Next Article in Journal
Numerical Characterization of Acoustic Cavitation Bubbles with Respect to the Bubble Size Distribution at Equilibrium
Next Article in Special Issue
Special Issue “Applications of Biosorption in Wastewater Treatment”
Previous Article in Journal
Cold Storage and Temperature Management of Olive Fruit: The Impact on Fruit Physiology and Olive Oil Quality—A Review
Previous Article in Special Issue
Green Synthesis of A Novel MXene–CS Composite Applied in Treatment of Cr(VI) Contaminated Aqueous Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biotransformation of Citrus Waste-II: Bio-Sorbent Materials for Removal of Dyes, Heavy Metals and Toxic Chemicals from Polluted Water

by
Neelima Mahato
1,*,
Pooja Agarwal
2,†,
Debananda Mohapatra
1,†,
Mukty Sinha
3,†,
Archana Dhyani
4,†,
Brajesh Pathak
5,†,
Manwendra K. Tripathi
6,† and
Subramania Angaiah
7
1
School of Chemical Engineering, Yeungnam University, Gyeongsan 38541, Gyeongsanbuk-do, Korea
2
Division of Chemistry, School of Basic and Applied Sciences, Galgotias University, Greater Noida 203 201, Uttar Pradesh, India
3
Department of Medical Devices, National Institute of Pharmaceutical Education and Research, Ahmedabad, Palej, Gandhinagar 382 355, Gujrat, India
4
Department of Applied Sciences, School of Engineering, University of Petroleum and Energy Studies, Dehradun 248 007, Uttarakhand, India
5
Department of Chemistry, Harishchandra PG College, Varanasi 221 005, Uttar Pradesh, India
6
Department of Metallurgical and Materials Engineering, National Institute of Technology, Raipur 492 010, Chhatisgarh, India
7
Electro-Materials Research Laboratory, Centre for Nanoscience and Technology, Pondicherry University, Puducherry 605 014, Puducherry, India
*
Author to whom correspondence should be addressed.
Contributed equally as first author.
Processes 2021, 9(9), 1544; https://doi.org/10.3390/pr9091544
Submission received: 3 August 2021 / Revised: 24 August 2021 / Accepted: 26 August 2021 / Published: 30 August 2021
(This article belongs to the Special Issue Applications of Biosorption in Wastewater Treatment)

Abstract

:
Industrial processes and anthropogenic activities generate huge amounts of wastes in the form of chemicals, such as heavy metals, dyes, fertilizers, pharmaceutically active chemicals, battery effluents and so on. When these chemicals are left untreated and discarded in the ground or surface waters, they not only cause pollution and harm the ecosystem but also cause toxic effects on the health of human beings, animals and food crops. There are several methods of removal of these toxic materials from the wastewaters, and adsorption by bio-sorbents has been demonstrated as one of the most inexpensive, efficient and convenient methods. Citrus is one of the largest grown fruit crops in the tropical and subtropical regions on the planet. After processing of the fruits at food processing industries, approximately half of the fruit mass is discarded as waste, which causes a number of pollution problems. Alternately, this biomass can be converted to bio-sorbents for the removal of harmful and toxic chemicals from the industrial effluents and wastewaters. The first part of this article contains a thorough review on the biotransformation of citrus waste for the production of biofuel and valuable compounds by fermentation involving microorganisms. The second and concluding part reviews the recent progress in biotransformation of citrus waste biomass (that may be remaining post-extraction of valuable compounds/biofuel generation) into efficient adsorbent substrate materials and their adsorption capacities. The article also includes the details of the synthesis process and mechanisms of adsorption processes.

1. Introduction

Citrus is one of the most popular and largest cultivated fruit crops in the tropical and subtropical regions on the planet, with an annual turnover exceeding 110–124 million tons. The fruits are largely processed in the food processing industries, and approximately half of the fruit mass (45–55%) is discarded as waste. The discarded mass consists of peels (flavedo and albedo), pith residue, seeds and parts inappropriate for human consumption. The waste, when directly discarded to the environment, causes huge problems in terms of pollution to the land and aquatic ecosystems, and underground as well as surface water resources. Alternately, this waste can be utilized as a sustainable and renewable natural resource and feedstock in a number of ways to obtain industrially important chemicals and valuable products by utilizing modern extraction methods, chemical processing techniques and biotransformation. On the one hand, utilization of the waste biomass provides an opportunity to produce valuable chemicals from green resources and avoid usage of harsh chemicals, and on the other hand, it helps in protecting the environment from the adverse effects of pollution. The authors have extensively reviewed the nature, properties and different technological processes of citrus waste valorization in their previously published articles [1,2,3,4,5,6,7,8].

1.1. Citrus Waste Pre-Treatment and Disposal

Citrus wastes produced by food processing industries are generally sent to waste-disposal plants, which require substantial transportation costs and availability of suitable sites for waste disposal. The most common waste management methods employed for the citrus wastes are composting, anaerobic digestion, incineration, thermolysis and gasification [9]. Commonly, in small processing units in the underdeveloped or developing countries, the wastewater is not believed to be toxic or harmful as sewage wastewater and is disposed of into drainage. When the dilute wastewaters, disposed of through drainage, reach streams, lakes, tidal waters, ponds or dumping wells, the organic solids carried by the wastewaters begin to decompose. During the process of decomposition, the dissolved oxygen present in the water is consumed. As a result, anaerobic or putrefaction reactions set in, and the aquatic organisms, such as insects, planktons (zooplanktons and phytoplanktons), fish, aquatic organisms, etc., die because of oxygen deficiency in the water [8]. Within the past three decades, there has been a multifold rise in the demand, and consequently, an increase in the production, supply, processing and applications of processed citrus products. At the same time, there has also been a huge rise in the quantities of dilute liquid and solid wastes, which are required to be disposed of appropriately. Large processing plants and units utilize large amounts of chlorinated waters to keep the processing tables, containers, vessels and equipment clean, and hence, the wastewater also contains chlorine. The effluent waters contain varied quantities of peel oil traces, pulp, juice sacs, follicles and organic materials [10]. Depending upon the concentration of materials and source of origin, the wastewater is divided into three categories, as follows:
(a)
Dilute wastewaters: This includes, (i) fruit-rinsing water, (ii) surface condenser water and (iii) water from barometric condensers of evaporators. This disposal contains good quantities of carbohydrates and a low concentration of nitrogen, unlike domestic sewage which has the opposite composition, i.e., low in carbohydrates and rich in nitrogen. This can be released to water bodies without any fear of adverse effects to the ecosystem.
(b)
Wastewater of intermediate concentration: This includes, (i) floor washing, (ii) equipment clean-up water and (iii) sectionizing wastewater. This contains solid waste concentrations ranging from trace amounts to ~2%. This requires some level of treatment prior to disposal to water bodies.
(c)
Concentrated wastewaters: This includes, (i) dripping waters from can closing and filling machines, (ii) effluent from peel-oil centrifugal and (iii) waste alkali from sectionizing or evaporator cleaning. This contains 2–6% of soluble solids and high concentrations of organic materials.
When dilute wastewater is disposed into water bodies, lakes and basins located in far-off places from the residential areas, sandy beds or sandy lands, the liquid in the wastewater can either percolate down into the soil or evaporate easily, leaving behind the solids at the surface. Later, the clogged wastewaters were also suggested for spray irrigation with somewhat diluted concentration to pasture lands [11,12]. One such experiment conducted at pasture land bearing leguminous cover crops has been reported to show good results [13]. However, spray irrigation to the wood lands, pastures and green vegetation showed negative effects. For disposal, usually, several ponds are constructed, sometimes in series, to manage the waste to flow from one pond to another. The discharge into groves or wells may result in defoliation of trees, probably because of the loss of oxygen in the soil around the roots of the trees. Wastes released into the city sewage system may contaminate the underground water sources and cause damage to pumps and piping, clogging of sand beds and foaming in primary settling tanks. Additionally, pumping the wastewaters into the wells is prohibited because of probable contamination of underground water supplies. Furthermore, fermentation causes gaseous build-up inside the dumping wells and sometimes the pressure mounts to an excessive level to blow back and cause fire outbreaks. Some of the workable solutions in such cases are treatment with nitrogen, extensive aeration and chemical flocculation followed by lagooning [14,15,16,17]. Chemical flocculation with lime combined with aeration in order to promote floating of solids has been observed to result in a 64% reduction of suspended solids and up to a 30% reduction in biological oxygen demand (BOD) [18,19,20]. Although this treatment does not reduce the BOD of the wastewaters, it helps in fulfilling the purpose of waste pre-treatment by regulating the pH to a less acidic consistency and clarifying the suspended solid particles. Common methods of citrus waste removal from the food processing plants and industries, and its adverse effects on the environment, are shown in Figure 1.
The semi-solid citrus wastes have very high-water content, and they are difficult to dry through conventional methods. Furthermore, these processes consume huge energy. The most common method of solid and semi-solid biowaste management methods is composting. However, in case of citrus waste, it causes additional problems. Composting or digesting the citrus waste is not a practical choice as these contain large amounts of essential oils, mainly limonene, which inhibits microbial growth and the fermentation process and affects decomposition. Therefore, extraction of oils from the peel waste is very important before disposal to the landfills. Murdock and Allen reported that the oils present in the orange peels are toxic to yeasts [19]. These have been found to exhibit inhibitory effects on the growth of several useful bacteria, yeast and molds, e.g., Bacillus subtilis, Saccharomyces cerevisiae and Aspergillus awamori [20]. Removal of oils from the citrus waste enables application of decomposition methods or anaerobic digestion, incineration, thermolysis and gasification. Detailed illustrations of practical methods of systematic and environment-friendly disposal of citrus wastes and their relative merits are displayed in Figure 2.
Thus, it appears that the extraction of oil from the peel waste streams not only provides useful by-products, but also contributes toward pollution abatement. Citrus wastes also contain large amounts of sugars, which along with moisture, invite bacteria to grow. Decaying waste causes visual displeasure, odoriferous environment and attracts flies. Furthermore, citrus wastes are required to be processed quickly before compositional changes occur. Therefore, waste disposal to the dumping grounds has additional disadvantages. The nitrogen content in the solid waste materials (<0.14%) is quite insufficient to support bacterial decomposition. Therefore, to support decomposition, it robs oxygen from the soil underneath, resulting in a deficiency of nitrogen in the soil. The effect can be witnessed in terms of de-coloration of the vegetation or grasses at or surrounding areas of the dumping ground. The problem can be overcome by adding N2 supplements in the form of chemical fertilizers. In this process, 200–400 pounds of calcium cyanide is added to each ton of ground waste, mixed thoroughly and allowed to dry until crumbly. Besides calcium cyanides, nitrates, ammonium sulfate and super phosphate are also sometimes added [10].

