol, and brilliant orange 3R dyes with removal percentages of more than 70%. On coffee shell activated carbon [139] and calcined eggshell [140], the reported figures for Remazol dyes are relatively low, at 65 and 15 mg/g for brilliant orange 3R and brilliant violet 5R, respectively.
Table 2 demonstrates that biochar derived from switchgrass [108] generated at 873 K displayed a poor orange G adsorption capacity of 8 mg/g. The poor surface area (255 m2/g) of the char could have contributed to this low capacity. Other published values include 9 mg/g for activated carbon derived from Thespesia populnea [141] and 19 mg/g for nanoporous activated carbon [142]. All of these values indicate that orange G dye is one to be treated. When it came to adsorbing MO, corn cob char [129] had an adsorption capacity potential of 85 mg/g, while amidoxime char [143] had a potential of 140 mg/g.

5.2. Dye Removal Using Adsorption onto Modified Biochars

Several investigations are now being conducted to improve the adsorption efficacy of biochars. Many papers on various biochar modification techniques are available. Treatment or activation of biochar with bases and acids, chemical impregnation, size alteration, and encapsulation are some of the modification techniques. The cationic dye adsorption characteristics and performance properties are shown in Table 3, whereas the anionic dye adsorption characteristics and performance properties are shown in Table 4.
On Fe/Mn impregnated fronds pyrolyzed at 973 K [147] and 1073 K thermally treated fronds [144], the biochar from date palm frond [104] exhibited an MB adsorption capacity of 205 mg/g, whereas the biochars obtained from modified date frond demonstrated capacities of 300 and 210 mg/g. Biochar made from tapioca skin has a 30% MG removal capacity and a 65% Rh B elimination potential [99]. The removal capacity of biochar made from sulfur-doped tapioca peel was 75 and 90%, respectively [99]. 30 and 33 mg/g were reported as the greatest adsorption potential.
The sulfur-doped tapioca skin was coated with chitosan at 873 K, which raised the dye absorption capabilities to 50 mg/g for MG and 40 mg/g for Rh B, respectively [145]. Unaltered areca plant biochar showed an MG capacity of 190 mg/g [106]. After activating with K2FeO4, the adsorption potential of biochar prepared from wakame increased to 250 mg/g. The biochar sourced from chlorella microalgae displayed an MG adsorption capacity of 110 mg/g and is one among many seaweed/algae species biochars with potentials lying between 25 and 130 mg/g. The adsorption capability of MG on biochar obtained from chlorella was close to 80 mg/g [151].
Only 10 mg/g capacity of Rh B was accepted by seaweed biochar. Biochar made from calcined wakame seaweed, on the other hand, has extraordinarily high adsorption capabilities for MG, Rh B, and MB, with 4065, 840, and 530 mg/g, respectively [148]. Biochar generated from rice husk had an MG adsorption capacity of 65 mg/g [100]; after alteration with Cu + Al [149], the potential elevated to 470 mg/g. Biswas et al. [109] found that biochar made from sugarcane bagasse had a potential of 70 mg/g, while biochars made from steam-activated bagasse and ZnCl2-modified bagasse had capacities of 5220 and 90 mg/g, respectively [109].
Table 4 shows the anionic dye adsorption properties on a variety of modified biochars. Biochar was prepared from rubber seeds and treated with NaOH at 1073 K. The adsorption potential of CR rose from 225 mg/g to 460 mg/g after activation [152]. The shorea robusta leaf extract offered a removal potential of only 2 mg/g for CR with an exceedingly poor surface area of 1.5 m2/g after heating at 573 K [153]. The addition of Ag nanoparticles to this biochar raised its capacity and surface area to 23 mg/g and 21 m2/g, respectively [153].
Table 4. Acid dyes (anionic) and their adsorption onto modified (altered) biochars.
Table 4. Acid dyes (anionic) and their adsorption onto modified (altered) biochars.
DyeBiochar Modified FeedstockPyrolysis ConditionsPore Volume (cm3/g)BET Surface Area (m2/g)Adsorption Capacity (mg/g) or Dye Removal (%)Isotherm TypeKinetic ModelAdsorbent ParametersMechanismReference
Temperature (K)Heating Rate (K/min)Time (min)pHEquilibrium Time (min)
CRRubber seeds + NaOH1073-360--460Langmuir, Freundlich, Dubinin-Radushkevich-6–7120-[152]
CRShorea robusta leaf extract + Agnps573-180-2120Langmuir, Freundlich, Temkin, Dubinin-RadushkevichPseudo I-order, Pseudo II-order, Intraparticle diffusion, Elovich2–1090Electrostatic attraction, Hydrogen bonding[153]
CRShorea robusta leaf extract573-180-12 -2–1060Electrostatic attraction[153]
MOFood waste + ultrasound + H2O25735420--69% -760-[43]
CRPhoenix dactylifera leaves + Mn673----120Langmuir, FreundlichPseudo I-order, Pseudo II-order2.5, 4.5, 5.8, 8.2, 11.2120Electrostatic interaction[125]
Reactive yellow (RY)Marine Chlorella + ultrasound723–9231060-35050Langmuir, Freundlich, TemkinPseudo I-order, Pseudo II-order2.0–10.02Electrostatic interaction, Electrostatic repulsion[154]
Reactive blue (RB 19)/Acid orange (AO) II/Direct redSludge-rice husk composite77371200.0583039, 42, 60Langmuir, Freundlich, Temkin, Dubinin-RadushkevichPseudo I-order, Pseudo II-order, Intraparticle diffusion, Elovich71440-[155]
CRArjuna (Terminalia Arjuna) seeds----17092 ± 5% ----[156]
CRCotton stalks673890--560Langmuir, Freundlich, Temkin, Dubinin-RadushkevichPseudo I-order, Pseudo II-order, Intraparticle diffusion2–10180Electrostatic attraction[126]
CROrange peel + CO2 + steam973-10-305140FreundlichPseudo II-order2–31440Electrostatic interaction[123]
CRLitchi peel + hydro-thermal1123 600.5881005400FreundlichElovich4720Hydrogen bonding, π-π interactions, Pore-filling, Electrostatic interaction[150]
CRSpirulina/alginate/paper72320120--40Langmuir, Freundlich, Temkin, Dubinin-RadushkevichPseudo I-order, Pseudo II-order,
Intraparticle diffusion
6–80–120Electrostatic attraction, Hydrogen bonding, π-π,[157]
Acid chrome blue/MOPine nutshell97310120--30, 10Langmuir, FreundlichPseudo I-order, Pseudo II-order31200Electrostatic interaction, π-π interactions[158]
A biochar made from magnetic food waste with a removal potential of 23% was produced for application in a Fenton-style wastewater treatment system [43]. The biochar adsorption potential increased by 33% after ultrasonic treatment. Finally, ultrasonics + H2O2 + biochar achieved a 97% removal in 3 h. Table 2 shows that dactylifera leaf biochar (without any modification) exhibited a CR adsorption potential of 25 mg/g before Mn activation, however, it increased to 120 mg/g after Mn activation [125]. Marine chlorella vulgaris algae were transformed to biochar at temperatures of 723, 823, and 923 K [154], yielding surface areas of 265, 350, and 150 m2/g, respectively. This adsorbent’s adsorption potential for the anionic RY-45 dye was 48 mg/g (experimental) and 58 mg/g at 823 K (Langmuir).
Biochars made from sludge have been extensively studied [159]. The Langmuir capacity of the anionic Remazol BB R dye is 125 mg/g [160]. A biochar made out of rice husk-sludge composite prepared at 773 K removed AO II, and the two additional anionic dyes, DR 4BS and RB 19, from solution exhibited adsorption potentials of 42, 60 and 39 mg/g successfully [155].
The separation of 90% of CR from the dye solution was achieved by utilizing a hybrid technique devised by using Arjuna seeds + microorganisms and the ozonation process [156]. Cotton stalk waste is copious around the world. The biochar from cotton waste has been employed to eliminate CR with a 250 mg/g capacity. The addition of ZnO nanoparticles to this biochar increased its adsorption potential to 555 mg/g [126]. The CR dye adsorption capacity of biochar made from orange peel was about 15 mg/g, nonetheless, two tailored biochars made by using steam and CO2 exhibited capacities of 135 and 90 mg/g, respectively. A one-hour hydrothermal carbonisation of litchi skin at 1073 K produced an altered biochar with a surface area of 1005 m2/g and a good CR adsorption potential of 400 mg/g [156].
Spirulina, paper, and seaweed alginate have been combined to create a modified biochar [157]. The goal of this research was to see if algae biorefinery waste and wastepaper could be used to make cost-effective and ecologically friendly xerogels for CR elimination. At the ideal pH value of 6–8, the produced biosorbents possessed a light and porous network structure and a quick dye uptake. The adsorption potential of CR was 40 mg/g.
Biochar was made from raw pine nut husks using fast pyrolysis at 973 K for 3 h [158]. Acid chrome blue K dye sorbates and anionic MO were used to examine the dye adsorption capabilities of the biochar. Before being exposed to a final FeCl3 alteration, the unaltered biochar was treated with modified magnetic biochar and cetyl trimethyl ammonium bromide. Acid chrome blue K and MO have adsorption potentials of 16 and 1 mg/g, respectively, on unaltered biochar. Following the modification, the capacities for acid chrome blue K and MO were 25 and 10 mg/g, respectively.

6. Conclusions

This review study discussed the use of biochars and modified biochars for dye removal from effluents. There are several hundred studies on this subject in the literature, and some of them have been cited and detailed in this review. Chemical species, chemical groups, and dyestuff qualities were discussed in this article. A quick overview of treatment technologies followed by a full discussion of the advantages of employing adsorption technology was also covered in this review. The review also highlighted the main features and applications of regular and modified (altered) biochars for color removal.
Certain performance trends can generally be observed and correlated with biochar properties, for example, when using the same raw material source as the fuel source:
The dye adsorption potential/capacity is highly linked to surface area, therefore, the greater the pyrolysis temperature, the greater the dye adsorption capacity;
The biochar yield decreases as the temperature rises. However, this is only true up to roughly 1073 K;
As the micropore and small mesopore walls burn away at temperatures beyond 1073–1123 K, pore volume increases, resulting in fewer but larger pores;
It is important to consider the reaction conditions of temperature, time, and heating rate based upon the type of dyestuff as well as the pyrolysis temperature. The thickness of the pores is also affected by both the type of raw materials and the pyrolysis temperature—dye molecules vary enormously in size, and even small dye molecules are relatively large in comparison to many chemical molecules;
Dye is strongly attracted to oppositely charged sites, so the nature of surface sites on the biochar, depending on raw material and temperature, is extremely important;
Slow pyrolysis yields the best biochar and has the best property control because it produces more biochar, guarantees better pore development, and has a narrower spectrum of pore size distribution.
There are so many different types of modified biochars that it is impossible to list them all; instead, a few brief examples are provided:
Acid or alkali chemical treatment produces biochars with negative and positive surface sites or groups; at temperatures above 823 K, these biochars are known as activated carbons;
Before pyrolysis treatment, sulfur doping of the feedstocks generates biochars carbons with a decent affinity for hazardous heavy metal ions;
Iron oxide-doped modified chars have demonstrated to be particularly attractive adsorbents for both anions and metal cations, as well as chromate, using ion exchange;
Coating the biochars has been exceedingly successful; for instance, coating with chitosan provided adsorption capabilities of 5-fold and 20-fold w/w, placing these altered biochars in the super-adsorbent category.
The future of biochar technology appears bright overall. Several biochars are made from waste biomass resources and are therefore carbon neutral, making them a cost-effective and environmentally friendly product with many potential applications. The cited literature contains numerous gaps that will be essential to address for designing treatment plants for dyehouse effluent biochar. However, recent articles have addressed the “tomorrow’s path” which should be followed in future adsorption studies.

Author Contributions

Original Draft, Writing—Review and Editing, P.P.; Original Draft, S.S. and J.S.; Original Draft, Writing—Review and Editing, M.A.; Funding acquisition, Conceptualization, Resources, Project administration, Supervision, Writing—Review and Editing, G.M. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thank Qatar National Research Fund for their support of this research through an award NPRP-11S-0117-180328, the Supreme Committee for Delivery and Legacy (SCDL), and to Hamad Bin Khalifa University, Qatar Foundation, for an award to H.F. Any opinions, findings, and conclusions, or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of HBKU or QF or SCDL.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors wish to thank Hamad Bin Khalifa University (HBKU), Qatar Foundation (QF), and Qatar University for their patronage.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Akakuru, O.U.; Iqbal, Z.M.; Wu, A. TiO2 Nanoparticles Properties and Applications. In TiO2 Nanoparticles: Applications in Nanobiotechnology and Nanomedicine; John Wiley & Sons, Ltd.: Weinhein, Germany, 2020; pp. 1–66. ISBN 9783527825448. [Google Scholar]
  2. Oyewo, O.A.; Elemike, E.E.; Onwudiwe, D.C.; Onyango, M.S. Metal oxide-cellulose nanocomposites for the removal of toxic metals and dyes from wastewater. Int. J. Biol. Macromol. 2020, 164, 2477–2496. [Google Scholar] [CrossRef] [PubMed]
  3. Zhu, S.; Xia, M.; Chu, Y.; Khan, M.A.; Lei, W.; Wang, F.; Muhmood, T.; Wang, A. Adsorption and Desorption of Pb(II) on l-Lysine Modified Montmorillonite and the simulation of Interlayer Structure. Appl. Clay Sci. 2019, 169, 40–47. [Google Scholar] [CrossRef]
  4. Khan, S.; Malik, A. Toxicity evaluation of textile effluents and role of native soil bacterium in biodegradation of a textile dye. Environ. Sci. Pollut. Res. 2018, 25, 4446–4458. [Google Scholar] [CrossRef] [PubMed]
  5. Kishor, R.; Purchase, D.; Saratale, G.D.; Saratale, R.G.; Ferreira, L.F.R.; Bilal, M.; Chandra, R.; Bharagava, R.N. Ecotoxicological and health concerns of persistent coloring pollutants of textile industry wastewater and treatment approaches for environmental safety. J. Environ. Chem. Eng. 2021, 9, 105012. [Google Scholar] [CrossRef]
  6. McKay, G.; Parthasarathy, P.; Sajjad, S.; Saleem, J.; Alherbawi, M. Dye removal using biochars. In Sustainable Biochar for Water and Wastewater Treatment; Mohan, D., Pittman, C.U., E.Mlsna, T., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 429–471. [Google Scholar]
  7. Forgacs, E.; Cserháti, T.; Oros, G. Removal of synthetic dyes from wastewaters: A review. Environ. Int. 2004, 30, 953–971. [Google Scholar] [CrossRef]
  8. Velusamy, S.; Roy, A.; Sundaram, S.; Kumar Mallick, T. A Review on Heavy Metal Ions and Containing Dyes Removal Through Graphene Oxide-Based Adsorption Strategies for Textile Wastewater Treatment. Chem. Rec. 2021, 21, 1570–1610. [Google Scholar] [CrossRef]
  9. Global Textile Dyes Industry Report 2015—Forecasts to 2020. Available online: https://www.prnewswire.com/news-releases/global-textile-dyes-industry-report-2015---forecasts-to-2020-498532981.html (accessed on 15 May 2021).
  10. Afroze, S.; Sen, T.K. A Review on Heavy Metal Ions and Dye Adsorption from Water by Agricultural Solid Waste Adsorbents. Water Air Soil Pollut. 2018, 229, 1–50. [Google Scholar] [CrossRef]
  11. Wang, K.; Wei, T.; Li, Y.; He, L.; Lv, Y.; Chen, L.; Ahmad, A.; Xu, Y.; Shi, Y. Flocculation-to-adsorption transition of novel salt-responsive polyelectrolyte for recycling of highly polluted saline textile effluents. Chem. Eng. J. 2021, 413, 127410. [Google Scholar] [CrossRef]
  12. Al-Degs, Y.; Khraisheh MA, M.; Allen, S.J.; Ahmad, M.N. Ahmad Effect of carbon surface chemistry on the removal of reactive dyes from textile effluent. Water Res. 2000, 34, 927–935. [Google Scholar] [CrossRef]
  13. Al-Degs, Y.S.; El-Barghouthi, M.I.; El-Sheikh, A.H.; Walker, G.M. Effect of solution pH, ionic strength, and temperature on adsorption behavior of reactive dyes on activated carbon. Dye. Pigment. 2008, 77, 16–23. [Google Scholar] [CrossRef]
  14. Chiou, M.S.; Kuo, W.S.; Li, H.Y. Removal of Reactive Dye from Wastewater by Adsorption Using ECH Cross-Linked Chitosan Beads as Medium. J. Environ. Sci. Heal. Part A 2007, 38, 2621–2631. [Google Scholar] [CrossRef] [PubMed]
  15. Chakraborty, J.N. Metal complex dyes. In Handbook of Textile and Industrial Dyeing Principles, Processes and Types of Dyes; Clark, M., Ed.; Woodhead Publishing Series in Textiles: Cambridge, UK, 2011; pp. 446–463. [Google Scholar]
  16. Yusuf, A. Vat Dyes-Properties-Dyeing Mechanism-A Comprehensive Look (2020). Available online: https://textiletuts.com/vat-dyes/ (accessed on 2 March 2022).