1.2. Pollutants: Dyes, Heavy Metals, Pharmaceutically Active Compounds (PACs) and Other Contaminants

Processes and manufacturing industries, such as metal plating, metal finishing, automotive, semiconductor manufacturing, pulp and paper production, mining operations, ceramics production, tanneries, radiator manufacturing, smelting and alloy manufacturing, battery manufacturing, corrosion of pipes and infrastructures, textiles and dye industries, etc., release a number of harmful and hazardous chemicals into wastewaters and effluents [21,22]. Researchers have reported that more than 700 kinds of pollutants, both organic and inorganic chemicals, mostly toxic and non-biodegradable in nature, are regularly discarded into the water bodies. The non-biodegradable pollutants are persistent in the environment. The heavy metals in the list of pollutants are cadmium, platinum, mercury, copper, lead, chromium, arsenic, antimony, etc., which have been observed to cause adverse health effects, such as gastrointestinal disorders, stomatitis, tumors, hemoglobinuria, ataxia, paralysis, diarrhea, neurological disorders, muscular dystrophy, vomiting, convulsions and so on [23,24,25]. Heavy metal pollution is a serious problem because of the metals’ persistent nature, and their ability to enter and accumulate in the food chain [26,27,28].
Dyes are organic compounds used for coloring textile materials, paper, plastic, paints and synthetic coloring materials. There are about 40,000 dyes and pigments with approximately 7000 different kinds of chemical structures known to chemists. A dye substance has two parts: chromophores and auxochromes. The chromophores impart color whereas the auxochromes impart intensity for the dye. The dyes are classified as acid, base, reactive, direct, disperse, solvent, sulfur, vat, etc., and involve a wide variety of applications as well as application methods. A vast majority of dyes and pigments are non-biodegradable and persistent in nature. Textile and fiber industries employ approximately 10,000 different kinds of dyes for dying and printing of clothes and fabrics. The concentration of the dye bath during the dying processes ranges between 10 and 200 mg L−1, which retains approximately 10–15% of its initial concentration post-dying and released into the effluent [29,30]. The dye concentration of approximately one ppm or less in the wastewaters has been considered as a potential threat to the environment and to human and aquatic lives [29]. Cyanides from industries released into the environment, particularly in wastewaters, have detrimental effects. Cyanides and their complexes have been demonstrated to have bio-accumulative properties that result in ecological deterioration [31]. Most of the industrial wastewaters contain F-CN, the simplest and most toxic form of cyanide, which is formed by dissociation of cyanide complexes during cyanide-based electroplating operations [32]. Chemical methods are usually employed for the conversion of free cyanide (F-CN) into a complex of NH4+ and CHOO, but most of the methods are expensive. Furthermore, they also produce harmful by-products which contribute to environmental contamination as well as being detrimental to the bioremediation processes [33]. Citrus wastes have been found to be useful for the conversion of F-CN into a complex of NH4+ and CHOO. Apart from F-CN, most of the wastewaters contain heavy metals and the presence of the metallic species (Ni, Zn, As, Cr, Hg and Cu), which have been observed to slow down the conversion of F-CN. The latter occurs as the metallic ions become attached to the hydroxyl groups of the absorbent material responsible for the removal. In addition, it has been reported that material obtained after acid hydrolysis of citrus solid wastes increases the catalytic conversion of F-CN by ~3.86-fold compared with the unhydrolyzed solid waste. The conversion has been found to increase linearly with an increase in pH and temperature [33]. Adsorption of heavy metal ions on the adsorbate material is fundamentally driven by electrostatic attraction between the oppositely charged ions. Both the heavy metal ions present in the wastewater and the functional groups (-COOH, -OH) present on the bio-sorbent material carry charges. The electrostatic attraction between the charges results in binding and ion exchange between Na+, Mg2+, K+, Ca2+ and heavy metal ions (M2+). However, the ion exchange predominates due to stronger binding forces between the charged species.
In recent years, pharmaceutically active compounds (PACs) are emerging as one of the most harmful contaminants in the natural and wastewater systems. PACs are the vast range of complex organic chemical formulations which possess a range of medicinal properties and are used in the treatment, control and eradication of diseases in both humans and animals. These include antibiotics, anti-inflammatories, cytotoxins, birth control pills, synthetic hormones and statins. The commonly encountered PACs in surface waters are erythromycin, metronidazole, sulfamethoxazole, trimethoprim, ciprafloxacin, amoxicillin, trimethoprim, tetracyclin, metformin, acetaminophen, diclofenac, ibuprofen, ketoprofen, naproxen, diazepam, fluoxetine and so on [34,35,36,37]. In the past few decades, consumption of pharmaceutical compounds has increased compared with many earlier decades in the century across the globe, resulting in a significant increase in the production of raw materials as well as the final PAC products. Manufacturing of PACs and life-saving drugs by the pharmaceutical companies and processing and packaging factories also releases huge amounts of compounds into the wastewaters. When excreted by humans or animals and disposed of inappropriately, these enter into the environment. In addition, PACs are also released by urban sewage, domestic hospital wastewaters, intensive livestock farming, liquid livestock manure production, sewage sludge from agricultural activities and effluents from sewage treatment plants [38]. The compounds undergo transformation into certain metabolites or breakdown compounds under the influence of temperature, light and vicinity of other chemical ingredients or microbes in nature under the process of biodegradation or photodegradation. Hence, the new molecules produced may sometimes be more toxic compared to their parent PACs. Furthermore, the transformation renders it difficult to monitor the presence of PACs in their original parent formulation in the environment (ground or surface waters). Presence of PACs in significant amounts in the drinking water has become an alarming concern in many countries. These compounds are capable of accumulating in the biological entities, primarily microbes, and making them drug-resistant. As a consequence, they lose their potency to act effectively against disease-causing microorganisms. Apart from drug resistance developments in disease-causing pathogens, PACs in the environment also cause feminization of male fish and amphibian species because of enhanced amounts of natural and synthetic estrogens in the habitat waters. In addition, increased amounts of PACs in irrigation waters to the crops produce food enriched with PACs. Such foods have been found to cause femaleness in human males, multiple organ complications/failure, genetic and hereditary diseases [38,39].
Diclofenac, a non-steroidal anti-inflammatory (NSAID) drug, has been extensively employed in veterinary usage, fever, pain and injury, livestock farming, fisheries and dairy industries, and has resulted in increased amounts of this chemical in the flesh of animals. When the animals die and are fed upon by predatory birds, diclofenac enters into their bodies and causes serious health consequences, leading to death and collapse of their population. Genus Gyps. is most severely affected and has been included in the category of global extinction risk. Some of the highly endangered species from genus Gyps. are oriental white-backed vulture (Gyps bengalensis), long-billed vulture (Gyps indicus) and slender-billed vulture (Gyps tenuirostris), and the decline in their population has been recorded to be more than 95% since the early 1990s. The decline has been observed to continue at an annual rate of 22% to 48% [40,41,42,43]. Diclofenac causes kidney damage, increases in the concentration of uric acid and crystal formation in serum and vital tissues, visceral gout, bone tumor and death of the predatory birds. Decline of predatory birds’ populations in the ecosystem has been reported to cause a severe imbalance in the food chain and an alarming threat to the survival of the human population in Europe and Asia. Absence of the predatory bird population (eagles and vultures), or scavengers of wild and domestic ungulate carcasses in the ecosystem, has led to an increase in the population of feral dogs (Canis familiaris), and consequently, an increased risk of rabies spread in the human population. Furthermore, there has also been an alarming increase in the population of rats (Ratus spp.) and increasing risk of transmission of various diseases observed, such as bubonic plague, brucellosis, tuberculosis and anthrax in humans and livestock. Disappearing eagle and vulture populations have also created a huge challenge for the Parsee and Tibetese communities in India, Nepal and Tibet, as they find it difficult to continue their burial ritual/practices at sky burial sites and towers of silence [39].

1.3. Health Hazards of Pollutants

Metal industries related to mechanical works and battery manufacturing carrying out electroplating, metal plating, etc., release substantial amounts of heavy metals into the wastewaters. Additionally, large amounts of poisonous dyes are released from pigments or printing industries [44]. The pigment and dye industries produce different kinds of dyes, such as methylene blue (MB), oxazine and xanthenes compounds, azo dyes, methyl violet, etc. Among various dyes, methylene blue and heavy metal salts are the most frequently used chemicals for dying silk, wood and cotton [45]. MB is a common cationic dye extensively used in medical, textile and printing industries. The wastewater effluents from these industries contain high amounts of dyes, which can cause severe environmental pollution. It contaminates the water bodies, such as rivers, ponds, lakes, ditches and even the ground waters. These chemicals are toxic and cause severe impacts on human health as they may cause nausea, vomiting, diarrhea, etc., when ingested. Thus, removal of MB from wastewater is of great concern not only from an environmental point of view, but also for the sake of human life. Various methods used for dye removal include adsorption, flocculation, precipitation, ion exchange, electro-kinetic coagulation, ozonization and so on. Among the above-mentioned processes, adsorption is one of the most efficient methods due to its simple design, ease of operation and insensitivity to toxic substances [46]. The molecular structures of some of the popular anionic and cationic dyes are shown in Figure 3. The hazardous effects of dyes and heavy metals on human health, the ecosystem and the environment, and the possible use of citrus waste-derived bio-sorbents for the removal of harmful chemicals from the wastewaters, are illustrated in Figure 4. The permissible limits and the hazardous effects of heavy metals, dyes and other contaminants on human health have been summarized in Table 1.

1.4. Citrus Peel-Derived Adsorbent Materials

The most popular adsorbent for the adsorption process is activated carbon. However, its use is still limited because it is expensive in terms of its high operational cost [57,58,59]. Advanced technology enables scientists to attempt to overcome the cost of the treatment process by using inexpensive, efficient and easily available adsorbents. In the literature, numerous studies have been reported to obtain low-cost activated carbons from agricultural wastes, such as wheat shells, rice husk, tea waste, neem leaf powder, cotton waste, banana peel and orange peel for the removal of heavy metals and other hazardous compounds from wastewaters [60,61,62,63,64,65]. However, the sorption potential of most of these low-cost sorbents is limited. The surface of the bio-sorbent can be modified to enhance its activity. Among many low-cost adsorbents, one such inexpensive and economical adsorbent precursor material is citrus fruit peel [66,67]. Citrus fruit peel has no use after the extraction of essential oils and other valuable chemicals, and thus, it is a commercially valuable and readily available resource for making adsorbent materials. Hence, conversion of citrus fruit peel to low-cost adsorbents serves multiple purposes. The unwanted waste can be converted to value-added products, such as low-cost adsorbent materials for the removal of heavy metals and dyes from aqueous solution. The latter efficiently alleviates the environmental pollution. Types of modification that can be introduced to the biomass-derived bio-sorbents are shown in Figure 5.
In this article, we focus on reviewing the recent progress in the field of citrus biomass research focused on producing energy, biofuels, important chemicals and bio-sorbent materials for the adsorption of dyes, heavy metals and toxic chemicals by processing of citrus waste biomass via biotransformation. Biotransformation is the modification of waste biomass or harmful chemical substances by certain microorganisms or chemical agents to render them either harmless to humans and the environment or for synthesis of useful products for safe consumption. The review has been presented in two parts under the titles “Biotransformation of Citrus Waste-I: Production of Biofuel and Valuable Compounds by Fermentation”, and “Biotransformation of Citrus Waste-II: Bio-sorbent Materials for the Removal of Dyes, Heavy Metals and Toxic Chemicals from Polluted Water”. The first part deals with the synthesis and production of biofuels (ethanol, methane and biodiesel) and valuable compounds, such as organic acids (citric, succinic, pyruvic, lactic, acetic), Vit-C, enzymes, single-cell proteins and prebiotics from fermentation of the citrus wastes [8]. In the second part of the article (present article), we attempt to review current and conventional trends of citrus waste disposal and their relative merits, and recent progress in biotransformation of citrus waste biomass into bio-sorbent material, employing physical, chemical or thermochemical methods for the adsorption of various pollutants, mainly heavy metals and dyes from polluted wastewaters and industrial effluents, and mechanisms and theoretical studies explored in the adsorption processes. The motivation behind this review is to conduct a detailed overview of the recent updates in this area of research and emphasize our focus towards the possibilities of harnessing the hidden potential of obtaining efficient products out of citrus biowaste, which otherwise is discarded in the dumping grounds as waste.