  17. ULLMANN’S Encyclopedia of Industrial Chemistry; Arpe, H.-J. (Ed.) Wiley: Weinheim, Germany, 2000. [Google Scholar]
  18. Asif Tahir, M.; 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]
  19. Gupta, V.K.; Mohan, D.; Sharma, S.; Sharma, M. Removal of Basic Dyes (Rhodamine B and Methylene Blue) from Aqueous Solutions Using Bagasse Fly Ash. Sep. Sci. Technol. 2007, 35, 2097–2113. [Google Scholar] [CrossRef]
  20. Allen, S.J.; Mckay, G.; Porter, J.F. Adsorption isotherm models for basic dye adsorption by peat in single and binary component systems. J. Colloid Interface Sci. 2004, 280, 322–333. [Google Scholar] [CrossRef]
  21. Fan, Q.; Hoskote, S.; Hou, Y. Reduction of colorants in nylon flock dyeing effluent. J. Hazard. Mater. 2004, 112, 123–131. [Google Scholar] [CrossRef]
  22. Ho, Y.S.; Chiang, C.C. Sorption Studies of Acid Dye by Mixed Sorbents. Adsorption 2001, 7, 139–147. [Google Scholar] [CrossRef]
  23. Valix, M.; Cheung, W.H.; McKay, G. Preparation of activated carbon using low temperature carbonisation and physical activation of high ash raw bagasse for acid dye adsorption. Chemosphere 2004, 56, 493–501. [Google Scholar] [CrossRef]
  24. Yu, M.C.; Skipper, P.L.; Tannenbaum, S.R.; Chan, K.K.; Ross, R.K. Arylamine exposures and bladder cancer risk. Mutat. Res. Mol. Mech. Mutagen. 2002, 506–507, 21–28. [Google Scholar] [CrossRef]
  25. Rovina, K.; Prabakaran, P.P.; Siddiquee, S.; Shaarani, S.M. Methods for the analysis of Sunset Yellow FCF (E110) in food and beverage products- a review. TrAC Trends Anal. Chem. 2016, 85, 47–56. [Google Scholar] [CrossRef]
  26. Nikfar, S.; Jaberidoost, M. Dyes and Colorants. In Encyclopedia of Toxicology: Third Edition; Academic Press: Cambridge, MA, USA, 2014; pp. 252–261. ISBN 9780123864543. [Google Scholar]
  27. Goswami, P.; Sulfur, M.B. Dyes. In Kirk-Othmer Encyclopedia of Chemical Technology; John Wiley & Sons: Hoboken, NJ, USA, 2000; pp. 1–31. [Google Scholar]
  28. Thakur, S.; Qanungo, K. Removal of aniline blue from aqueous solution using adsorption: A mini review. Mater. Today Proc. 2021, 37, 2290–2293. [Google Scholar] [CrossRef]
  29. Baker, J.R. Principles of Biological Microtechnique; Methuen: London, UK, 1958. [Google Scholar]
  30. Gupta, V.K. Suhas Application of low-cost adsorbents for dye removal—A review. J. Environ. Manage. 2009, 90, 2313–2342. [Google Scholar] [CrossRef] [PubMed]
  31. Ip, A.W.M.; Barford, J.P.; McKay, G. Biodegradation of Reactive Black 5 and bioregeneration in upflow fixed bed bioreactors packed with different adsorbents. J. Chem. Technol. Biotechnol. 2010, 85, 658–667. [Google Scholar] [CrossRef]
  32. Mcyotto, F.; Wei, Q.; Macharia, D.K.; Huang, M.; Shen, C.; Chow, C.W.K. Effect of dye structure on color removal efficiency by coagulation. Chem. Eng. J. 2021, 405, 126674. [Google Scholar] [CrossRef]
  33. Bayramoglu, M.; Eyvaz, M.; Kobya, M. Treatment of the textile wastewater by electrocoagulation: Economical evaluation. Chem. Eng. J. 2007, 128, 155–161. [Google Scholar] [CrossRef]
  34. Hasani, G.; Maleki, A.; Daraei, H.; Ghanbari, R.; Safari, M.; McKay, G.; Yetilmezsoy, K.; Ilhan, F.; Marzban, N. A comparative optimization and performance analysis of four different electrocoagulation-flotation processes for humic acid removal from aqueous solutions. Process Saf. Environ. Prot. 2019, 121, 103–117. [Google Scholar] [CrossRef]
  35. Al-Ahmed, Z.A.; Al-Radadi, N.S.; Ahmed, M.K.; Shoueir, K.; El-Kemary, M. Dye removal, antibacterial properties, and morphological behavior of hydroxyapatite doped with Pd ions. Arab. J. Chem. 2020, 13, 8626–8637. [Google Scholar] [CrossRef]
  36. Robinson, T.; McMullan, G.; Marchant, R.; Nigam, P. Remediation of dyes in textile effluent: A critical review on current treatment technologies with a proposed alternative. Bioresour. Technol. 2001, 77, 247–255. [Google Scholar] [CrossRef]
  37. Samsami, S.; Mohamadi, M.; Sarrafzadeh, M.H.; Rene, E.R.; Firoozbahr, M. Recent advances in the treatment of dye-containing wastewater from textile industries: Overview and perspectives. Process Saf. Environ. Prot. 2020, 143, 138–163. [Google Scholar] [CrossRef]
  38. Yang, L.; Wang, Z.; Zhang, J. Zeolite imidazolate framework hybrid nanofiltration (NF) membranes with enhanced permselectivity for dye removal. J. Memb. Sci. 2017, 532, 76–86. [Google Scholar] [CrossRef]
  39. Tang, L.; Xiao, F.; Wei, Q.; Liu, Y.; Zou, Y.; Liu, J.; Sand, W.; Chow, C. Removal of active dyes by ultrafiltration membrane pre-deposited with a PSFM coagulant: Performance and mechanism. Chemosphere 2019, 223, 204–210. [Google Scholar] [CrossRef]
  40. Ellouze, E.; Souissi, S.; Jrad, A.; Amar, R.B.; Salah, A. Ben Performances of nanofi ltration and reverse osmosis in textile industry waste water treatment. Desalin. Water Treat. 2012, 22, 182–186. [Google Scholar] [CrossRef]
  41. Ammar, A.; Dofan, I.; Jegatheesan, V.; Muthukumaran, S.; Shu, L. Comparison between nanofiltration and forward osmosis in the treatment of dye solutions. Desalin. Water Treat. 2014, 54, 853–861. [Google Scholar] [CrossRef]
  42. Balcik-Canbolat, C.; Sengezer, C.; Sakar, H.; Karagunduz, A.; Keskinler, B. Recovery of real dye bath wastewater using integrated membrane process: Considering water recovery, membrane fouling and reuse potential of membranes. Environ. Technol. 2016, 38, 2668–2676. [Google Scholar] [CrossRef]
  43. Chu, J.H.; Kang, J.K.; Park, S.J.; Lee, C.G. Application of magnetic biochar derived from food waste in heterogeneous sono-Fenton-like process for removal of organic dyes from aqueous solution. J. Water Process Eng. 2020, 37, 101455. [Google Scholar] [CrossRef]
  44. Anantha, M.S.; Olivera, S.; Hu, C.; Jayanna, B.K.; Reddy, N.; Venkatesh, K.; Muralidhara, H.B.; Naidu, R. Comparison of the photocatalytic, adsorption and electrochemical methods for the removal of cationic dyes from aqueous solutions. Environ. Technol. Innov. 2020, 17, 100612. [Google Scholar] [CrossRef]
  45. Cheremisinoff, P.N. Biomanagement of Wastewater and Wastes; Prentice Hall: New York, NY, USA, 1993. [Google Scholar]
  46. Santos, A.; Yustos, P.; Rodríguez, S.; Garcia-Ochoa, F.; De Gracia, M. Decolorization of Textile Dyes by Wet Oxidation Using Activated Carbon as Catalyst. Ind. Eng. Chem. Res. 2007, 46, 2423–2427. [Google Scholar] [CrossRef]
  47. Chang, D.J.; Chen, I.P.; Chen, M.T.; Lin, S.S. Wet air oxidation of a reactive dye solution using CoAlPO4-5 and CeO2 catalysts. Chemosphere 2003, 52, 943–949. [Google Scholar] [CrossRef]
  48. Boudissa, F.; Mirilà, D.; Arus, V.A.; Terkmani, T.; Semaan, S.; Proulx, M.; Nistor, I.D.; Roy, R.; Azzouz, A. Acid-treated clay catalysts for organic dye ozonation—Thorough mineralization through optimum catalyst basicity and hydrophilic character. J. Hazard. Mater. 2019, 364, 356–366. [Google Scholar] [CrossRef]
  49. Faria, P.C.C.; Órfão, J.J.M.; Pereira, M.F.R. Activated carbon and ceria catalysts applied to the catalytic ozonation of dyes and textile effluents. Appl. Catal. B Environ. 2009, 88, 341–350. [Google Scholar] [CrossRef]
  50. Chaudhuri, S.K.; Sur, B. Oxidative Decolorization of Reactive Dye Solution Using Fly Ash as Catalyst. J. Environ. Eng. 2000, 126, 583–594. [Google Scholar] [CrossRef]
  51. Christie, R.M. Handbook of Textile and Industrial Dyeing- 17-Fluorescent Dyes; M.Clark, Ed.; Woodhead Publishing: Cambridge, UK, 2011; ISBN 9781845696955. [Google Scholar]
  52. Raghavacharya, C. Colour removal from industrial effluents. Chem. Eng. World 1997, 32, 53–54. [Google Scholar]
  53. Sarasa, J.; Roche, M.P.; Ormad, M.P.; Gimeno, E.; Puig, A.; Ovelleiro, J.L. Treatment of a wastewater resulting from dyes manufacturing with ozone and chemical coagulation. Water Res. 1998, 32, 2721–2727. [Google Scholar] [CrossRef]
  54. Farzaneh, H.; Loganathan, K.; Saththasivam, J.; McKay, G. Ozone and ozone/hydrogen peroxide treatment to remove gemfibrozil and ibuprofen from treated sewage effluent: Factors influencing bromate formation. Emerg. Contam. 2020, 6, 225–234. [Google Scholar] [CrossRef]
  55. Oliveira, G.A.R.; Ferraz, E.R.A.; Chequer, F.M.D.; Grando, M.D.; Angeli, J.P.F.; Tsuboy, M.S.; Marcarini, J.C.; Mantovani, M.S.; Osugi, M.E.; Lizier, T.M.; et al. Chlorination treatment of aqueous samples reduces, but does not eliminate, the mutagenic effect of the azo dyes Disperse Red 1, Disperse Red 13 and Disperse Orange 1. Mutat. Res. Toxicol. Environ. Mutagen. 2010, 703, 200–208. [Google Scholar] [CrossRef]
  56. Martínez, F.; Calleja, G.; Melero, J.A.; Molina, R. Heterogeneous photo-Fenton degradation of phenolic aqueous solutions over iron-containing SBA-15 catalyst. Appl. Catal. B Environ. 2005, 60, 181–190. [Google Scholar] [CrossRef]
  57. Silvestri, S.; Stefanello, N.; Sulkovski, A.A.; Foletto, E.L. Preparation of TiO2 supported on MDF biochar for simultaneous removal of methylene blue by adsorption and photocatalysis. J. Chem. Technol. Biotechnol. 2020, 95, 2723–2729. [Google Scholar] [CrossRef]
  58. Mian, M.M.; Liu, G. Recent progress in biochar-supported photocatalysts: Synthesis, role of biochar, and applications. RSC Adv. 2018, 8, 14237–14248. [Google Scholar] [CrossRef] [Green Version]
  59. Pinna, M.; Binda, G.; Altomare, M.; Marelli, M.; Dossi, C.; Monticelli, D.; Spanu, D.; Recchia, S. Biochar Nanoparticles over TiO2 Nanotube Arrays: A Green Co-Catalyst to Boost the Photocatalytic Degradation of Organic Pollutants. Catalysts 2021, 11, 1048. [Google Scholar] [CrossRef]
  60. Chen, Q.; Lin, G.; Meng, L.; Zhou, L.; Hu, L.; Nong, J.; Li, Y.; Wang, J.; Hu, K.; Yu, Q. Enhanced photoelectric performance of TiO2 nanotubes sensitized with carbon dots derived from bagasse. Chem. Phys. Lett. 2020, 749, 137428. [Google Scholar] [CrossRef]
  61. Hislop, K.A.; Bolton, J.R. The Photochemical Generation of Hydroxyl Radicals in the UV−vis/Ferrioxalate/H2O2 System. Environ. Sci. Technol. 1999, 33, 3119–3126. [Google Scholar] [CrossRef]
  62. Jeong, J.; Yoon, J. pH effect on OH radical production in photo/ferrioxalate system. Water Res. 2005, 39, 2893–2900. [Google Scholar] [CrossRef] [PubMed]
  63. Houas, A.; Lachheb, H.; Ksibi, M.; Elaloui, E.; Guillard, C.; Herrmann, J.M. Photocatalytic degradation pathway of methylene blue in water. Appl. Catal. B Environ. 2001, 31, 145–157. [Google Scholar] [CrossRef]
  64. Vaghela, S.S.; Jethva, A.D.; Mehta, B.B.; Dave, S.P.; Adimurthy, S.; Ramachandraiah, G. Laboratory Studies of Electrochemical Treatment of Industrial Azo Dye Effluent. Environ. Sci. Technol. 2005, 39, 2848–2855. [Google Scholar] [CrossRef] [PubMed]
  65. Bell, J.; Buckley, C.A. Treatment of a textile dye in the anaerobic baffled reactor. Water SA 2003, 29, 129–134. [Google Scholar] [CrossRef] [Green Version]
  66. Shoukat, R.; Khan, S.J.; Jamal, Y. Hybrid anaerobic-aerobic biological treatment for real textile wastewater. J. Water Process Eng. 2019, 29, 100804. [Google Scholar] [CrossRef]
  67. Banat, F.; Al-Asheh, S.; Al-Makhadmeh, L. Utilization of raw and activated date pits for the removal of phenol from aqueous solutions. Chem. Eng. Technol. 2004, 27, 80–86. [Google Scholar] [CrossRef]
  68. Aragaw, T.A.; Bogale, F.M. Biomass-Based Adsorbents for Removal of Dyes From Wastewater: A Review. Front. Environ. Sci. 2021, 9, 558. [Google Scholar] [CrossRef]
  69. McKay, G. Waste colour removal from textile effluents. Am. Dyest. Report. 1980, 68, 29–36. [Google Scholar]
  70. Saleem, J.; Shahid, U.B.; Hijab, M.; Mackey, H.; McKay, G. Production and applications of activated carbons as adsorbents from olive stones. Biomass Convers. Biorefinery 2019, 9, 775–802. [Google Scholar] [CrossRef] [Green Version]
  71. Cooney, D. Adsorption Design for Wastewater Treatment; CRC Press; Lewis Publishers: Boca Raton, FL, USA, 1998; ISBN 978-1566703338. [Google Scholar]
  72. Dotto, G.L.; McKay, G. Current scenario and challenges in adsorption for water treatment. J. Environ. Chem. Eng. 2020, 8, 103988. [Google Scholar] [CrossRef]
  73. Zhu, S.; Khan, M.A.; Kameda, T.; Xu, H.; Wang, F.; Xia, M.; Yoshioka, T. New insights into the capture performance and mechanism of hazardous metals Cr3+ and Cd2+ onto an effective layered double hydroxide based material. J. Hazard. Mater. 2022, 426, 128062. [Google Scholar] [CrossRef] [PubMed]
  74. Zhu, S.; Chen, Y.; Khan, M.A.; Xu, H.; Wang, F.; Xia, M. In-Depth Study of Heavy Metal Removal by an Etidronic Acid-Functionalized Layered Double Hydroxide. ACS Appl. Mater. Interfaces 2022, 14, 7450–7463. [Google Scholar] [CrossRef] [PubMed]
  75. Zhu, S.; Khan, M.A.; Wang, F.; Bano, Z.; Xia, M. Exploration of adsorption mechanism of 2-phosphonobutane-1,2,4-tricarboxylic acid onto kaolinite and montmorillonite via batch experiment and theoretical studies. J. Hazard. Mater. 2021, 403, 123810. [Google Scholar] [CrossRef] [PubMed]
  76. Yang, R.T. Adsorbents: Fundamentals and Applications; John Wiley & Sons: New York, NY, USA, 2003; ISBN 9780471297413. [Google Scholar]
  77. Crini, G. Non-conventional low-cost adsorbents for dye removal: A review. Bioresour. Technol. 2006, 97, 1061–1085. [Google Scholar] [CrossRef] [PubMed]
  78. Abdolali, A.; Guo, W.S.; Ngo, H.H.; Chen, S.S.; Nguyen, N.C.; Tung, K.L. Typical lignocellulosic wastes and by-products for biosorption process in water and wastewater treatment: A critical review. Bioresour. Technol. 2014, 160, 57–66. [Google Scholar] [CrossRef] [PubMed]
  79. Gonzalez, M.T.; Rodriguez-Reinoso, F. Garcia, A.N.; Marcilla, A. CO2 activation of olive stones carbonized different experimental conditions. Carbons 1997, 35, 159–162. [Google Scholar] [CrossRef]