2. Methods of Preparation of Bio-Sorbents

Bio-sorbents from citrus wastes have been developed in a number of ways, such as (a) mechanical shredding/grinding, (b) physicochemical treatment, (c) thermochemical treatment and (d) biochemical methods using enzymes. Based upon the treatment, the different kinds of bio-sorbents obtained from citrus waste can be classified into the following categories: (i) native peel bio-sorbent, (ii) protonated peel bio-sorbent, (iii) peel pectic acid bio-sorbent, (iv) de-pectinated peel bio-sorbent, (v) carbonized peel-activated carbon bio-sorbent, (vi) chemically modified bio-sorbent and (vii) biochemically or enzymatically modified bio-sorbent. The native peel bio-sorbents are obtained from physical or mechanical treatment of citrus peel biowaste. The biomass is washed with tap water to remove dirt, followed by washing with distilled or deionized (DI) water, or nano-pure (NP) water, 3–5 times. Washing is a common step and is essentially included in all the techniques of biomaterial preparation. The native peel bio-sorbent is developed from drying the washed peel waste at 323 K, until it attains a constant weight, followed by grinding and sieving to a desired particle size of ~0.5–1.0 mm, appropriate for adsorption studies. Peel pectic acid is prepared from albedo of the citrus peel waste. The peels are treated with hot acidified water (HCl, pH 1.5) at 358 K and stirred for 2 h. The treatment at elevated temperature facilitates extraction of pectic acid from the peel tissues. The extracted pectic acid is then filtered and coagulated with an equal volume of ethanol (95%). The coagulate is washed with ethanol and dried until constant weight. The peel pectic acid is insoluble in aqueous medium and remains stable during the adsorption tests [40]. The de-pectinated peel adsorbents are obtained from the residue remaining after the extraction of pectin from peel waste. The pectin is mainly extracted from the albedo (white) part of the peels. The residue is washed thoroughly and dried in an oven until constant weight [1,2].
Chemical modification of bio-sorbents is carried out in order to enhance the adsorption capacity by introducing active functional groups by means of chemical reactions. Furthermore, it helps in enhancing the chemical stability and mechanical properties of the sorbent material. The latter prohibits the release of pollutants into the adsorption system. Chemical treatment is carried out via several methods. Some of the well-known methods are: (a) protonation (inorganic acid: HCl, H2SO4, HNO3), (b) alkali saponification (using NaOH, Ca(OH)2, CaCl2), (c) phosphorylation, (d) blocking of functional groups using chemical reagents, (e) organic acids (citric acid, oxalic acid), (f) H2O2 treatment, etc. Protonated peel bio-sorbents are obtained from protonation of the citrus peel waste collected after washing and drying. The protonation is carried out by treating the material with acids, such as HCl or HNO3. For this, the material is suspended in the acid and stirred or shaken for 6–8 h, followed by filtration, washing until neutral pH and drying until constant weight. Protonation is employed to remove excess of cations, such as Ca2+ or Na+, present on the biomaterial surface, which interfere with the metal sorption process. These cations are replaced by protons, which enhances the binding of heavy metals by decreasing the competition between Na+, K+ and Ca2+ ions with heavy metals such as Cd2+, Hg2+, Pb2+, Zn2+, Se2+, As2+ and so on. Replacement of Ca2+ and Na+ ions by protons has exhibited enhancements in the adsorption of desired heavy metal ions and their removal from wastewater. Treatment with NaOH and citric acid is employed to introduce carboxyl groups on the bio-sorbent surface, which interact with heavy metals to form a complex, and the resultant structure helps in the removal of the toxic elements from the wastewater. Phosphorylation is employed to introduce abundant alcoholic “-OH” groups and phosphoric groups into the bio-sorbent material. The latter, possessing a high affinity for ferric iron, enhances the loading capacity for iron. Iqbal et al. carried out experiments to block “-COOH” and “-OH” functional groups using anhydrous CH3OH, and concentrated HCl and HCHO, respectively. The modified bio-sorbents with blocked functional groups were found to exhibit a reduced adsorption capacity of Ni2+ by 78.57% and Zn2+ by 73.31%, confirming the main contribution of carboxyl and hydroxyl functional groups in the adsorption of heavy metal ions [69].
In the carbonization process, the dried peels are subjected to a very high temperature of ~773 K, followed by acid oxidation. An inert atmosphere, such as N2, is employed to prohibit fire or rigorous oxidation. The positively charged amine groups present on the surface of the adsorbent material facilitate binding to anionic RMB reactive dye by electrostatic attraction [70]. Bhatnagar et al. prepared a bio-sorbent from lemon peels by thermal activation at 323 K in the presence of air, which converted the peels to ash. Treating bio-sorbents with H2O2 has been employed to avoid the release of color in bio-sorbents [71]. Treating with 1% NaOH and ethanol removes lignin and colored pigments. Carbonization through the chemical activation method is one of the most favored methods for preparation of adsorbent material from citrus fruit peel. Weight ratios of peel vs. activating agent, temperature and time of carbonization are the selected parameters for optimizing the preparation of an efficient adsorbent material [72]. Generally, the dried citrus fruit peel is fed to a mixer grinder, and the ground powder is mixed with activating agents such as ortho-phosphoric acid, zinc chloride or sulfuric acid. This is then carbonized in a muffle furnace by heating it at an elevated temperature of ~723–823 K, up to a duration of 0.75 to 1.5 h. The weight ratio of dried citrus fruit peel to activating agent varies in the range of 1:1 to 3:1. The charred material is then cooled and washed with dilute ammonia solution and distilled water. This removes any unconverted activating agent from the carbonaceous material. The washing of the sample is continued until the pH becomes neutral. The charred material is then left for drying overnight under ambient conditions. The dried samples are then crushed and fractioned into different sizes [72].
Cross-linked hydrogel adsorbents for the removal of dye molecules can be prepared by treating the ground peel powder with N-vinyl-2 pyrrolidone (NVP), followed by irradiation with gamma rays. Mahmoud et al. demonstrated that a gamma irradiation dose of 30 kGy to the hydrogel precursor composition of NVP and orange peels in the ratio of 1:1 results in optional homogeneity of the bio-sorbent material with appropriate properties for practical applications. Usage of hydrogel (cross-linked polymerized hydrogel) enables the material to adsorb and retain large volumes of water and facilitate in increasing the contact time between the pollutant dyes and adsorbent material by the virtue of its cross-linked three-dimensional network structure. A porous material possesses additional benefits in terms of extended surface area for adsorption [73]. Chemical treatment by formaldehyde and urea carried out by Rabia et al. demonstrated an enhancement in roughness and unevenness, with apparent pores and canals of irregular shapes in a 3D network structure of the material surface to facilitate extended contact time between the adsorbate molecules and adsorbent material, and enhance physisorption [74]. Treating with HCHO improves the shelf-life of the bio-sorbent and prevents microbial damage [75]. While the functional groups, such as carboxylic acid, amine and sulfonic acid groups participate in electrostatic interaction and binding with dye molecules, the bulky structure of the hydrogel can be conveniently collected, separated and regenerated either by washing with water or treating with acid for further usage. Dev et al. reported on the adsorption of selenium from wastewater by citrus peel-based bio-sorbents chemically modified by calcium alginate. The chemical modification provides structural stability to the adsorbent material, which allows its reusability. Besides, it enhances the number of “-COOH” and “-OH” functional groups due to alginate beads, which enables the bio-sorbent material to absorb and remove other metallic ions in addition to SeO42− and SeO32− [76].
Kam and Lee carried out adsorption of amoxicillin onto the activated carbon surface prepared from citrus peel waste from aqueous solution containing the antibiotic, and reported an efficient adsorption within 30 min and attainment of equilibrium in 90 min. The waste citrus peel-based activated carbon showed a maximum adsorption capacity of 125 mg/g of the adsorbent at 293 K [77]. Putra et al. studied adsorption and removal of amoxicillin using commercial activated carbon [78]. Moussavi et al. reported on the adsorption studies of amoxicillin on commercial activated carbon and activated carbon derived from pomegranate wood [79]. Ding et al. developed activated carbon from sewage sludge and oil sludge and reported adsorption and removal of oxytetracycline and chlortetracycline [80]. Baccar et al. obtained activated carbon from olive-waste cake to absorb naproxen, ketoprofen, diclofenac and ibuprofen from aqueous solution containing the contaminants [81]. Ahmed et al. produced activated carbon from Siris seed pods and carried out adsorption and removal of metronidazole from contaminated water [82]. On the other hand, activated carbon developed from vine wood by Pouretedal and Sedech et al. showed efficient adsorption of amoxicillin, cephalexin, tetracycline and penicillin from contaminated water [83].
The adsorption of heavy metal ions onto the adsorbent surface is influenced by several factors, such as nature of the material, charge on the chelating metal ion, size of the ion, nature of donor atom present in the ligand, buffering environment during the adsorption process and exchange of ions, nature and properties of the solid support, and so on [84]. In this direction, Li et al. carried out chemical modification employing 20% isopropyl alcohol to remove coloring compounds from orange peels, along with polar compounds. This was followed by saponification by addition of 0.1 M NaOH/0.1 M NH4OH/saturated solution of Ca(OH)2. The Na+/NH4+ or Ca2+ ions become attached to the cellulose molecules of the adsorbent material and facilitate an ion exchange mechanism between Na+, NH4+ or Ca2+ ions and the bivalent heavy metal ions. In the next step, the saponified orange peels are treated with 0.6 M acid at an elevated temperature of 353 K. The heat is required to produce a condensation product and acid anhydride. The latter combines with cellulose hydroxyl groups and results in the formation of ester linkage and introduction of carbonyl groups to the cellulose molecule. The additional carbonyl functional groups introduced to the cellulose molecule enhance metal ion adsorption [85]. In another method, Liang et al. treated orange peels with NaOH followed by mercapto-acetic acid (C2H4O2S) in order to convert the hydroxyl groups present in the cellulose molecules into mercapto groups. The latter exhibited a higher affinity towards heavy metal adsorption (Cu2+ and Cd2+) from aqueous solutions [86]. The final different sized samples are used for adsorption purposes. The different methods of bio-sorbent pre-treatment by heat, chemical(s) and enzymes are summarized in Figure 6.
Cameron et al. carried out an elaborated study on the adsorption of Pb2+ ions by adsorbents synthesized from citrus peels and peel-derived pectin and concluded that fragmentation of larger molecules into smaller fragments and demethylation of the same occur. The latter plays an important role in the enhancement of the sorption capacity of pectin and derived materials. The fragmentation can be carried out either chemically or enzymatically. Pectin, a polysaccharide present in the citrus peels, can be modified via enzymatic or chemical conversion to develop a suitable bio-sorbent as well as fine-tune the desired properties, e.g., ion exchange and adsorption properties. Pectin polymer is made of galacturonic acid (GA) monomer units, a major sugar found in citrus fruits. It is basically concentrated in the linear homogalacturonan region (HG), which is pectin’s dominant structural domain. The carboxylic acid functional groups present in the GA molecular structure interact with the heavy metal cation present in the wastewater or industrial effluents and require removal. While the polyanionic character of pectin is crucial for adsorption, masking of the negative charge present on the carbonyl group in the GA molecule by means of methyl esterification at C-6 position hence alters the overall functionality of the pectin molecule. In other words, the pectin functionality is dependent upon the total amount of methylation of GA units or degree of methylation (DM) and distribution of methylated GAs and non-methylated GAs (GAs with unmasked or free carboxylic acid functional groups) in the pectin polymeric chain in the HG region [89]. The de-esterification of the GAs or de-methylation of GAs can induce ordered or random distribution of de-methylated GAs in the pectin polymer chain. Both kinds of specific properties obtained post-modification, e.g., degree of methylation (DM) and degree of polymerization (DP), in pectin molecule have been reported to exhibit effects in terms of interaction with cation and sorption properties [87,88]. Pectin extraction can be carried out via aqueous extraction using an aqueous acid or base followed by purification of liquid extracts containing hydrocolloids, and isolating the extracted pectin from the mixture [1,2]. There is a basic difference in the product quality of pectin obtained post-acid/alkali treatment. The pectin obtained from the alkali extraction process contains a low degree of esterification (low DE pectin). The latter results from saponification of the ester groups present in the polymer molecule by alkali. On the other hand, pectin obtained from the acid extraction process contains a high degree of esterification (with high DE pectin ~50% and greater) [89]. The methods of obtaining bio-sorbent materials from citrus peel-derived oligosaccharides and enzymatically modified pectin have been summarized in Figure 7a, and their respective adsorption capacities towards Pb2+ are shown in Figure 7b.

3. Adsorption Experiments and Mechanism

The bio-sorption of harmful chemicals from wastewaters has been focused on achieving two prime targets: (a) development of novel bio-sorbent substrate material, and optimization of the adsorption process in terms of uptake of pollutant molecules/ions in a batch reactor containing a single target metal, and (b) enhancement of sorption uptake capacity by suitable processing [92,93]. From the viewpoint of carrying out adsorption processes/experiments, either or both of the two main techniques are employed, namely, batch adsorption test and/or column adsorption test. In the batch adsorption process, ion exchange has been identified as the main mechanism for the adsorption and removal of heavy metal ions or dye molecules. The carboxyl and hydroxyl functional groups present on the adsorbent substrate have been demonstrated to play key roles in the adsorption process. Other functional groups, such as amide, sulfonate, phosphate and amino groups, have also been reported to participate in the adsorption process. The adsorption process is a complex interplay between a number of mechanisms, such as complexation, coordination, chelation, ion adsorption or exchange, micro-precipitation, electrostatic interaction, H-bonding and so on. In the experimental part, a fixed amount of bio-sorbent substrate/material is placed with a definite volume of solution containing the pollutant or toxic metal ions/dyes, and stirred/shaken for a specified duration in a beaker or flask. After the adsorption process is completed, the solution is filtered and the adsorbent is regenerated or recycled by washing with water or treating with chemicals, such as acids, alkali or organic solvent(s). Alternately, the bio-sorbent material can also be regenerated by physical or thermo-physical treatments, such as heating, microwaving or sonication [93,94,95].
In some cases, the test solution has been observed to develop a brown color, which is explained by leaching out of carotenoids from the citrus peel biomass. Carotenoids are responsible for binding the heavy metal ions present in the test solution and forming complexes. The latter has been found to be unable to adsorb onto the bio-sorbent surface, and hence, left in the test solution imparting color. This problem can be solved by chemical pre-treatment of the bio-sorbent material, which helps in leaching out the carotenoid or other colored materials, such as chlorophyll, from the citrus peel biomass. Treating the biomass with isopropanol helps in the removal of soluble compounds without any adverse effect on the nature of biomass material or ion-binding sites on the surface. Citric acid treatment helps in the dissolution of polysaccharides present in the cell wall of the biomass. This facilitates in opening up the physical structure of the biomass and thereby increases the number of adsorption/binding sites, i.e., functional groups. Alkali treatment has been observed to impart a stronger effect on cell wall rupturing and facilitate the exposure of functional groups. Furthermore, the hydroxyl or carboxylic acid groups are converted to their salt forms, thereby helping to enhance the adsorption process by an ion exchange mechanism. An increase in temperature during alkali or acid pre-treatment has been observed in biomass loss. The factors influencing the bio-sorption process are initial pH of the test solution, concentration of the test solution (heavy metal ions, dye molecules, etc.), dosage of adsorbent, pre-treatment of the bio-sorbent, temperature during the adsorption process and duration of the contact time between the sorbent material and the test solution. The advantages of the batch adsorption technique are short analysis time, low operational costs, simple maintenance and conductance and that it can be operated with locally available adsorbent materials with satisfactory efficiency [96].
A next-level technique, which can be employed for large sample sizes or industrial scale purposes regarding adsorption and removal of pollutants, is fixed-bed reactors or the column adsorption method. It is also called a continuous flow column system. The fixed-bed column is a commercially viable technique, and at present, employed in industries fitted with ion-exchange resins or other commercially available adsorbent materials. The column material is usually made of clear-extruded acrylic. The length of the column ranges between 25 and 30 cm with an internal diameter of 1.3–1.5 cm. The bottom layer of the column is filled with spherical glass beads (of diameter 3 mm) to facilitate even distribution of influent flowing across the column length. A fiber screen is placed at the top of the column to prohibit the scattering of the adsorbent material in the fixed-bed and confine it appropriately in a fixed position during the entire operation. The wet-packing of the column is usually performed in such a way so as to allow a void space inside the column up to 70%. The bed-height is hence adjusted up to 24 cm in order to enhance the length of the fixed-bed, and the column length can be increased by supplementing an additional number of columns in series. During operation, a certain amount of material by weight is wet-packed in the column and the column is conditioned with a suitable solvent (e.g., acidified water of pH 4.0–5.5) overnight. Post-conditioning, the test solution (prepared at the same pH) is percolated through the column at a constant flow rate using a peristaltic pump or a micro-tube pump. The test solution passes through the column contents, i.e., packed adsorption bed, and the latter adsorbs the heavy metal ions or dye molecules, leaving behind the solution with a lesser or minimal concentration of pollutants. If the driving force (concentration variation between sorbent and influent) remains high, the sorbent shows high metal ion uptake. This is due to the adsorbent material being saturated at a relatively high influent concentration, whereas the progressing metal solution comes repeatedly in contact with a fresher and more efficiently adsorbing surface, and thus the effluent leaving the column becomes virtually free of metal ions. The saturated column bed can be regenerated and recycled. Regeneration of the column bed is carried out by eluting the bed with a suitable desorption solution. The effluent samples are collected at a fixed interval of time using a fraction collector [97,98]. The schematic representation of the experimental set-up for the fixed-bed adsorption column is shown in Figure 8. The important reported results of experiments on bio-sorbent materials developed from citrus peel waste are listed in Table 2, Table 3 and Table 4. The various mechanisms of adsorption of heavy metals and dyes on the biomass derived bio-sorbent surface are shown in Figure 9.

4. Kinetics and Thermodynamics

4.1. Kinetics

The adsorption kinetics helps in defining the rate of efficiency of adsorption. Kinetic parameters are useful for designing and modeling the bio-sorption processes. Several types of kinetic models have been proposed by researchers to study the mechanism and rate-regulating steps. The Lagergren’s rate equation is one of the most commonly used models to explain the adsorption of an adsorbate from the liquid phase [143]. The linear form of the pseudo-first-order equation is given as [144]:
ln ( q e q t ) =   lnq e k 1 t
where, ‘qt’ and ‘qe’ are amount of adsorbed material (mg/g) at time ‘t’ and at equilibrium, respectively. ‘k1’ (min−1) is the rate constant for the pseudo-first-order reaction. The pseudo-second-order kinetics explain the involvement of both the adsorbate and adsorbent in the rate-limiting step [113,144]. The equation for the linearized pseudo-second-order reaction is given as:
t q t = 1 k 2   q e 2 + t q e
where, ‘k2′ is the rate constant (g/mg−1 min−1) of the second-order kinetics [113,145]. The initial adsorption rate ‘h’ (mg/g min−1) is defined as:
h   =   k 2 q 2 e
the values of ‘k2′ and ‘h’ can be obtained from the intercept of the plot based on the second-order equation [146]. The kinetics defined by Elovich’s model is based on the principle that adsorption sites increase exponentially with progress in the adsorption process, suggesting multilayer adsorption [147]. The linear form of Elovich’s equation is given as:
q t = 1 β ln ( α β ) + 1 β ln t
where, ‘α’ is the initial rate of adsorption (mg/g−1 min−1) and ‘β’ is related to surface coverage (g/mg−1). Weber and Morris proposed an intra-particle diffusion model [117,148]. The equation for the linear form of this model is given as:
q t =   kt 1 / 2 + C
where, ‘C’ is the intercept and ‘k’ is the intra-particle diffusion constant. The value of ‘k’ can be calculated from the slope of the linear plot of ‘qt’ vs. ‘t1/2’. If intra-particle diffusion is involved in adsorption, then there would be a linear plot for of ‘qt’ against ‘t1/2’. In cases where the line passes through the origin, it shows that intra-particle diffusion is the rate-controlling step [132].