  80. Rodriguez-Reinoso, F.; Molina-Sabio, M. Activated carbons from lignocellulosic materials by chemical and or physical activation—an Overview. Carbon N. Y. 1992, 30, 1111–1118. [Google Scholar] [CrossRef]
  81. Allen, S.J.; McKay, G.; Khader, K.Y.H. Equilibrium adsorption isotherms for basic dyes onto lignite. J. Chem. Technol. Biotechnol. 1989, 45, 291–302. [Google Scholar] [CrossRef]
  82. Rouquerol, J.; Sing, K.S.W. Adsorption by Powders and Porous Solids: Principles, Methodology And Applications; Academic Press: San Diego, CA, USA, 1999. [Google Scholar]
  83. Lam, K.F.; Yeung, K.L.; McKay, G. A rational approach in the design of mesoporous adsorbents. Langmuir 2006, 22, 9632–9641. [Google Scholar] [CrossRef]
  84. Parthasarathy, P.; Al-Ansari, T.; Mackey, H.R.; Sheeba Narayanan, K.; McKay, G. A review on prominent animal and municipal wastes as potential feedstocks for solar pyrolysis for biochar production. Fuel 2022, 316, 123378. [Google Scholar] [CrossRef]
  85. Tomczyk, A.; Sokołowska, Z.; Boguta, P. Biochar physicochemical properties: Pyrolysis temperature and feedstock kind effects. Rev. Environ. Sci. Biotechnol. 2020, 19, 191–215. [Google Scholar] [CrossRef] [Green Version]
  86. Kumar, A.; Balouch, A.; Abdullah. Remediation of toxic fluoride from aqueous media by various techniques. Int. J. Environ. Anal. Chem. 2019, 101, 482–505. [Google Scholar] [CrossRef]
  87. Mui, E.L.K.; Cheung, W.H.; Valix, M.; McKay, G. Dye adsorption onto activated carbons from tyre rubber waste using surface coverage analysis. J. Colloid Interface Sci. 2010, 347, 290–300. [Google Scholar] [CrossRef] [PubMed]
  88. Osagie, C.; Othmani, A.; Ghosh, S.; Malloum, A.; Kashitarash Esfahani, Z.; Ahmadi, S. Dyes adsorption from aqueous media through the nanotechnology: A review. J. Mater. Res. Technol. 2021, 14, 2195–2218. [Google Scholar] [CrossRef]
  89. Mohan, D.; Sarswat, A.; Ok, Y.S.; Pittman, C.U. Organic and inorganic contaminants removal from water with biochar, a renewable, low cost and sustainable adsorbent—A critical review. Bioresour. Technol. 2014, 160, 191–202. [Google Scholar] [CrossRef]
  90. Mui, E.L.K.; Cheung, W.H.; Valix, M.; McKay, G. Dye adsorption onto char from bamboo. J. Hazard. Mater. 2010, 177, 1001–1005. [Google Scholar] [CrossRef]
  91. Patel, M.; Kumar, R.; Pittman, C.U.; Mohan, D. Ciprofloxacin and acetaminophen sorption onto banana peel biochars: Environmental and process parameter influences. Environ. Res. 2021, 201, 111218. [Google Scholar] [CrossRef]
  92. Vimal, V.; Patel, M.; Mohan, D. Aqueous carbofuran removal using slow pyrolyzed sugarcane bagasse biochar: Equilibrium and fixed-bed studies. RSC Adv. 2019, 9, 26338–26350. [Google Scholar] [CrossRef] [Green Version]
  93. Gupta, V.K.; Tyagi, I.; Agarwal, S.; Singh, R.; Chaudhary, M.; Harit, A.; Kushwaha, S. Column operation studies for the removal of dyes and phenols using a low cost adsorbent. Glob. J. Environ. Sci. Manag. 2016, 2, 1–10. [Google Scholar] [CrossRef]
  94. Yang, G.; Wu, L.; Xian, Q.; Shen, F.; Wu, J.; Zhang, Y. Removal of Congo Red and Methylene Blue from Aqueous Solutions by Vermicompost-Derived Biochars. PLoS ONE 2016, 11, e0154562. [Google Scholar] [CrossRef]
  95. Sewu, D.D.; Boakye, P.; Woo, S.H. Highly efficient adsorption of cationic dye by biochar produced with Korean cabbage waste. Bioresour. Technol. 2017, 224, 206–213. [Google Scholar] [CrossRef] [PubMed]
  96. Nautiyal, P.; Subramanian, K.A.; Dastidar, M.G. Adsorptive removal of dye using biochar derived from residual algae after in-situ transesterification: Alternate use of waste of biodiesel industry. J. Environ. Manage. 2016, 182, 187–197. [Google Scholar] [CrossRef] [PubMed]
  97. Khan, N.; Chowdhary, P.; Ahmad, A.; Shekher Giri, B.; Chaturvedi, P. Hydrothermal liquefaction of rice husk and cow dung in Mixed-Bed-Rotating Pyrolyzer and application of biochar for dye removal. Bioresour. Technol. 2020, 309, 123294. [Google Scholar] [CrossRef] [PubMed]
  98. Huang, W.; Zhang, M.; Wang, Y.; Chen, J.; Zhang, J. Biochars prepared from rabbit manure for the adsorption of rhodamine B and Congo red: Characterisation, kinetics, isotherms and thermodynamic studies. Water Sci. Technol. 2020, 81, 436–444. [Google Scholar] [CrossRef]
  99. Vigneshwaran, S.; Sirajudheen, P.; Karthikeyan, P.; Meenakshi, S. Fabrication of sulfur-doped biochar derived from tapioca peel waste with superior adsorption performance for the removal of Malachite green and Rhodamine B dyes. Surfaces Interfaces 2021, 23, 100920. [Google Scholar] [CrossRef]
  100. Ganguly, P.; Sarkhel, R.; Das, P. Synthesis of pyrolyzed biochar and its application for dye removal: Batch, kinetic and isotherm with linear and non-linear mathematical analysis. Surfaces Interfaces 2020, 20, 100616. [Google Scholar] [CrossRef]
  101. Dai, L.; Zhu, W.; He, L.; Tan, F.; Zhu, N.; Zhou, Q.; He, M.; Hu, G. Calcium-rich biochar from crab shell: An unexpected super adsorbent for dye removal. Bioresour. Technol. 2018, 267, 510–516. [Google Scholar] [CrossRef]
  102. Yang, S.S.; Chen, Y.D.; Kang, J.H.; Xie, T.R.; He, L.; Xing, D.F.; Ren, N.Q.; Ho, S.H.; Wu, W.M. Generation of high-efficient biochar for dye adsorption using frass of yellow mealworms (larvae of Tenebrio molitor Linnaeus) fed with wheat straw for insect biomass production. J. Clean. Prod. 2019, 227, 33–47. [Google Scholar] [CrossRef]
  103. Di Chen, Y.; Lin, Y.C.; Ho, S.H.; Zhou, Y.; Ren, N. qi Highly efficient adsorption of dyes by biochar derived from pigments-extracted macroalgae pyrolyzed at different temperature. Bioresour. Technol. 2018, 259, 104–110. [Google Scholar] [CrossRef]
  104. Zubair, M.; Mu’azu, N.D.; Jarrah, N.; Blaisi, N.I.; Aziz, H.A.; Al-Harthi, A.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, 1–19. [Google Scholar] [CrossRef]
  105. Yu, K.L.; Lee, X.J.; Ong, H.C.; Chen, W.H.; Chang, J.S.; Lin, C.S.; Show, P.L.; Ling, T.C. Adsorptive removal of cationic methylene blue and anionic Congo red dyes using wet-torrefied microalgal biochar: Equilibrium, kinetic and mechanism modeling. Environ. Pollut. 2021, 272, 115986. [Google Scholar] [CrossRef] [PubMed]
  106. Yin, Z.; Liu, N.; Bian, S.; Li, J.; Xu, S.; Zhang, Y. Enhancing the adsorption capability of areca leaf biochar for methylene blue by K2FeO4-catalyzed oxidative pyrolysis at low temperature. RSC Adv. 2019, 9, 42343–42350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. dos Santos, K.J.L.; dos Santos, G.E.; de Sá, Í.M.G.L.; Ide, A.H.; Duarte, J.L.d.S.; de Carvalho, S.H.V.; Soletti, J.I.; Meili, L. Wodyetia bifurcata biochar for methylene blue removal from aqueous matrix. Bioresour. Technol. 2019, 293, 122093. [Google Scholar] [CrossRef] [PubMed]
  108. Park, J.H.; Wang, J.J.; Meng, Y.; Wei, Z.; DeLaune, R.D.; Seo, D.C. Adsorption/desorption behavior of cationic and anionic dyes by biochars prepared at normal and high pyrolysis temperatures. Colloids Surfaces A Physicochem. Eng. Asp. 2019, 572, 274–282. [Google Scholar] [CrossRef]