4.2. Thermodynamic Observations

Appropriate study and explanation of adsorption isotherms is very significant and crucial for the overall development of the adsorption mechanism and effective design of the adsorption system. It helps to explain the mechanism of interaction between adsorbate molecules with the adsorbent surface. There are many models and operational designs available to understand the batch adsorption system. The most commonly employed methods are Langmuir and Freundlich models. The Langmuir adsorption isotherm model explains monolayer adsorption equilibrium between the adsorbate and the adsorbent [149]. This model is suitable for explaining the chemisorption when there is covalent or ionic bond formation between the adsorbate and the adsorbent. Many systems followed the equation to explain the binary adsorption system. The Langmuir model in its linear form may be expressed as:
1 q e = 1 K L C e q m + 1 q m
In this equation, ‘qe’ is adsorption capacity (mg/g−1) at equilibrium, ‘Ce’ is the equilibrium concentration (mg L−1) of the adsorbate, ‘qm’ is maximum adsorption capacity (mg/g−1) and ‘KL’ is the Langmuir constant (L mg−1). The Freundlich isotherm explains the multi-layered adsorption phenomenon. It is applicable for reversible adsorption of the adsorbate on the surface of the adsorbent [120,150]. It states that the surface of the adsorbent should be heterogeneous in nature for multilayer adsorption [115,145]. This model states that the surface of the adsorbent has a diverse binding energy spectrum. The linear form of the Freundlich isotherm equation can be expressed as:
log q e =   logK F + 1 n logC e
where, ‘KF’ is the Freundlich isotherm constant (mg1−1/n L1/n g−1), and shows the adsorption efficiency of per unit mass of adsorbent. The 1/n value expresses the heterogeneity factor.
Thermodynamic parameters, e.g., entropy change, enthalpy change and standard free energy, are signification parameters to assess and evaluate the viability of the adsorption process along with the nature of adsorption. The negative value of change in enthalpy (ΔH°) shows the exothermic nature of the adsorption process, while the positive value of change in entropy (ΔS°) stipulates the increased randomness of the process at the interface. It explains that the process is entropy-driven. The Gibbs free energy change of the adsorption process is subsequent to ‘Kc’ and given by the following equation:
Δ G ° = RTlnK c
Here, ‘Kc’ can be expressed as shown in the following equation:
lnK c = Δ S ° R Δ H ° RT
Standard free energy (ΔG°), enthalpy change (ΔH°) and entropy change (ΔS°) can be determined by using the following equation:
Δ G ° =   Δ H ° T Δ S °
where, ‘R’ is universal gas constant (8.314 J.mol−1 K−1), ‘T’ is temperature in Kelvin and ‘Kc’ is equilibrium constant. Change in enthalpy (ΔH°) and entropy (ΔS°) can be calculated from the slope of the plot of (ΔG°) vs. T. Reported values of ΔH° for physical adsorption range from −4 to −40 kJ.mol−1. Bhatnagar et al. has calculated thermodynamic parameters to check the adsorption nature of cobalt using lemon peel as a bio-sorbent. The value of ΔH (−21.2 kJ.mol−1) found in the range shows the physical adsorption [119]. The value of ΔG calculated indicates the spontaneity in the process. A thermodynamic study of the bio-sorption of methylene blue from C. sinensis bagasse was observed by Bhatti et al., and the calculated value of ΔH° (51.9 kJ/mol) shows a similar physical adsorption in the process as a purely physical or chemical one [132]. The role of physisorption can be explained on the basis of the heat involved, which is >40 kJ.mol−1, whereas, for chemisorption, it is reported in a range of 80–200 kJ.mol−1. Similar results have been reported for the adsorption of Reamzol Brilliant Blue using an orange peel adsorbent [137]. The process was efficient as the negative value of free energy denotes the feasibility of the process. Additionally, the positive values of ΔH and ΔS shown are in favor of the adsorption process. These results also show the affinity of the adsorbent towards the dye. However, the adsorption of dyes has been reported as an exothermic phenomenon in many studies. Bio-sorption of La and Ce using peels of Citrus reticulata was also found to be a thermodynamically feasible and spontaneous process. It is shown to be a process of endothermic nature in the temperature range of 293–323 K, and the overall entropy increases due to the exchange of the metal ions with more mobile ions [107]. Malachite green dye adsorbed by Citrus grandis peels revealed the change in ΔG° from −21.55 to −24.22 kJ.mol−1, in the temperature range of 303 to 333 K, indicating enhanced spontaneity at high temperatures [140]. Similar results were also observed during the removal of fluoride using a Citrus limetta peels adsorbent activated with FeCl3 [56]. A thermodynamic study for the removal of methylene blue dye using Citrus limetta peel waste exhibited Gibbs free energy (ΔG°) values in favor of the process [48] (Table 5). Kaffir lime peels were used to reduce graphene oxide to prepare reduced graphene oxide (RGO), and applied for the absorption of methylene blue [151]. Several kinetic models fitted for the citrus peel as adsorbent have been shown in Table 5.

5. Design of Experiments

Response surface methodology (RSM) is a very popular tool for the optimization of process variables. It has been adopted in various studies for the design and analysis of the experiments. Principally, it is a mathematical and statistical technique for the design of experiments using relations between a cluster of controlled experimental variables and the measured properties, created on one or more selected conditions [156]. Numerical or physical experimental data are calculated by an expression that is generally a low-order polynomial. Usually, a second-order polynomial equation is fitted to analyze the experimental data by means of RSM, which can be represented as:
Y   =   b 0 + i   = 1 n b i x i + ( i   = 1 n b ii x i ) 2 + ( i   = 1 n 1 j   =   i + 1 n b ij x i x j ) + ε  
The results are obtained as 2D contours and 3D plots. This method is very competent, and uses the experimental data and interactions between the factors [157,158]. This process is based on three key steps, which involve statistically designed experiments, determination of the coefficients through estimation of response via mathematical modeling and investigating the competency of the model [159]. The ANOVA program is used to calculate the statistical parameters along with the optimization of independent parameters and dependent output responses. Dutta et al. analyzed the result of each run and correlated the responses with three individual factors for preparation of an adsorbent using an empirical second-degree polynomial, as shown above. Optimized conditions obtained as responses for carbonization of citrus fruit peel were weight ratio of the peel to the activating agent, temperature of carbonization and time of carbonization, which have the values of 3:1, 798 K and 0.75 h, respectively (Table 5) [72]. An experimental design for the removal of MB dye by charred citrus fruit peel has also been attained. Numeric parameters selected were initial concentration of MB, amount of adsorbent and pH of the solution, and results obtained from statistical design were maintained during the experiment and found to be fitted for the removal of dye [72]. The model adequacy of the Cr(VI) adsorption by Musambi peels was also found to be statistically viable [152]. In recent years, RSM modeling has been applied and reported in several adsorbent-based materials. It is an efficient and useful procedure which can help maximize the performance along with responses based on combinations of variables. Therefore, RSM offers an extensive scope for modeling the parameters’ optimization along with the percent removal of heavy metal/dye by citrus peel waste-based adsorbent materials. Some of the experimental models applied for citrus peel adsorbents are recorded in Table 6.

6. Summary and Conclusions

Biotransformation of citrus waste into valuable compounds and adsorbent substrate materials for the purpose of adsorption of heavy metals, dyes and toxic chemicals from industrial wastewaters is among hugely adopted research projects around the world. Hazards of pollution in water are not only restricted to the aquatic ecosystem but are also found to spread to the underground water tables, crops and crop products, human/livestock/birds’ health and microbial ecosystems on the land. Adsorbents from citrus wastes can be developed by a number of methods, namely physical processes, chemical methods, thermal and thermo-chemical techniques. Protonated adsorbents have demonstrated efficiencies better than native peel bio-sorbents. In addition, chemically treated bio-sorbents exhibit several advantages, such as greater chemical and mechanical stabilities in the test solution. Furthermore, they help in improving the surface properties with additional functional groups or active adsorption sites and enhance the adsorptive capacities of the resultant sorption materials. Thermochemical activation gives rise to activated carbon materials, which show enhanced porosity to facilitate physisorption along with chemical adsorption. The sorption process is governed by a number mechanisms, such as physical adsorption on the adsorbent surface by van der Waals forces of attraction, hydrogen bonding, dipole-induced dipole moments and electrostatic attraction between charged species, i.e., cationic charges on heavy metal ions and polyanionic charges on the bio-sorbent surface. Two main methods of carrying out sorption processes are popular and widely adopted in experiments: batch adsorption tests and fixed-bed adsorption columns. The latter has an advantage of installing an additional number of columns in order to increase the length of the adsorption bed for enhanced capacity of removal of pollutants from wastewater. Theoretical studies, including kinetics, thermodynamics, simulation and modeling, add a greater in-depth understanding of the adsorption mechanism. Bio-sorbents derived from citrus wastes, the largest fruit crop grown on the planet, provide an inexpensive, natural, renewable and sustainable means of obtaining resourceful as well as fruitful products.

Author Contributions

This work was completed with the contributions of 8 authors. N.M. and P.A. designed and wrote the manuscript; M.S. and A.D. contributed to analyzing recent studies on the valorization of citrus waste and synthesis of bio-adsorbents from citrus peel waste; P.A. compiled the theoretical studies reported in this subject area; B.P. and D.M. contributed to the summary and interpretation of reports and relevance to the current research progress issues; M.K.T. and S.A. performed the final proof-reading of the manuscript; N.M. carried out the final editing, revision and supervision of the project to bring it to its final format. All authors have read and agreed to the published version of the manuscript.

Funding

Authors acknowledge support from the National Research Foundation of Korea (NRF) Korean Government (Ministry of Science and ICT) (NRF-2020R1G1A1015243).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mahato, N.; Sharma, K.; Sinha, M.; Baral, E.R.; Koteswararao, R.; Dhyani, A.; Cho, M.H.; Cho, S. Bio-sorbents, industrially important chemicals and novel materials from citrus processing waste as a sustainable and renewable bioresource: A review. J. Adv. Res. 2020, 23, 61–82. [Google Scholar] [CrossRef] [PubMed]
  2. Mahato, N.; Sinha, M.; Sharma, K.; Koteswararao, R.; Cho, M.H. Cho Modern Extraction and Purification Techniques for Obtaining High Purity Food-Grade Bioactive Compounds and Value-Added Co-Products from Citrus Wastes. Foods 2019, 8, 523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Mahato, N.; Sharma, K.; Sinha, M.; Cho, M.H. Citrus waste derived nutra-/pharmaceuticals for health benefits: Current trends and future perspectives. J. Funct. Foods 2018, 40, 307–316. [Google Scholar] [CrossRef]
  4. Mahato, N.; Sharma, K.; Nabybaccus, F.; Cho, M.H. Pharma-/Nutraceutical Applications. Eras J. Med. Res. 2016, 3, 20–32. [Google Scholar]
  5. Mahato, N.; Sharma, K.; Rao, R.K.; Sinha, M.; Baral, E.R.; Cho, M.H. Citrus essential oils: Extraction, authentication and application in food preservation. Crit. Rev. Food Sci. Nutr. 2017, 59, 611–625. [Google Scholar] [CrossRef] [PubMed]
  6. Sharma, K.; Mahato, N.; Cho, M.H.; Lee, Y.R. Converting citrus wastes into value-added products: Economic and environ-mently friendly approaches. Nutrition 2017, 34, 29–46. [Google Scholar] [CrossRef]
  7. Sharma, K.; Mahato, N.; Lee, Y.R. Extraction, characterization and biological activity of citrus flavonoids. Rev. Chem. Eng. 2019, 35, 265–284. [Google Scholar] [CrossRef]
  8. Mahato, N.; Sharma, K.; Sinha, M.; Dhyani, A.; Pathak, B.; Jang, H.; Park, S.; Pashikanti, S.; Cho, S. Biotransformation of Citrus Waste-I: Production of Biofuel and Valuable Compounds by Fermentation. Processes 2021, 9, 220. [Google Scholar] [CrossRef]
  9. Crowe, M.; Nolan, K.; Collins, C.; Carty, G.; Donlon, B.; Brøgger, M. European Environment Agency Topic Report: Biodegradable Municipal Waste Management in Europe; EEA: Copenhagen, Denmark, 2002; p. 123. [Google Scholar]
  10. United States Department of Agriculture. Chemistry and Technology of Citrus. In Citrus Products Agriculture Handbook; USDA: Washington, DC, USA, 1956; Volume 98, pp. 6–24. [Google Scholar]
  11. O’Neal, B.F. Citrus Waste Research Project FLA; State Bd Health Bur. Sanit. Engin. Final Rpt; USDA: Washington, DC, USA, 1953; p. 82. [Google Scholar]
  12. Annonymous. Grappling with that problem of citrus waste odors. Food Eng. 1954, 26, 145. [Google Scholar]
  13. Sanborn, N. Spray irrigation as a means of cannery waste disposal. Cann. Trade 1952, 74, 17–21. [Google Scholar]
  14. Ingols, R. The citrus canning disposal problem in Florida. Sew. Work. J. 1945, 17, 320–329. [Google Scholar]
  15. McNary, R.R. Citrus Canning Industry. Ind. Eng. Chem. 1947, 39, 625–627. [Google Scholar] [CrossRef]
  16. Sanborn, N. Nitrate treatment of cannery wastes. Fruit Prod. J. Am. Vineg. Ind. 1941, 20, 207–208. [Google Scholar]
  17. Hooper, C.H. Citrus Fruits. Nature 1915, 96, 58–59. [Google Scholar] [CrossRef]
  18. Wakefield, J. The results of research on citrus waste disposal. Proc. Fla. State Hortic. Soc. 1953, 246–254. [Google Scholar]
  19. Murdock, D.I.; Allen, W.E. Germicidal Effect of Orange Peel Oil and D-Limonene in Water and Orange Juice. Food Technol. 1960, 14, 441–445. [Google Scholar]
  20. Subba, M.S.; Soumithri, T.C.; Rao, R.S. Antimicrobial Action of Citrus Oils. J. Food Sci. 1967, 32, 225–227. [Google Scholar] [CrossRef]
  21. Ahmaruzzaman, M. Industrial wastes as low-cost potential adsorbents for the treatment of wastewater laden with heavy metals. Adv. Colloid Interface Sci. 2011, 166, 36–59. [Google Scholar] [CrossRef]
  22. Alshabanat, M.; Alsenani, G.; Almufarij, R. Removal of crystal violet dye from aqueous solutions onto date palm fiber by ad-sorption technique. J. Chem. 2013, 2013, 1001. [Google Scholar] [CrossRef] [Green Version]
  23. Brusick, D. Genotoxicity of Phenolic Antioxidants. Toxicol. Ind. Health 1993, 9, 223–230. [Google Scholar] [CrossRef]
  24. Dich, J.; Zahm, S.H.; Hanberg, A.; Adami, H.O. Pesticides and cancer. Cancer Causes Control 1997, 8, 420–443. [Google Scholar] [CrossRef] [PubMed]
  25. Hossain, M.; Ngo, H.; Guo, W.; Nguyen, T.V.; Vigneswaran, S. Performance of cabbage and cauliflower wastes for heavy metals removal. Desalin. Water Treat. 2014, 52, 844–860. [Google Scholar] [CrossRef]
  26. Dudgeon, D.; Arthington, A.H.; Gessner, M.O.; Kawabata, Z.-I.; Knowler, D.J.; Lévêque, C.; Naiman, R.J.; Prieur-Richard, A.-H.; Soto, D.; Stiassny, M.L.J.; et al. Freshwater biodiversity: Importance, threats, status and conservation challenges. Biol. Rev. 2006, 81, 163–182. [Google Scholar] [CrossRef] [PubMed]
  27. Lintern, A.; Leahy, P.J.; Heijnis, H.; Zawadzki, A.; Gadd, P.; Jacobsen, G.; Deletic, A.; Mccarthy, D.T. Identifying heavy metal levels in historical flood water deposits using sediment cores. Water Res. 2016, 105, 34–46. [Google Scholar] [CrossRef]
  28. Liu, M.; Du, P.; Yu, C.; He, Y.; Zhang, H.; Sun, X.; Lin, H.; Luo, Y.; Xie, H.; Guo, J.; et al. Increases of Total Mercury and Methylmercury Releases from Municipal Sewage into Environment in China and Implications. Environ. Sci. Technol. 2018, 52, 124–134. [Google Scholar] [CrossRef]
  29. Crini, G. Non-conventional low-cost adsorbents for dye removal: A review. Bioresour. Technol. 2006, 97, 1061–1085. [Google Scholar] [CrossRef]
  30. Daneshvar, N.; Oladegaragoze, A.; Djafarzadeh, N. Decolorization of basic dye solutions by electrocoagulation: An investigation of the effect of operational parameters. J. Hazard. Mater. 2006, 129, 116–122. [Google Scholar] [CrossRef]
  31. Patil, Y.; Paknikar, K. Development of a process for biodetoxification of metal cyanides from waste waters. Process. Biochem. 2000, 35, 1139–1151. [Google Scholar] [CrossRef]
  32. Nesbitt, A.B. Recovery of Metal Cyanides Using a Fluidized Bed of Resin. Master’s Thesis, Cape Peninsula University of Technology, Cape Town, South Africa, December 1996. [Google Scholar]
  33. Santos, B.A.Q.; Ntwampe, S.K.O.; Doughari, J.H.; Muchatibaya, G. Application of Citrus sinensis Solid Waste as a Pseudo-Catalyst for Free Cyanide Conversion under Alkaline Conditions. Bioresources 2013, 8, 3461–3467. [Google Scholar] [CrossRef] [Green Version]
  34. Kostich, M.S.; Batt, A.L.; Lazorchak, J. Concentrations of prioritized pharmaceuticals in effluents from 50 large wastewater treatment plants in the US and implications for risk estimation. Environ. Pollut. 2014, 184, 354–359. [Google Scholar] [CrossRef]
  35. Lessa, E.F.; Nunes, M.L.; Fajardo, A.R. Chitosan/waste coffee-grounds composite: An efficient and eco-friendly adsorbent for removal of pharmaceutical contaminants from water. Carbohydr. Polym. 2018, 189, 257–266. [Google Scholar] [CrossRef]
  36. Larsson, D.G.J. Pollution from drug manufacturing: Review and perspectives. Philos. Trans. R. Soc. B Biol. Sci. 2014, 369, 20130571. [Google Scholar] [CrossRef] [Green Version]
  37. Yan, Z.; Lu, G.; Liu, J.; Jin, S. An integrated assessment of estrogenic contamination and feminization risk in fish in Taihu Lake, China. Ecotoxicol. Environ. Saf. 2012, 84, 334–340. [Google Scholar] [CrossRef]
  38. Adewuyi, A. Chemically modified bio-sorbents and their role in the removal of emerging pharmaceutical waste in the water system. Water 2020, 12, 1551. [Google Scholar] [CrossRef]
  39. Swan, G.; Naidoo, V.; Cuthbert, R.; Green, R.E.; Pain, D.J.; Swarup, D.; Prakash, V.; Taggart, M.; Bekker, L.; Das, D.; et al. Removing the Threat of Diclofenac to Critically Endangered Asian Vultures. PLoS Biol. 2006, 4, e66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Oaks, J.L.; Gilbert, M.; Virani, M.Z.; Watson, R.T.; Meteyer, C.U.; Rideout, B.A. Diclofenac residues as the cause of vulture population decline in Pakistan. Nature 2004, 427, 630–633. [Google Scholar] [CrossRef]
  41. Shultz, S.; Baral, H.S.; Charman, S.; Cunningham, A.A.; Das, D.; Ghalsasi, G.R.; Goudar, M.S.; Green, R.E.; Jones, A.; Nighot, P.; et al. Diclofenac poisoning is widespread in declining vulture populations across the Indian subcontinent. Proc. R. Soc. B Boil. Sci. 2004, 271, S458–S460. [Google Scholar] [CrossRef]
  42. Green, R.E.; Newton, I.; Shultz, S.; Cunningham, A.A.; Pain, D.J.; Prakash, V. Diclofenac Poisoning as a Cause of Vulture Population Declines across the Indian Subcontinent. J. Appl. Ecol. 2004, 41, 793–800. [Google Scholar] [CrossRef]
  43. Prakash, V.; Pain, D.; Cunningham, A.; Donald, P.; Verma, A.; Gargi, R.; Sivakumar, S.; Rahmani, A. Catastrophic collapse of Indian white-backed Gyps bengalensis and long-billed Gyps indicus vulture populations. Biol. Conserv. 2003, 109, 381–390. [Google Scholar] [CrossRef]
  44. Patterson, J. Industrial Wastewater Treatment Technology, 2nd ed.; Patterson, J., Ed.; Butterworth Publishers: Stoneham, MA, USA, 1985. [Google Scholar]
  45. Hameed, B.; Tan, I.A.W.; Ahmad, A.L. Optimization of basic dye removal by oil palm fibre-based activated carbon using response surface methodology. J. Hazard. Mater. 2008, 158, 324–332. [Google Scholar] [CrossRef]
  46. Lata, H.; Garg, V.; Gupta, R. Removal of a basic dye from aqueous solution by adsorption using Parthenium hysterophorus: An agricultural waste. Dye Pigment. 2007, 74, 653–658. [Google Scholar] [CrossRef]
  47. Tan, X.; Liu, Y.; Zeng, G.; Wang, X.; Hu, X.; Gu, Y.; Yang, Z. Application of biochar for the removal of pollutants from aqueous solutions. Chemosphere 2015, 125, 70–85. [Google Scholar] [CrossRef]
  48. Shakoor, S.; Nasar, A. Removal of methylene blue dye from artificially contaminated water using Citrus limetta peel waste as a very low cost adsorbent. J. Taiwan Inst. Chem. Eng. 2016, 66, 154–163. [Google Scholar] [CrossRef]
  49. World Health Organization. Guidelines for Drinking—Water Quality; WHO: Geneva, Switzerland, 2011. [Google Scholar]
  50. Purkayastha, D.; Mishra, U.; Biswas, S. A comprehensive review on Cd(II) removal from aqueous solution. J. Water Process. Eng. 2014, 2, 105–128. [Google Scholar] [CrossRef]
  51. Yadav, S.; Sinha, S.; Singh, D.K. Chromium(VI) removal from aqueous solution and industrial wastewater by modified date palm trunk. Environ. Prog. Sustain. Energy 2014, 34, 452–460. [Google Scholar] [CrossRef]
  52. Bilal, M.; Shah, J.A.; Ashfaq, T.; Mubashar, S.; Gardazi, H.; Tahir, A.A. Waste biomass adsorbents for copper removal from industrial wastewater-A review. J. Hazard. Mater. 2013, 263, 322–333. [Google Scholar] [CrossRef]
  53. Zahra, N. Lead Removal from Water by Low Cost Adsorbents: A Review. Pak. J. Anal. Environ. Chem. 2012, 13, 1–8. [Google Scholar]
  54. Kanawade, S.M.; Gaikwad, R.W. Removal of Zinc Ions from Industrial Effluent by Using Cork Powder as Adsorbent. Int. J. Chem. Eng. Appl. 2011, 2, 199–201. [Google Scholar] [CrossRef] [Green Version]
  55. Singh, T.P.; Majumder, C.B. Flouride removal from sewage water using Citrus limetta peel as bio-sorbent. Int. J. Pharm. Pharm. Sci. 2016, 8, 86–92. [Google Scholar]
  56. Siddique, A.; Nayak, A.K.; Singh, J. Synthesis of FeCl3-activated carbon derived from waste Citrus limetta peels for removal of fluoride: An eco-friendly approach for the treatment of groundwater and bio-waste collectively. Groundw. Sustain. Dev. 2020, 10, 100339. [Google Scholar] [CrossRef]
  57. Dolas, H.; Sahin, O.; Saka, C.; Demir, H. A new method on producing high surface area activated carbon: The effect of salt on the surface area and the pore size distribution of activated carbon prepared from pistachio shell. Chem. Eng. J. 2011, 166, 191–197. [Google Scholar] [CrossRef]
  58. Özdemir, M.; Bolgaz, T.; Saka, C.; Şahin, Ö. Preparation and characterization of activated carbon from cotton stalks in a two-stage process. J. Anal. Appl. Pyrolysis 2011, 92, 171–175. [Google Scholar] [CrossRef]
  59. Saka, C.; Sahin, O.; Celik, M.S. The Removal of Methylene Blue from Aqueous Solutions by Using Microwave Heating and Pre-boiling Treated Onion Skins as a New Adsorbent. Energy Sources Part A Recover. Util. Environ. Eff. 2012, 34, 1577–1590. [Google Scholar] [CrossRef]
  60. Bulut, Y.; Aydın, H. A kinetics and thermodynamics study of methylene blue adsorption on wheat shells. Desalination 2006, 194, 259–267. [Google Scholar] [CrossRef]
  61. Vadivelan, V.; Kumar, K.V. Equilibrium, kinetics, mechanism, and process design for the sorption of methylene blue onto rice husk. J. Colloid Interface Sci. 2005, 286, 90–100. [Google Scholar] [CrossRef]