  109. Biswas, S.; Mohapatra, S.S.; Kumari, U.; Meikap, B.C.; Sen, T.K. Batch and continuous closed circuit semi-fluidized bed operation: Removal of MB dye using sugarcane bagasse biochar and alginate composite adsorbents. J. Environ. Chem. Eng. 2020, 8, 103637. [Google Scholar] [CrossRef]
  110. Vyavahare, G.; Jadhav, P.; Jadhav, J.; Patil, R.; Aware, C.; Patil, D.; Gophane, A.; Yang, Y.H.; Gurav, R. Strategies for crystal violet dye sorption on biochar derived from mango leaves and evaluation of residual dye toxicity. J. Clean. Prod. 2019, 207, 296–305. [Google Scholar] [CrossRef]
  111. Côrtes, L.N.; Druzian, S.P.; Streit, A.F.M.; Godinho, M.; Perondi, D.; Collazzo, G.C.; Oliveira, M.L.S.; Cadaval, T.R.S.; Dotto, G.L. Biochars from animal wastes as alternative materials to treat colored effluents containing basic red 9. J. Environ. Chem. Eng. 2019, 7, 103446. [Google Scholar] [CrossRef]
  112. Lafi, W.K. Production of activated carbon from acorns and olive seeds. Biomass Bioenergy 2001, 20, 57–62. [Google Scholar] [CrossRef]
  113. 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]
  114. Sayğili, H.; Güzel, F. Performance of new mesoporous carbon sorbent prepared from grape industrial processing wastes for malachite green and congo red removal. Chem. Eng. Res. Des. 2015, 100, 27–38. [Google Scholar] [CrossRef]
  115. Salamat, S.; Hadavifar, M.; Rezaei, H. Preparation of nanochitosan-STP from shrimp shell and its application in removing of malachite green from aqueous solutions. J. Environ. Chem. Eng. 2019, 7, 103328. [Google Scholar] [CrossRef]
  116. Li, Z.; Chen, K.; Chen, Z.; Li, W.; Biney, B.W.; Guo, A.; Liu, D. Removal of malachite green dye from aqueous solution by adsorbents derived from polyurethane plastic waste. J. Environ. Chem. Eng. 2021, 9, 104704. [Google Scholar] [CrossRef]
  117. Danish, M.; Ahmad, T.; Hashim, R.; Said, N.; Akhtar, M.N.; Mohamad-Saleh, J.; Sulaiman, O. Comparison of surface properties of wood biomass activated carbons and their application against rhodamine B and methylene blue dye. Surfaces Interfaces 2018, 11, 1–13. [Google Scholar] [CrossRef]
  118. Da Silva Lacerda, V.; López-Sotelo, J.B.; Correa-Guimarães, A.; Hernández-Navarro, S.; Sánchez-Báscones, M.; Navas-Gracia, L.M.; Martín-Ramos, P.; Martín-Gil, J. Rhodamine B removal with activated carbons obtained from lignocellulosic waste. J. Environ. Manage. 2015, 155, 67–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Fabryanty, R.; Valencia, C.; Soetaredjo, F.E.; Putro, J.N.; Santoso, S.P.; Kurniawan, A.; Ju, Y.H.; Ismadji, S. Removal of crystal violet dye by adsorption using bentonite—alginate composite. J. Environ. Chem. Eng. 2017, 5, 5677–5687. [Google Scholar] [CrossRef] [Green Version]
  120. Pal, A.; Pan, S.; Saha, S. Synergistically improved adsorption of anionic surfactant and crystal violet on chitosan hydrogel beads. Chem. Eng. J. 2013, 217, 426–434. [Google Scholar] [CrossRef]
  121. Duman, O.; Tunç, S.; Gürkan Polat, T. Adsorptive removal of triarylmethane dye (Basic Red 9) from aqueous solution by sepiolite as effective and low-cost adsorbent. Microporous Mesoporous Mater. 2015, 210, 176–184. [Google Scholar] [CrossRef]
  122. Kizilkaya, B. Usage of Biogenic Apatite (Fish Bones) on Removal of Basic Fuchsin Dye from Aqueous Solution. J. Dispers. Sci. Technol. 2012, 33, 1596–1602. [Google Scholar] [CrossRef]
  123. Yek, P.N.Y.; Peng, W.; Wong, C.C.; Liew, R.K.; Ho, Y.L.; Wan Mahari, W.A.; Azwar, E.; Yuan, T.Q.; Tabatabaei, M.; Aghbashlo, M.; et al. Engineered biochar via microwave CO2 and steam pyrolysis to treat carcinogenic Congo red dye. J. Hazard. Mater. 2020, 395, 122636. [Google Scholar] [CrossRef]
  124. Das, L.; Sengupta, S.; Das, P.; Bhowal, A.; Bhattacharjee, C. Experimental and Numerical modeling on dye adsorption using pyrolyzed mesoporous biochar in Batch and fixed-bed column reactor: Isotherm, Thermodynamics, Mass transfer, Kinetic analysis. Surfaces Interfaces 2021, 23, 100985. [Google Scholar] [CrossRef]
  125. Iqbal, J.; Shah, N.S.; Sayed, M.; Niazi, N.K.; Imran, M.; Khan, J.A.; Khan, Z.U.H.; Hussien, A.G.S.; Polychronopoulou, K.; Howari, F. Nano-zerovalent manganese/biochar composite for the adsorptive and oxidative removal of Congo-red dye from aqueous solutions. J. Hazard. Mater. 2021, 403, 123854. [Google Scholar] [CrossRef] [PubMed]
  126. Iqbal, M.M.; Imran, M.; Hussain, T.; Naeem, M.A.; Al-Kahtani, A.A.; Shah, G.M.; Ahmad, S.; Farooq, A.; Rizwan, M.; Majeed, A.; et al. Effective sequestration of Congo red dye with ZnO/cotton stalks biochar nanocomposite: MODELING, reusability and stability. J. Saudi Chem. Soc. 2021, 25, 101176. [Google Scholar] [CrossRef]
  127. Gurav, R.; Bhatia, S.K.; Choi, T.R.; Choi, Y.K.; Kim, H.J.; Song, H.S.; Lee, S.M.; Lee Park, S.; Lee, H.S.; Koh, J.; et al. Application of macroalgal biomass derived biochar and bioelectrochemical system with Shewanella for the adsorptive removal and biodegradation of toxic azo dye. Chemosphere 2021, 264, 128539. [Google Scholar] [CrossRef] [PubMed]
  128. Gokulan, R.; Avinash, A.; Prabhu, G.G.; Jegan, J. Remediation of remazol dyes by biochar derived from Caulerpa scalpelliformis—An eco-friendly approach. J. Environ. Chem. Eng. 2019, 7, 103297. [Google Scholar] [CrossRef]
  129. Zhang, Z.; Wang, G.; Li, W.; Zhang, L.; Chen, T.; Ding, L. Degradation of methyl orange through hydroxyl radical generated by optically excited biochar: Performance and mechanism. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 601, 125034. [Google Scholar] [CrossRef]
  130. Janoš, P.; Buchtová, H.; Rýznarová, M. Sorption of dyes from aqueous solutions onto fly ash. Water Res. 2003, 37, 4938–4944. [Google Scholar] [CrossRef]
  131. Janoš, P.; Šedivý, P.; Rýznarová, M.; Grötschelová, S. Sorption of basic and acid dyes from aqueous solutions onto oxihumolite. Chemosphere 2005, 59, 881–886. [Google Scholar] [CrossRef]
  132. Janoš, P.; Coskun, S.; Pilařová, V.; Rejnek, J. Removal of basic (Methylene Blue) and acid (Egacid Orange) dyes from waters by sorption on chemically treated wood shavings. Bioresour. Technol. 2009, 100, 1450–1453. [Google Scholar] [CrossRef]
  133. Bouhemal, N.; Addoun, F. Adsorption of dyes from aqueous solution onto activated carbons prepared from date pits: The effect of adsorbents pore size distribution. Desalin. Water Treat. 2009, 7, 242–250. [Google Scholar] [CrossRef]
  134. Bouchemal, N.; Azoudj, Y.; Merzougui, Z.; Addoun, F. Adsorption modeling of orange G dye on mesoporous activated carbon prepared from algerian date pits using experimental designs. Desalin. Water Treat. 2012, 45, 284–290. [Google Scholar] [CrossRef]