  62. Uddin, M.T.; Islam, M.A.; Mahmud, S.; Rukanuzzaman, M. Adsorptive removal of methylene blue by tea waste. J. Hazard. Mater. 2009, 164, 53–60. [Google Scholar] [CrossRef]
  63. Annadurai, G.; Juang, R.-S.; Lee, D.-J. Use of cellulose-based wastes for adsorption of dyes from aqueous solutions. J. Hazard. Mater. 2002, 92, 263–274. [Google Scholar] [CrossRef]
  64. Bhattacharya, K.G.; Sharma, A. Kinetics and thermodynamics of Methylene Blue adsorption on Neem (Azadirachta indica) leaf powder. Dye Pigment. 2005, 65, 51–59. [Google Scholar] [CrossRef]
  65. McKay, G.; Porter, J.F.; Prasad, G.R. The Removal of Dye Colours from Aqueous Solutions by Adsorption on Low-cost Materials. Water Air Soil Pollut. 1999, 114, 423–438. [Google Scholar] [CrossRef]
  66. Asgher, M.; Bhatti, H.N. Removal of reactive blue 19 and reactive blue 49 textile dyes by citrus waste biomass from aqueous solution: Equilibrium and kinetic study. Can. J. Chem. Eng. 2011, 90, 412–419. [Google Scholar] [CrossRef]
  67. Bhatti, H.N.; Bajwa, I.I.; Hanif, M.A.; Bukhari, I.H. Removal of lead and cobalt using lignocellulosic fiber derived from Citrus reticulata waste biomass. Korean J. Chem. Eng. 2010, 27, 218–227. [Google Scholar] [CrossRef]
  68. Won, S.W.; Kotte, P.; Wei, W.; Lim, A.; Yun, Y.S. Bio-sorbents for recovery of precious metals. Bioresour. Technol. 2014, 160, 203–212. [Google Scholar] [CrossRef]
  69. Iqbal, M.; Schiewer, S.; Cameron, R. Mechanistic elucidation and evaluation of biosorption of metal ions by grapefruit peel using FTIR spectroscopy, kinetics and isotherms modeling, cations displacement and EDX analysis. J. Chem. Technol. Biotechnol. 2009, 84, 1516–1526. [Google Scholar] [CrossRef]
  70. Suresh Babu, C.; Chakrapani, C.; Somasekhara Rao, K. Equilibrium and kinetic studies of reactive red 2 dye adsorption onto prepared activated carbons. J. Chem. Pharm. Res. 2011, 3, 428–439. [Google Scholar]
  71. Chen, S.; Luo, J.; Hu, M.; Lai, K.; Geng, P.; Huang, H. Enhancement of cypermethrin degradation by a coculture of Bacillus cereus ZH-3 and Streptomyces aureus HP-S-01. Bioresour. Technol. 2012, 110, 97–104. [Google Scholar] [CrossRef]
  72. Dutta, S.; Bhattacharyya, A.; Ganguly, A.; Gupta, S.; Basu, S. Application of Response Surface Methodology for preparation of low-cost adsorbent from citrus fruit peel and for removal of Methylene Blue. Desalination 2011, 275, 26–36. [Google Scholar] [CrossRef]
  73. Mahmoud, G.A.; Abdel-Aal, S.E.; Badway, N.A.; Elbayaa, A.A.; Ahmed, D.F. A novel hydrogel based on agricultural waste for removal of hazardous dyes from aqueous solution and reuse process in a secondary adsorption. Polym. Bull. 2016, 74, 337–358. [Google Scholar] [CrossRef]
  74. Rehman, R.; Zafar, J.; Nisar, H. Adsorption Studies of Removal of Indigo Caramine Dye from Water by Formaldehyde and Urea Treated Cellulosic Waste of Citrus reticulata Peels. Asian J. Chem. 2014, 26, 43–47. [Google Scholar] [CrossRef]
  75. Rehman, R.; Salman, M.; Mahmud, T.; Kanwal, F.; Waheed-Uz-Zama. Utilization of chemically modified Citrus reticulata peels for biosorptive removal of acid yellow-73 dye from water. J. Chem. Soc. Pak. 2013, 35, 611–616. [Google Scholar]
  76. Dev, S.; Khamkhash, A.; Ghosh, T.; Aggarwal, S. Adsorptive Removal of Se(IV) by Citrus Peels: Effect of Adsorbent Entrapment in Calcium Alginate Beads. ACS Omega 2020, 5, 17215–17222. [Google Scholar] [CrossRef] [PubMed]
  77. Kam, S.K.; Lee, M.G. Adsorption characteristics of antibiotics amoxicillin in aqueous solution with activated carbon prepared from waste citrus peel. Appl. Chem. Eng. 2018, 29, 369–375. [Google Scholar]
  78. Putra, E.K.; Pranowo, R.; Sunarso, J.; Indraswati, N.; Ismadji, S. Performance of activated carbon and bentonite for adsorption of amoxicillin from wastewater: Mechanisms, isotherms and kinetics. Water Res. 2009, 43, 2419–2430. [Google Scholar] [CrossRef]
  79. Moussavi, G.; Alahabadi, A.; Yaghmaeian, K.; Eskandari, M. Preparation, characterization and adsorption potential of the NH4Cl-induced activated carbon for the removal of amoxicillin antibiotic from water. Chem. Eng. J. 2013, 217, 119–128. [Google Scholar] [CrossRef]
  80. Ding, R.; Zhang, P.; Seredych, M.; Bandosz, T.J. Removal of antibiotics from water using sewage sludge- and waste oil sludge-derived adsorbents. Water Res. 2012, 46, 4081–4090. [Google Scholar] [CrossRef]
  81. Baccar, R.; Sarra, M.; Bouzid, J.; Feki, M.; Blánquez, P. Removal of pharmaceutical compounds by activated carbon prepared from agricultural by-product. Chem. Eng. J. 2012, 211–212, 310–317. [Google Scholar] [CrossRef]
  82. Ahmed, M.J.; Theydan, S.K. Microporous activated carbon from Siris seed pods by microwave-induced KOH activation for metronidazole adsorption. J. Anal. Appl. Pyrolysis 2013, 99, 101–109. [Google Scholar] [CrossRef]
  83. Pouretedal, H.R.; Sadegh, N. Effective removal of Amoxicillin, Cephalexin, Tetracycline and Penicillin G from aqueous solutions using activated carbon nanoparticles prepared from vine wood. J. Water Process. Eng. 2014, 1, 64–73. [Google Scholar] [CrossRef]
  84. Camel, V. Solid phase extraction of trace elements. Spectrochim. Acta Part B At. Spectrosc. 2003, 58, 1177–1233. [Google Scholar] [CrossRef]
  85. Li, X.; Tang, Y.; Cao, X.; Lu, D.; Luo, F.; Shao, W. Preparation and evaluation of orange peel cellulose adsorbents for effective removal of cadmium, zinc, cobalt and nickel. Colloids Surf. A Physicochem. Eng. Asp. 2008, 317, 512–521. [Google Scholar] [CrossRef]
  86. Sha, L.; Xueyi, G.; Ningchuan, F.; Qinghua, T. Adsorption of Cu2+ and Cd2+ from aqueous solution by mercapto-acetic acid modified orange peel. Colloids Surf. B Biointerfaces 2009, 73, 10–14. [Google Scholar] [CrossRef] [PubMed]
  87. Kohn, R. Binding of divalent cations to oligomeric fragments of pectin. Carbohydr. Res. 1987, 160, 343–353. [Google Scholar] [CrossRef]
  88. Malovíková, A.; Kohn, R. 702 Binding of Cadmium Cations to Pectin. Collect. Czech. Chem. Commun. 1982, 47, 702–708. [Google Scholar] [CrossRef]
  89. Cameron, R.G.; Iqbal, M.; Widmer, W.; Haven, A.S.N.W.W. Biosorption Properties of Citrus Peel Derived Oligogalac-turonides, Enzyme-modified Pectin and Peel Hydrolysis Residues. Proc. Fla. State Hortic. Soc. 2008, 5, 311–314. [Google Scholar]
  90. Haven, W. A Refereed Paper Production of Narrow-Range Size-Classes of Polygalacturonic Acid Oligomers. Proc. Fla. State Hortic. Soc. 2005, 118, 406–409. [Google Scholar]
  91. Joye, D.; Luzio, G. Process for selective extraction of pectins from plant material by differential pH. Carbohydr. Polym. 2000, 43, 337–342. [Google Scholar] [CrossRef]
  92. Arief, V.O.; Trilestari, K.; Sunarso, J.; Indraswati, N.; Ismadji, S. Recent progress on biosorption of heavy metals from liquids using low cost bio-sorbents: Characterization, biosorption parameters and mechanism studies. Clean Soil Air Water 2008, 36, 937–962. [Google Scholar] [CrossRef]
  93. Vijayaraghavan, K.; Yun, Y.S. Bacterial bio-sorbents and biosorption. Biotechnol. Adv. 2008, 26, 266–291. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, J.; Chen, C. Biosorption of heavy metals by Saccharomyces cerevisiae: A review. Biotechnol. Adv. 2006, 24, 427–451. [Google Scholar] [CrossRef] [PubMed]
  95. Mahvi, A.H. Application of agricultural fibers in pollution removal from aqueous solution. Int. J. Environ. Sci. Technol. 2008, 5, 275–285. [Google Scholar] [CrossRef] [Green Version]
  96. Suryavanshi, U.; Shukla, S.R. Adsorption of Pb2+ by Alkali-Treated Citrus limetta Peels. Ind. Eng. Chem. Res. 2010, 49, 11682–11688. [Google Scholar] [CrossRef]
  97. Chatterjee, A.; Schiewer, S. Biosorption of Cadmium(II) Ions by Citrus Peels in a Packed Bed Column: Effect of Process Pa-rameters and Comparison of Different Breakthrough Curve Models. Clean Soil Air Water 2011, 39, 874–881. [Google Scholar] [CrossRef]
  98. Chatterjee, A.; Schiewer, S. Effect of Competing Cations (Pb, Cd, Zn, and Ca) in Fixed-Bed Column Biosorption and Desorption from Citrus Peels. Water Air Soil Pollut. 2014, 225, 1854. [Google Scholar] [CrossRef]
  99. Ghimire, K.N.; Inoue, K.; Yamaguchi, H.; Makino, K.; Miyajima, T. Adsorptive separation of arsenate and arsenite anions from aqueous medium by using orange waste. Water Res. 2003, 37, 4945–4953. [Google Scholar] [CrossRef]
  100. Gondhalekar, S.C.; Shukla, S.R. Equilibrium and kinetics study of uranium(VI) from aqueous solution by Citrus limetta peels. J. Radioanal. Nucl. Chem. 2014, 302, 451–457. [Google Scholar] [CrossRef]
  101. Khaskheli, M.I.; Memon, S.Q.; Parveen, S.; Khuhawar, M.Y. Citrus paradisi: An Effective bio-adsorbent for Arsenic(V) Remediation. Pak. J. Anal. Environ. Chem. 2014, 15, 35–41. [Google Scholar]
  102. Khaskheli, M.I.; Memon, S.Q.; Siyal, A.N.; Khuhawar, M.Y. Use of Orange Peel Waste for Arsenic Remediation of Drinking Water. Waste Biomass Valoriz. 2011, 2, 423–433. [Google Scholar] [CrossRef]
  103. Gönen, F. Adsorption study on orange peel: Removal of Ni(II) ions from aqueous solution. Afr. J. Biotechnol. 2012, 11, 1250–1258. [Google Scholar] [CrossRef]
  104. Marín, A.B.P.; Aguilar, M.I.; Ortuño, J.F.; Meseguer, V.F.; Sáez, J.; Llorens, M. Biosorption of Zn(II) by orange waste in batch and packed-bed systems. J. Chem. Technol. Biotechnol. 2010, 85, 1310–1318. [Google Scholar] [CrossRef]
  105. Pavan, F.A.; Lima, I.S.; Lima, É.C.; Airoldi, C.; Gushikem, Y. Use of Ponkan mandarin peels as bio-sorbent for toxic metals uptake from aqueous solutions. J. Hazard. Mater. 2006, 137, 527–533. [Google Scholar] [CrossRef]
  106. Abdul Sattar, J.A. Toxic Metal Pollution Abatement Using Sour Orange Biomass. J. Al-Nahrain Univ. Sci. 2013, 16, 56–64. [Google Scholar] [CrossRef]
  107. Torab-Mostaedi, M. Biosorpcija lantana i cerijuma iz vodenih rastvora pomoću kore mandarine (Citrus reticulata): Ravnotežna, kinetička i termodinamička ispitivanja. Chem. Ind. Chem. Eng. Q. 2013, 19, 79–88. [Google Scholar] [CrossRef]
  108. Pavan, F.A.; Mazzocato, A.C.; Jacques, R.A.; Dias, S.L.P. Ponkan peel: A potential bio-sorbent for removal of Pb(II) ions from aqueous solution. Biochem. Eng. J. 2008, 40, 357–362. [Google Scholar] [CrossRef]
  109. Pérez-Marín, A.; Ballester, A.; González, F.; Blázquez, M.; Muñoz, J.; Sáez, J.; Zapata, V.M. Study of cadmium, zinc and lead biosorption by orange wastes using the subsequent addition method. Bioresour. Technol. 2008, 99, 8101–8106. [Google Scholar] [CrossRef]
  110. Rane, N.M.; Sapkal, R.S. Chromium(Vi) Removal By Using Orange Peel Powder in Batch Adsorption. Int. J. Chem. Sci. Appl. 2014, 5, 2278–6015. [Google Scholar]
  111. Santos, C.M.; Dweck, J.; Viotto, R.S.; Rosa, A.H.; de Morais, L.C. Application of orange peel waste in the production of solid biofuels and bio-sorbents. Bioresour. Technol. 2015, 196, 469–479. [Google Scholar] [CrossRef]
  112. Njikam, E.; Schiewer, S. Optimization and kinetic modeling of cadmium desorption from citrus peels: A process for bio-sorbent regeneration. J. Hazard. Mater. 2012, 213–214, 242–248. [Google Scholar] [CrossRef] [PubMed]
  113. Sowmya Lakshmi, K.B.; Munilakshmi, N. Adsorptive removal of colour from aqueous solution of disazo dye by using organic adsorbents. Int. J. Chem. Technol. Res. 2016, 9, 407–415. [Google Scholar]
  114. Lasheen, M.R.; Ammar, N.S.; Ibrahim, H.S. Adsorption/desorption of Cd(II), Cu(II) and Pb(II) using chemically modified orange peel: Equilibrium and kinetic studies. Solid State Sci. 2012, 14, 202–210. [Google Scholar] [CrossRef]
  115. Balaria, A.; Schiewer, S. Assessment of biosorption mechanism for Pb binding by citrus pectin. Sep. Purif. Technol. 2008, 63, 577–581. [Google Scholar] [CrossRef]
  116. Biswas, B.K.; Inoue, K.; Ghimire, K.N.; Ohta, S.; Harada, H.; Ohto, K.; Kawakita, H. The adsorption of phosphate from an aquatic environment using metal-loaded orange waste. J. Colloid Interface Sci. 2007, 312, 214–223. [Google Scholar] [CrossRef]
  117. Feng, N.-C.; Guo, X. Characterization of adsorptive capacity and mechanisms on adsorption of copper, lead and zinc by modified orange peel. Trans. Nonferrous Met. Soc. China 2012, 22, 1224–1231. [Google Scholar] [CrossRef]
  118. Mandina, S. Removal of chromium (VI) from aqueous solution using chemically modified orange (citrus cinensis) peel. IOSR J. Appl. Chem. 2013, 6, 66–75. [Google Scholar] [CrossRef]
  119. Bhatnagar, A.; Minocha, A.K.; Sillanpää, M. Adsorptive removal of cobalt from aqueous solution by utilizing lemon peel as bio-sorbent. Biochem. Eng. J. 2010, 48, 181–186. [Google Scholar] [CrossRef]