  135. Li, Y.; Meas, A.; Shan, S.; Yang, R.; Gai, X. Production and optimization of bamboo hydrochars for adsorption of Congo red and 2-naphthol. Bioresour. Technol. 2016, 207, 379–386. [Google Scholar] [CrossRef] [PubMed]
  136. Li, C.; Zhang, L.; Xia, H.; Peng, J.; Zhang, S.; Cheng, S.; Shu, J. Kinetics and isotherms studies for congo red adsorption on mesoporous Eupatorium adenophorum-based activated carbon via microwave-induced H3PO4 activation. J. Mol. Liq. 2016, 224, 737–744. [Google Scholar] [CrossRef]
  137. Absalan, G.; Asadi, M.; Kamran, S.; Sheikhian, L.; Goltz, D.M. Removal of reactive red-120 and 4-(2-pyridylazo) resorcinol from aqueous samples by Fe3O4 magnetic nanoparticles using ionic liquid as modifier. J. Hazard. Mater. 2011, 192, 476–484. [Google Scholar] [CrossRef] [PubMed]
  138. Oueslati, K.; Lima, E.C.; Ayachi, F.; Cunha, M.R.; Ben Lamine, A. Modeling the removal of Reactive Red 120 dye from aqueous effluents by activated carbon. Water Sci. Technol. 2020, 82, 651–662. [Google Scholar] [CrossRef] [PubMed]
  139. Ahmad, M.A.; Rahman, N.K. Equilibrium, kinetics and thermodynamic of Remazol Brilliant Orange 3R dye adsorption on coffee husk-based activated carbon. Chem. Eng. J. 2011, 170, 154–161. [Google Scholar] [CrossRef]
  140. Rápó, E.; Posta, K.; Suciu, M.; Szép, R.; Tonk, S. Adsorptive Removal of Remazol Brilliant Violet-5R Dye from Aqueous Solutions using Calcined Eggshell as Biosorbent. Acta Chim. Slov. 2019, 66, 648–658. [Google Scholar] [CrossRef]
  141. Arulkumar, M.; Sathishkumar, P.; Palvannan, T. Optimization of Orange G dye adsorption by activated carbon of Thespesia populnea pods using response surface method-ology. J. Hazard. Mater. 2011, 186, 827–834. [Google Scholar] [CrossRef]
  142. Kundu, S.; Chowdhury, I.H.; Naskar, M. Synthesis of hexagonal shaped nanoporous carbon for efficient adsorption of methyl orange dye. J. Mol. Liq. 2017, 234, 417–423. [Google Scholar] [CrossRef]
  143. Rahman, N.; Dafader, N.C.; Miah, A.R.; Shahnaz, S. Efficient removal of methyl orange from aqueous solution using amidoxime adsorbent. Int. J. Environ. Stud. 2018, 76, 594–607. [Google Scholar] [CrossRef]
  144. Ghany, H.M.A. Production of a new activated carbon prepared from palm fronds by thermal activation. Int. J. Eng. Technol. Manag. Res. 2019, 6, 34–43. [Google Scholar] [CrossRef]
  145. Vigneshwaran, S.; Sirajudheen, P.; Nikitha, M.; Ramkumar, K.; Meenakshi, S. Facile synthesis of sulfur-doped chitosan/biochar derived from tapioca peel for the removal of organic dyes: Isotherm, kinetics and mechanisms. J. Mol. Liq. 2021, 326, 115303. [Google Scholar] [CrossRef]
  146. Carrier, M.; Hardie, A.G.; Uras, Ü.; Görgens, J.; Knoetze, J. Production of char from vacuum pyrolysis of South-African sugar cane bagasse and its characterization as activated carbon and biochar. J. Anal. Appl. Pyrolysis 2012, 96, 24–32. [Google Scholar] [CrossRef]
  147. Zubair, M.; Manzar, M.S.; Mu’azu, N.D.; Anil, I.; Blaisi, N.I.; Al-Harthi, M.A. Functionalized MgAl-layered hydroxide intercalated date-palm biochar for Enhanced Uptake of Cationic dye: Kinetics, isotherm and thermodynamic studies. Appl. Clay Sci. 2020, 190, 105587. [Google Scholar] [CrossRef]
  148. Yao, X.; Ji, L.; Guo, J.; Ge, S.; Lu, W.; Chen, Y.; Cai, L.; Wang, Y.; Song, W. An abundant porous biochar material derived from wakame (Undaria pinnatifida) with high adsorption performance for three organic dyes. Bioresour. Technol. 2020, 318, 124082. [Google Scholar] [CrossRef] [PubMed]
  149. Palapa, N.R.; Taher, T.; Rahayu, B.R.; Mohadi, R.; Rachmat, A.; Lesbani, A. CuAl LDH/Rice husk biochar composite for enhanced adsorptive removal of cationic dye from aqueous solution. Bull. Chem. React. Eng. Catal. 2020, 15, 525–537. [Google Scholar] [CrossRef]
  150. Wu, J.; Yang, J.; Feng, P.; Huang, G.; Xu, C.; Lin, B. High-efficiency removal of dyes from wastewater by fully recycling litchi peel biochar. Chemosphere 2020, 246, 125734. [Google Scholar] [CrossRef]
  151. Tsai, W.T.; Chen, H.R. Removal of malachite green from aqueous solution using low-cost chlorella-based biomass. J. Hazard. Mater. 2010, 175, 844–849. [Google Scholar] [CrossRef]
  152. Nizam, N.U.M.; Hanafiah, M.M.; Mahmoudi, E.; Halim, A.A.; Mohammad, A.W. The removal of anionic and cationic dyes from an aqueous solution using biomass-based activated carbon. Sci. Rep. 2021, 11, 8623. [Google Scholar] [CrossRef]
  153. Shaikh, W.A.; Islam, R.U.; Chakraborty, S. Stable silver nanoparticle doped mesoporous biochar-based nanocomposite for efficient removal of toxic dyes. J. Environ. Chem. Eng. 2021, 9, 104982. [Google Scholar] [CrossRef]
  154. Hernández-Zamora, M.; Cristiani-Urbina, E.; Martínez-Jerónimo, F.; Perales-Vela, H.V.; Ponce-Noyola, T.; Montes-Horcasitas, M. del C.; Cañizares-Villanueva, R.O. Bioremoval of the azo dye Congo Red by the microalga Chlorella vulgaris. Environ. Sci. Pollut. Res. 2015, 22, 10811–10823. [Google Scholar] [CrossRef]
  155. Chen, S.; Qin, C.; Wang, T.; Chen, F.; Li, X.; Hou, H.; Zhou, M. Study on the adsorption of dyestuffs with different properties by sludge-rice husk biochar: Adsorption capacity, isotherm, kinetic, thermodynamics and mechanism. J. Mol. Liq. 2019, 285, 62–74. [Google Scholar] [CrossRef]
  156. Goswami, M.; Chaturvedi, P.; Kumar Sonwani, R.; Dutta Gupta, A.; Rani Singhania, R.; Shekher Giri, B.; Nath Rai, B.; Singh, H.; Yadav, S.; Sharan Singh, R. Application of Arjuna (Terminalia arjuna) seed biochar in hybrid treatment system for the bioremediation of Congo red dye. Bioresour. Technol. 2020, 307, 123203. [Google Scholar] [CrossRef] [PubMed]
  157. Fawzy, M.A.; Gomaa, M. Use of algal biorefinery waste and waste office paper in the development of xerogels: A low cost and eco-friendly biosorbent for the effective removal of congo red and Fe (II) from aqueous solutions. J. Environ. Manage. 2020, 262, 110380. [Google Scholar] [CrossRef]
  158. Wang, H.; Wang, S.; Gao, Y. Cetyl trimethyl ammonium bromide modified magnetic biochar from pine nut shells for efficient removal of acid chrome blue K. Bioresour. Technol. 2020, 312, 123564. [Google Scholar] [CrossRef] [PubMed]
  159. Hadi, P.; Xu, M.; Ning, C.; Sze Ki Lin, C.; McKay, G. A critical review on preparation, characterization and utilization of sludge-derived activated carbons for wastewater treatment. Chem. Eng. J. 2015, 260, 895–906. [Google Scholar] [CrossRef]
  160. Raj, A.; Yadav, A.; Rawat, A.P.; Singh, A.K.; Kumar, S.; Pandey, A.K.; Sirohi, R.; Pandey, A. Kinetic and thermodynamic investigations of sewage sludge biochar in removal of Remazol Brilliant Blue R dye from aqueous solution and evaluation of residual dyes cytotoxicity. Environ. Technol. Innov. 2021, 23, 101556. [Google Scholar] [CrossRef]
Figure 1. (A) Dye house discharge. (B) River quality affected by dyestuffs.