  120. De Souza, J.V.T.M.; Diniz, K.M.; Massocatto, C.L.; Tarley, C.R.T.; Caetano, J.; Dragunski, D.C. Removal of Pb (II) from aqueous solution with orange sub-products chemically modified as biosorbent. BioResources 2012, 7, 2300–2318. [Google Scholar] [CrossRef] [Green Version]
  121. Tasaso, P. Adsorption of Copper Using Pomelo Peel and Depectinated Pomelo Peel. J. Clean Energy Technol. 2014, 2, 154–157. [Google Scholar] [CrossRef] [Green Version]
  122. Schiewer, S.; Iqbal, M. The role of pectin in Cd binding by orange peel bio-sorbents: A comparison of peels, depectinated peels and pectic acid. J. Hazard. Mater. 2010, 177, 899–907. [Google Scholar] [CrossRef] [PubMed]
  123. Schiewer, S.; Patil, S.B. Modeling the effect of pH on biosorption of heavy metals by citrus peels. J. Hazard. Mater. 2008, 157, 8–17. [Google Scholar] [CrossRef]
  124. Singh, S.; Shukla, S.R. Adsorptive removal of cobalt ions on raw and alkali-treated lemon peels. Int. J. Environ. Sci. Technol. 2015, 13, 165–178. [Google Scholar] [CrossRef] [Green Version]
  125. Crockett, M.; Evans, A.; Worthington, M.J.H.; Albuquerque, I.S.; Slattery, A.D.; Gibson, C.; Campbell, J.; Lewis, D.A.; Bernardes, G.; Chalker, J.M. Sulfur-Limonene Polysulfide: A Material Synthesized Entirely from Industrial By-Products and Its Use in Removing Toxic Metals from Water and Soil. Angew. Chem. Int. Ed. 2016, 55, 1714–1718. [Google Scholar] [CrossRef] [PubMed]
  126. Namasivayam, C.; Muniasamy, N.; Gayatri, K.; Rani, M.; Ranganathan, K. Removal of dyes from aqueous solutions by cellulosic waste orange peel. Bioresour. Technol. 1996, 57, 37–43. [Google Scholar] [CrossRef]
  127. Khaled, A.; El Nemr, A.; El-Sikaily, A.; Abdelwahab, O. Treatment of artificial textile dye effluent containing Direct Yellow 12 by orange peel carbon. Desalination 2009, 238, 210–232. [Google Scholar] [CrossRef]
  128. Saeed, A.; Sharif, M.; Iqbal, M. Application potential of grapefruit peel as dye sorbent: Kinetics, equilibrium and mechanism of crystal violet adsorption. J. Hazard. Mater. 2010, 179, 564–572. [Google Scholar] [CrossRef]
  129. Foo, K.Y.; Hameed, B.H. Microwave assisted preparation of activated carbon from pomelo skin for the removal of anionic and cationic dyes. Chem. Eng. J. 2011, 173, 385–390. [Google Scholar] [CrossRef]
  130. Jayarajan, M.; Arunachala, R.; Annadurai, G. Use of Low Cost Nano-porous Materials of Pomelo Fruit Peel Wastes in Removal of Textile Dye. Res. J. Environ. Sci. 2011, 5, 434–443. [Google Scholar] [CrossRef]
  131. Bhatnagar, A.; Kumar, E.; Minocha, A.K.; Jeon, B.-H.; Song, H.; Seo, Y.-C. Removal of Anionic Dyes from Water usingCitrus limonum(Lemon) Peel: Equilibrium Studies and Kinetic Modeling. Sep. Sci. Technol. 2009, 44, 316–334. [Google Scholar] [CrossRef]
  132. Bhatti, H.N.; Akhtar, N.; Saleem, N. Adsorptive Removal of Methylene Blue by Low-Cost Citrus sinensis Bagasse: Equilibrium, Kinetic and Thermodynamic Characterization. Arab. J. Sci. Eng. 2011, 37, 9–18. [Google Scholar] [CrossRef]
  133. Kučić, D.; Miljanić, S.; Rožić, M. Sorption of Methylene Blue onto Orange and Lemon Peel. Holist. Approach Environ. 2011, 1, 39–49. [Google Scholar]
  134. Rosales, E.; Meijide, J.; Tavares, T.; Pazos, M.; Sanromán, M.A. Grapefruit peelings as a promising bio-sorbent for the removal of leather dyes and hexavalent chromium. Process. Saf. Environ. Prot. 2016, 101, 61–71. [Google Scholar] [CrossRef] [Green Version]
  135. Hameed, B.H.; Mahmoud, D.; Ahmad, A.L. Sorption of basic dye from aqueous solution by pomelo (Citrus grandis) peel in a batch system. Colloids Surf. A Physicochem. Eng. Asp. 2008, 316, 78–84. [Google Scholar] [CrossRef]
  136. Patil, P.R. Adsorption of Erichrome black T from aqueous solutions on activated carbon prepared from mosambi peel. J. Appl. Sci. Environ. Sanit. 2014, 6, 149–154. [Google Scholar]
  137. Mafra, M.R.; Igarashimafra, L.; Zuim, D.R.; Vasques, É.C.; Ferreira, M.A. Adsorption of remazol brilliant blue on an orange peel adsorbent. Braz. J. Chem. Eng. 2013, 30, 657–665. [Google Scholar] [CrossRef]
  138. Pavlović, M.D.; Nikolić, I.R.; Milutinović, M.D.; Dimitrijević-Branković, S.I.; Šiler-Marinković, S.S.; Antonović, D.G. Iskorišćenje sirovog otpada iz restorana za adsorpciju boja. Hem. Ind. 2015, 69, 667–677. [Google Scholar]
  139. Rahmat, N.A.; Ali, A.A.; Salmiati; Hussain, N.; Muhamad, M.S.; Kristanti, R.; Hadibarata, T. Removal of Remazol Brilliant Blue R from Aqueous Solution by Adsorption Using Pineapple Leaf Powder and Lime Peel Powder. Water Air Soil Pollut. 2016, 227, 1–11. [Google Scholar] [CrossRef]
  140. Bello, O.S.; Ahmad, M.A.; Semire, B. Scavenging malachite green dye from aqueous solutions using pomelo (Citrus grandis) peels: Kinetic, equilibrium and thermodynamic studies. Desalin. Water Treat. 2015, 56, 521–535. [Google Scholar] [CrossRef]
  141. Sivaraj, R.; Namasivayam, C.; Kadirvelu, K. Orange peel as an adsorbent in the removal of Acid violet 17 (acid dye) from aqueous solutions. Waste Manag. 2001, 21, 105–110. [Google Scholar] [CrossRef]
  142. Salman, T.A.; Ali, M.I. Potential Application of Natural and Modified Orange Peel as an Eco‒friendly Adsorbent for Methylene Blue Dye. Iraqi J. Sci. 2016, 57, 1–13. [Google Scholar]
  143. Lagergren, S. Zur Theorie der Sogenannten Adsorption Gelöster Stoffe. Sven. Vetensk. Handl. 1998, 24, 1–39. [Google Scholar]
  144. Ho, Y.S.; McKay, G. Pseudo-second order model for sorption processes. Process. Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  145. Shakoor, S.; Nasar, A. Adsorptive treatment of hazardous methylene blue dye from artificially contaminated water using cucumis sativus peel waste as a low-cost adsorbent. Groundw. Sustain. Dev. 2017, 5, 152–159. [Google Scholar] [CrossRef]
  146. Levankumar, L.; Muthukumaran, V.; Gobinath, M. Batch adsorption and kinetics of chromium (VI) removal from aqueous solutions by Ocimum americanum L. seed pods. J. Hazard. Mater. 2009, 161, 709–713. [Google Scholar] [CrossRef]
  147. Elovich, S.Y.; Larinov, O.G. Theory of Adsorption from Solutions of Non Electrolytes on Solid (I) Equation Adsorption from Solutions and the Analysis of Its Simplest Form, (II) Verification of the Equation of Adsorption Isotherm from Solutions. Izv. Akad. Nauk. SSSR 1962, 2, 209–216. [Google Scholar]
  148. Weber, W.J.; Morris, J.C. Advances in water pollution research: Removal of biologically resistant pollutant from waste water by adsorption. In Proceedings of the 1st International Conference on Water Pollution Symposium; Pregamon Press: Oxford, UK, 1962; Volume 2, pp. 231–266. [Google Scholar]
  149. Tan, I.A.W.; Hameed, B.H.; Ahmad, A.L. Equilibrium and kinetic studies on basic dye adsorption by oil palm fibre activated carbon. Chem. Eng. J. 2007, 127, 111–119. [Google Scholar] [CrossRef]
  150. Freundlich, H.M.F. Over the adsorption in solution. J. Phys. Chem. 1906, 57, 385–471. [Google Scholar]
  151. Wijaya, R.; Andersan, G.; Santoso, S.P.; Irawaty, W. Green Reduction of Graphene Oxide using Kaffir Lime Peel Extract (Citrus hystrix) and Its Application as Adsorbent for Methylene Blue. Sci. Rep. 2020, 10, 1–9. [Google Scholar] [CrossRef]
  152. Mondal, N.K.; Basu, S.; Sen, K.; Debnath, P. Potentiality of mosambi (Citrus limetta) peel dust toward removal of Cr(VI) from aqueous solution: An optimization study. Appl. Water Sci. 2019, 9, 116. [Google Scholar] [CrossRef] [Green Version]
  153. Ahmed, M.; Mashkoor, F.; Nasar, A. Development, characterization, and utilization of magnetized orange peel waste as a novel adsorbent for the confiscation of crystal violet dye from aqueous solution. Groundw. Sustain. Dev. 2020, 10, 100322. [Google Scholar] [CrossRef]
  154. Sharifzade, G.; Asghari, A.; Rajabi, M. Highly effective adsorption of xanthene dyes (rhodamine B and erythrosine B) from aqueous solutions onto lemon citrus peel active carbon: Characterization, resolving analysis, optimization and mechanistic studies. RSC Adv. 2017, 7, 5362–5371. [Google Scholar] [CrossRef] [Green Version]
  155. Jawad, A.H.; Al-Heetimi, D.T.A.; Mastuli, M.S. Biochar from orange (Citrus sinensis) peels by acid activation for methylene blue adsorption. Iran J. Chem. Chem. Eng. 2019, 38, 91–105. [Google Scholar]
  156. López-Manchado, M.; Arroyo, M. Effect of the incorporation of pet fibers on the properties of thermoplastic elastomer based on PP/elastomer blends. Polymer 2001, 42, 6557–6563. [Google Scholar] [CrossRef]
  157. Ziabari, M.; Mottaghitalab, V.; Haghi, A.K. A new approach for optimization of electrospun nanofiber formation process. Korean J. Chem. Eng. 2010, 27, 340–354. [Google Scholar] [CrossRef]
  158. Agarwal, P.; Mishra, P.K.; Srivastava, P. Statistical optimization of the electrospinning process for chitosan/polylactide nanofabrication using response surface methodology. J. Mater. Sci. 2012, 47, 4262–4269. [Google Scholar] [CrossRef]
  159. Agarwal, P.; Mondal, A.; Mishra, P.; Srivastava, P. A predictive model for the synthesis of polylactide from lactic acid by response surface methodology. ePolymers 2011, 11, 739. [Google Scholar] [CrossRef]
  160. Sudamalla, P.; Pichiah, S.; Manickam, M. Responses of surface modeling and optimization of Brilliant Green adsorption by adsorbent prepared from Citrus limetta peel. Desalin. Water Treat. 2012, 50, 367–375. [Google Scholar] [CrossRef]
  161. Kam, S.K.; Lee, M.G. Response surface modeling for the adsorption of dye eosin Y by activated carbon prepared from waste citrus peel. Appl. Chem. Eng. 2018, 29, 270–277. [Google Scholar]
  162. Lee, M.G.; Kam, S.K. Study on the adsorption of antibiotics trimethoprim in aqueous solution by activated carbon prepared from waste citrus peel using box-behnken design. Korean Chem. Eng Res. 2018, 56, 568–576. [Google Scholar]
Figure 1. Conventional methods of waste removal from citrus processing plants and deteriorating effects to the environment.
Figure 1. Conventional methods of waste removal from citrus processing plants and deteriorating effects to the environment.
Processes 09 01544 g001
Figure 2. Methods of systematic and environment-friendly disposal of citrus wastes, conversion into useful materials and their relative merits.
Figure 2. Methods of systematic and environment-friendly disposal of citrus wastes, conversion into useful materials and their relative merits.
Processes 09 01544 g002
Figure 3. Structures of the anionic and cationic dyes.
Figure 3. Structures of the anionic and cationic dyes.
Processes 09 01544 g003
Figure 4. Hazardous effects of dyes and heavy metals released in the effluents from chemical and textile industries on humans and the environment, and the possible use of citrus waste-derived bio-sorbents for the removal of harmful chemicals from the wastewaters. Artwork developed from the information provided in [47].
Figure 4. Hazardous effects of dyes and heavy metals released in the effluents from chemical and textile industries on humans and the environment, and the possible use of citrus waste-derived bio-sorbents for the removal of harmful chemicals from the wastewaters. Artwork developed from the information provided in [47].
Processes 09 01544 g004
Figure 5. Basic principle of surface modification of citrus waste-derived bio-sorbent material by chemical pre-treatment for the removal and recovery of heavy/precious metals. Artwork created from the information provided in [68].
Figure 5. Basic principle of surface modification of citrus waste-derived bio-sorbent material by chemical pre-treatment for the removal and recovery of heavy/precious metals. Artwork created from the information provided in [68].
Processes 09 01544 g005
Figure 6. (at) Different methods of pre-treatment of the precursor material for the preparation of bio-sorbents from citrus wastes [69,70,71,72,73,74,75,76,84,85,86,87,88].
Figure 6. (at) Different methods of pre-treatment of the precursor material for the preparation of bio-sorbents from citrus wastes [69,70,71,72,73,74,75,76,84,85,86,87,88].
Processes 09 01544 g006aProcesses 09 01544 g006b
Figure 7. (a) Preparation of bio-sorbent from biochemical treatment of citrus peel waste. (b) Bio-sorption capacities of different bio-sorbent substrates prepared in (a). Adapted from the information provided in References [89,90,91].
Figure 7. (a) Preparation of bio-sorbent from biochemical treatment of citrus peel waste. (b) Bio-sorption capacities of different bio-sorbent substrates prepared in (a). Adapted from the information provided in References [89,90,91].
Processes 09 01544 g007
Figure 8. Schematic representation of experimental set-up for fixed-bed adsorption column. Artwork developed from References [97,98,99].
Figure 8. Schematic representation of experimental set-up for fixed-bed adsorption column. Artwork developed from References [97,98,99].