Figure 1. (A) Dye house discharge. (B) River quality affected by dyestuffs.
Separations 09 00139 g001
Figure 2. Color-producing chromophores or groups.
Figure 2. Color-producing chromophores or groups.
Separations 09 00139 g002
Figure 3. Typical mechanisms for dyeing cellulose.
Figure 3. Typical mechanisms for dyeing cellulose.
Separations 09 00139 g003
Figure 4. Reactive dyeing mechanism on cellulose.
Figure 4. Reactive dyeing mechanism on cellulose.
Separations 09 00139 g004
Figure 5. Disperse Blue 6 dye compound.
Figure 5. Disperse Blue 6 dye compound.
Separations 09 00139 g005
Figure 6. Structure of the vat red 13 dye.
Figure 6. Structure of the vat red 13 dye.
Separations 09 00139 g006
Figure 7. Structure of the direct yellow 24 dye.
Figure 7. Structure of the direct yellow 24 dye.
Separations 09 00139 g007
Figure 8. Structure of basic crystal violet (CV) dye.
Figure 8. Structure of basic crystal violet (CV) dye.
Separations 09 00139 g008
Figure 9. Structure of the acid blue 78 dye.
Figure 9. Structure of the acid blue 78 dye.
Separations 09 00139 g009
Figure 10. Typical structure of chrysoidine azo dye.
Figure 10. Typical structure of chrysoidine azo dye.
Separations 09 00139 g010
Figure 11. Structure of the sulfur blue 7.
Figure 11. Structure of the sulfur blue 7.
Separations 09 00139 g011
Figure 12. Structure of the mauveine A aniline dye.
Figure 12. Structure of the mauveine A aniline dye.
Separations 09 00139 g012
Figure 13. Structure of the dithiolene metal complex dye.
Figure 13. Structure of the dithiolene metal complex dye.
Separations 09 00139 g013
Figure 14. Structure of mordant red 11.
Figure 14. Structure of mordant red 11.
Separations 09 00139 g014
Table 3. Basic dyes (cationic) and their adsorption onto modified (altered) biochars.
Table 3. Basic dyes (cationic) and their adsorption onto modified (altered) biochars.
DyeModified Biochar FeedstockPyrolysis ConditionsPore Volume (cm3/g)BET Surface Area (m2/g)Adsorption Capacity (mg/g) or Dye Removal (%)Isotherm TypeKinetic ModelAdsorbent ParametersMechanismReference
Temperature (K)Heating Rate (K/min)TimepHEquilibrium Time (min)
MBDate palm fronds107320240-70210--7180-[144]
MGTapioca peel + S- doped107310180-14530Langmuir, FreundlichPseudo I-order, Pseudo II-order2–101080-[99]
Rh BTapioca peel + S- doped107310180-14530Langmuir, FreundlichPseudo I-order, Pseudo II-order2–101080-[99]
MBAreca leaf + K2FeO4473560-20250Langmuir, FreundlichPseudo I-order, Pseudo II-order7720Electrostatic attraction[106]
MGChitosan-tapioca peel + S-doped873-120-12050Langmuir, FreundlichPseudo I-order, Pseudo II-order2–12160Electrostatic attraction, Hydrogen bonding[145]
Rh BChitosan-tapioca peel + S-doped873-120-12040Langmuir, FreundlichPseudo I-order, Pseudo II-order2–12160Electrostatic attraction, Hydrogen bonding, π-π interactions[145]
MBSugarcane bagasse + steam1073101200.356570 5220 Langmuir, Freundlich-7.4180-[146]
MBDate palm fronds with Fe/Mn9733240-430300Langmuir, FreundlichPseudo I-order, Pseudo II-order, Intraparticle diffusion, Elovich4–10240Surface adsorption, π-π interactions, Ion exchange, Pore-filling[147]
MBWakame Undaria pinnatifida leaves with calcination107310120-1160840Langmuir, FreundlichPseudo I-order, Pseudo II-order2–12300Surface adsorption, Hydrogen bonding, π-π interactions, Pore-filling[148]
Rh BWakame Undaria pinnatifida leaves with calcination107310120-1160530Langmuir, FreundlichPseudo I-order, Pseudo II-order2–12300Surface adsorption, Hydrogen bonding, π-π interactions, Pore-filling[148]
MGWakame Undaria pinnatifida leaves with calcination107310120-11604065Langmuir, FreundlichPseudo I-order, Pseudo II-order2–12300Surface adsorption, Hydrogen bonding, π-π interactions, Pore-filling[148]
MGCorn straw773-180-35 520Langmuir, Freundlich, TemkinPseudo I-order, Pseudo II-order, Intra diffusion2–920 [76]
MGRice husk + Cu + Al353-600.350200470Langmuir, Freundlich 9200Pore-filling, π- π interactions[149]
MGLitchi peel + HC1123 600.58810102470FreundlichElovich8720Hydrogen bonding, π-π interactions, Pore-filling, Electrostatic interaction[150]
MGSugarcane bagasse + ZnCl21073-1200.02355090FreundlichPseudo II-order8-Boyd[124]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Parthasarathy, P.; Sajjad, S.; Saleem, J.; Alherbawi, M.; Mckay, G. A Review of the Removal of Dyestuffs from Effluents onto Biochar. Separations 2022, 9, 139. https://doi.org/10.3390/separations9060139

AMA Style

Parthasarathy P, Sajjad S, Saleem J, Alherbawi M, Mckay G. A Review of the Removal of Dyestuffs from Effluents onto Biochar. Separations. 2022; 9(6):139. https://doi.org/10.3390/separations9060139

Chicago/Turabian Style

Parthasarathy, Prakash, Samra Sajjad, Junaid Saleem, Mohammad Alherbawi, and Gordon Mckay. 2022. "A Review of the Removal of Dyestuffs from Effluents onto Biochar" Separations 9, no. 6: 139. https://doi.org/10.3390/separations9060139

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