Processes 09 01544 g008
Figure 9. Mechanism of adsorption of dyes and heavy metals from industrial wastewaters by citrus biomass-derived bio-sorbent. Artwork developed from the information provided in [47].
Figure 9. Mechanism of adsorption of dyes and heavy metals from industrial wastewaters by citrus biomass-derived bio-sorbent. Artwork developed from the information provided in [47].
Processes 09 01544 g009
Table 1. The permissible limits and the hazardous effects of heavy metals, dyes and other contaminants on human health.
Table 1. The permissible limits and the hazardous effects of heavy metals, dyes and other contaminants on human health.
Heavy MetalsAllowed Limits (WHO/EPA); mgL−1Source of ContaminationAdverse Effects on HealthRef.
Arsenic0.01Erosion of natural deposits, runoff from soil in orchards, glass and electronics manufacturing waste, tanneriesSkin damage, deformation of digits, cancer, deterioration of circulatory system[48]
Beryllium0.004Metal refineries, coal burning, discharge from electrical, aerospace and metal finishing factories/industriesGastrointestinal disorders[49]
Cadmium0.005Corrosion of pipes and infrastructures, erosion of metal in refineries, battery waste, paints and dyesKidney damage, cancer of lungs, malfunctioning of vital organs, proteinuria[50]
Chromium0.05–0.25Steel industries, metal finishing factories, pulp mills and corrosion of stainless steel pipes and infrastructures, tanneriesAllergy, dermatitis, hemolysis, kidney failure, carcinomas, mutagenic diseases[51]
Copper1.0–1.3Corrosion of household utensils and artefacts, plumbing system, erosion from copper minesGastrointestinal disorders, abdominal irritation, liver and kidney damage[52]
Lead0.005–0.015Battery waste, corrosion of pipes and plumbing system, solder joints, erosion from natural depositsRetardation of growth in children, abnormality in mental health and physical growth, anemia, vomiting, kidney and liver damage, high blood pressure[49,53]
Mercury0.002Erosion from natural deposits, discharge from refineries and factories, runoff from landfills, croplandsHypersensitivity, fever, vomiting, neurological disorders[49]
Selenium0.01–0.05Petroleum refineries, erosion from natural deposits and mining sitesLoss of hair and finger nails, red skin, numbness in fingers and toes, burns, circulatory system disorders[49]
Zinc5Electroplating industry, galvanization of metals, motor oil, battery waste, hydraulic fluid, tire dustCorrosive to skin and eyes, zinc pox, sweet taste, throat dryness, cough, weakness, generalized aching, chills, fever, nausea, vomiting[54]
DyesLess than 1.0 ppmEffluents from fiber and textile industries, paper and pulp industries, plastic industries, paint industriesNausea, vomiting, diarrhea
Fluoride1.5Fluoride deposits (rocks such as topaz, cryolite and fluorapatite, etc.) on earth. Electroplating, glass, ceramics, steel manufacturing and phosphate fertilizer production; semiconductor manufacturing factories, pharmaceutical companies, beryllium extraction plants, aluminum smelters, fertilizer manufacturing and mining industriesDental and skeletal fluorosis, crippling[55,56]
PACs
(Pharmaceutically active compounds)
50 ngL−1 to 0.1 μgL−1Pharmaceutical manufacturing plants, urban sewage, domestic hospital wastewaters, intensive livestock farming, liquid livestock manure production, sewage sludge from agricultural activities and effluents from sewage treatment plantsGenotoxicity and carcinogenicity, multiple drug resistance in pathogenic microorganisms, femaleness in males, tumor and vital organ failure in predatory birds[49]
Table 2. Citrus waste-based bio-sorbents for removal of heavy metals from wastewater (BAT: Batch Adsorption Test; DAT: Direct Adsorption Test (Column Adsorption)).
Table 2. Citrus waste-based bio-sorbents for removal of heavy metals from wastewater (BAT: Batch Adsorption Test; DAT: Direct Adsorption Test (Column Adsorption)).
AdsorbentHeavy MetalsProcessing MethodMaximum Adsorption Capacity
(Per Gram Adsorbent)
Ref.
Citrus limetta peelsUraniumWash; Dry—333.15 K, 24 h; Grind, P.S. = 850 μm–1.2 mm;
BAT—0.1 g BS in 40 mL MS; [U(VI)] = 250 mg/L; pH 2.0–7.0; CT—3 h; SRs—150 rpm
Maximum adsorption of 75.33 mg/g of bio-sorbent material at pH 4.0. Adsorbed species detected are UO22+, UO2OH+, (UO2)3(OH)5+, (UO2)2(OH)22+[100]
Citrus paradisi
(Grapefruit) peels
ArsenicWash; Dry—333.15 K, 6 h; Grind, P.S.=100–125 µm; Wash; Dry—333 K, 24 h
BAT—0.2 g BS in 15–20 mL MS; [As(V)] = 0.1–50 mg/L; pH 4.0–7.0; CT—15 min–2 h; SRs—100–200 rpm
Maximum adsorption of 37.76 mg/g of bio-sorbent material at pH 4.0, 318 K. 76–94% removal in different polluted water sources[101,102]
Orange peelsNickelWash; Dry—373 K, 12 h; Grind, P.S. = 1.80 mm
BAT—0.2 g BS in 100 mL MS; [Ni2+] = 10–200 mg/L;
pH 5.0; SRs—200 rpm
1.05–29.04 mg/g adsorbent from metal solution of concentration 10–200 mg/L. Maximum adsorbing capacity of 33.14%[103]
Orange peelsZincWash; Dry—333.15 K, CW; Grind, P.S. = 0.15–1.5 mm
BAT—0.2 g in 50 mL MS, [Zn2+] = 100 mg/L; pH 4.0–6.0; CT—3 h, stirring
DAT—Acrylic column, 50 cm length; 2.2 cm diameter; 24 g peel pack, [Zn2+] = 20, 30, 40 mg/L, pH 4.0; FL—8.5 mL/min
BAT—0.664 mmol/g (75%) at pH 6.0
DAT—0.42–0.44 mmol/g
[104]
Citrus reticulata (Ponkan mandarin) peelsNickel, Cobalt, CopperWash; Dry—333.15 K, 24 h; Grind, P.S. ≤ 0.6 mm
BAT—0.1 g BS in 25 mL MS, [M2+] = 0.01 M, pH = 4.8, Stirring—2 h
DAT—50 cm length, 0.5 cm diameter, 1.0 g BS, [M2+] = 5 × 10−4 M, FS—3.5 mL/min, pH 4.8
BAT—Nickel—1.92 mmol/g; Cobalt—1.37 mmol/g; Copper—1.31 mmol/g
DAT—Nickel—1.85 mmol/g; Cobalt—1.35 mmol/g; Copper—1.30 mmol/g
[105]
Citrus aurantium (Bitter Orange) fruit partsCobaltWash; Sundry; Grind, P.S.= 250 µm
BAT—2 g BS in 50 mL MS; [Co3+] = 5 mg/L; pH = 2.0, CT = 90 min; stirring
Flavedo = 57.99%; Albedo = 20.11%; Juice = 15.63%; Segment membrane = 20.90%; Seeds = 10.06%[106]
Citrus reticulata TangerineLanhanum
Cerium
Wash; Dry 75–353 K, 24 h; Grind, P.S.= 355 µm
BAT—0.5–3 g/l BS; [M3+] = 10–200 mg/L; pH 2.0–6.0; CT = 5–150 min; SRs—200 rpm; T 20–323 K
La(III)—154.86 mg/g
Ce(III)—162.79 mg/g
[107]
Ponkan mandarin peelsLeadWash; Sun dry—7 d; Grind, P.S. ≤ 600 µm
BAT—0.2 g BS in 25 mL MS; [M2+] = 0.5–1 g/L; pH 5.0;
CT = 120 min; SRs—120 rpm; T 298 K
112.1 mg/g[108]
Orange wasteBinary systems
Cd2+--Zn2+
Cd2+--Pb2+,
Pb2+--Zn2+
Wash; Dry; Grind, P.S.= 0.6–1.5 mm;
BAT—0.4 g BS in 100 mL MS; [M12+–M22+] = 15–100 mg/L —15–25 mg/L added subsequently in 30 combinations.
Pb2+ > Zn2+ > Cd2+
Maximum uptake of 0.25 mmol/g adsorbent
[109]
Orange PeelsChromiumWash; Dry for 3 weeks; Grind
BAT—2 g BS in 250 mL MS; [Cr(VI)] = 0.001 M; CT—5–360 min; Stir—180 rpm
Removal of up to 98% from synthetic chromium solution[110]
Table 3. Citrus waste-derived bio-sorbents via heat/chemical and enzyme pre-treatment for removal of heavy metals from wastewater.
Table 3. Citrus waste-derived bio-sorbents via heat/chemical and enzyme pre-treatment for removal of heavy metals from wastewater.
Bio-Sorbent Modification by Heat/Enzyme/Chemical Treatment (BPT- Bio-Sorbent Pre-Treatment)
AdsorbentHeavy MetalsProcessing MethodMaximum Adsorption Capacity (per Gram Adsorbent)Ref.
Citrus limetta peelsLeadWash; Dry—343 K, 24 h; Grind, P.S. ≤ 1 mm
BPT—10% isopropanol, 303 K, 24 h; 1 M citric acid, 303 K, 1 h, 353 K, 1 h; 0.1 M NaOH, 303 K, 6 h, 353 K, 0.5 h; oxidation with 50% H2O2, 353 K, 2 h at pH 11
BAT—0.1 g PTBS in 50 mL MS; [M2+] = 100 mg/L; pH = 4.0, T = 303 K; RSs—100 rpm; CT—6 h
630 mg/g adsorbent. Cold alkali treatment increases adsorption by 87% (80% in first 15 min)[96]
Citrus tamurana
Citrus latifolia peels
Nickel,
Cadmium, Lead from Oriza sativa (rice)
Wash; Sundry; Grind, P.S. = 250 µm;
BPT—(a) Soaking in 1% w/v citric acid for 10 min, drained, dried at 423 K, 24 h CTBS (citric acid-treated BS)
(b) CTBS heated to 673 K, powdered BSAC (BS active carbon)
(c) BSAC treated with 1% w/v phosphoric acid, dried and sieved ACPA (active carbon treated with phosphoric acid)
BAT—0.1 g CTBS, BSAC and ACPA added to 5 g of raw and rinsed rice, soaked in 250 mL DI with 2% NaCl at pH 6.3, 298 K, 1 h
Rice soaked with ACPA showed maximum reduction in heavy metal concentration
Cd is reduced by 96.4%, Ni by 67.9%, Pb by 90.11%
[111]
Orange, Grapefruit peelsCadmiumWash; Dry; Grind, P.S. = 1–1.1 mm;
BPT—Protonation —20 g BS in 1 L of 0.1 M HNO3, 240 min stirring, rinsed with DI, dried at 313 K for 740 min, protonated BS
BAT—0.05 g protonated BS in 50 mL MS; [Cd2+] = 10–1000 mg/L (0.089–0.89 mM); CT—180 min; pH 5.0
Adsorption of >90% Cd in 50 min. Desorption of the bio-sorbent material using 0.1 M HNO3 + 0.1 M Ca(NO3)2 shows 90% recovery in 60 min[112]
Orange peelsCadmiumWash; Dry; Grind, P.S. = 1–1.1 mm
BPT—Protonation—10 g BS in 500 mL 0.1 N HNO3, stirred for 4 h at 120 rpm, 298 K, rinsed with DI till pH 4.0, dried at 318 K for 12 h; PS = 1–2 mm
DAT—Acrylic bed column; length 30 cm, diameter 1.3 cm; packed with 5.0 g Protonated BS; Bed height = 24–75 cm; [Cd2+] = 5–15 mg/L; FR = 2–15.5 mL/min; pH 5.5; T = 298 K
0.40 mmol/g adsorbent[97]
Orange juice residue wasteArsenate, ArseniteWash; Vacuum dry; Grind, P.S. = 208 µm
BPT—Step I—Decolorization: 10 g BS in 500 mL in 80% EtOH, stirred for 24 h, 298 K, filtered, washed with EtOH until colorless
Vacuum-dried for 24 h, DBSG (de-colorized bio-sorbent gel)
Step II—Cross-linking: 5.22 g DBSG, stirred with 200 mL DMSO for 24 h,
298 K; added 40 mL epichlorhydrin, stirred for 2 h; added
50 mL of 5 M NaOH, stood for 5 h at 323 K; cooled, filtered and washed with 70% EtOH, 0.5 M HCl, again with EtOH to pH 7.0; vacuum-dried, 24 h,
cross-linked BS
Step III—Phosphorylation: cross-linked BS soaked in 200 mL DMF overnight, filtered and immersed again in 200 mL DMF+
5.04 g urea; added 3.1 g phosphoric acid drop-wise with constant stirring; stirred for 1 h. Temperature raised to 423 K and stirred for 2 h; cooled to RT, filtered, washed with 70% EtOH and DI until pH 7.0; washed with 0.1 M HCl and DI until pH 7.0; vacuum-dried for 2 days. Phosphorylated BS
Step IV—Fe(III) loading: treated with Fe(III) solution of concentration 55.85 mg/L (= 1 mM), Fe(III)-loaded BS
BAT—25 mg Fe(III)-loaded BS in 15 mL MS ([Arsenite/Arsenate] = 15 mg/L, 24 h, T = 303 K)
DAT—column packed with 0.1 g Fe(III)-loaded BS and conditioned with pH 4 water overnight. [Arsenate/Arsenite] = 15 mg/L; FR—0.098 mL/min
Bio-sorbent pre-treatment and Fe(III) loading enables direct removal of arsenite and arsenate together, without oxidizing arsenite into arsenate
DAT—0.91 mmol/g adsorbent; 99% removal of arsenic compared to 80% removal by cellulose control
DAT—Maximum arsenic adsorbed on the packed bed = 1.1 mg. Elution with 0.1 M HCl recovers 0.62 mg arsenic (60% recovery)
[113]
Orange peelsCadmium, Copper, LeadWash; Sun dry, 6 days; Grind, P.S. = 0.2 mm
BPT—Protonation—10 g dried peel soaked in 1 L of 0.1 M HNO3, 6 h; filtered, rinsed with DI; sun-dried for 6 days. Protonated BS
BAT—0.1–1 g protonated BS in 25 mL MS; [M2+] = 20 mg/L; CT—5–120 min;
T 98 K; pH 5.0. Shaking at 200 rpm
Maximum adsorption of Pb = 73.53 mg/g; Cu = 15.27 mg/g; Cd = 13.7 mg/g. Pb(99.5%) > Cu(89.57%) > Cd(81.03%) at [M2+] = 20 mg/L and BS loading of 4 g/L
Pb(96.3%) > Cu(93.3%) > Cd(85%) at [M2+] = 100–600 mg/L
[114]
Orange peelsLead, Cadmium, ZincWash; Dry; Grind, P.S. = 1–2 mm;
BPT—Protonation—10 g BS in 500 mL 0.1 N HNO3, stirred for 4 h at 120 rpm, 298 K, rinsed with DI until pH = 4.0, dried at 318 K
for 12 h; PS = 1–2 mm, protonated BS
DAT—Acrylic column; length 30 cm, diameter—1.3 cm, 5 g protonated BS wet-packed. Feed concentration [Pb2+] = 10.36 mg/L; [Cd2+] = 5.62 mg/L; [Zn2+] = 3.27 mg/L. Total feed to the column = 20 L; FR = 9 mL/min, pH 5.0, T = 298 K
Pb (85 mg/g) > Cd (44 mg/g) > Zn (20 mg/g)[98]
Citrus paradisi (Grapefruit) peelsZinc, NickelWash; Dry, 323 K until constant weight; Grind, P.S. 0.5–1.0 mm
BPT—(a) Blocking of –COOH group—9.0 g BS suspended in 633 mL CH3OH, and 5.4 mL HCl; stirred at 100 rpm, 6 h. Centrifuged, washed, freeze-dried. (b) Blocking of –OH group—5.0 g BS suspended in 100 mL HCHO, stirred at 100 rpm, 6 h. Centrifuged, washed and freeze-dried
BAT—100 mg BS in 100 mL MS; [M2+] = 300 mg/L, pH 5.0; CT = 120 min;
SRs—100 rpm; 298 K
Native peel BS
Ni2+—1.331 meq/g (84.73%)
Zn2+—1.512 meq/g (92.46%)
-COOH blocking reduces Ni2+ sorption by 78.57%, Zn2+ sorption by 73.31%
-OH blocking reduces Ni2+ sorption by 22.63% and Zn2+ sorption by 28.54%
[69]
Citrus peel pectinLeadCitrus peel pectins
(a) Low methoxylated (LM) pectin (methoxyl content 9%) and (b) high methoxylated (HM) pectin (methoxyl content 64%)
BAT/KS—0.02 g BS in 200 mL MS; [M2+] = 0.1–1.0 mM; pH ≤ 5.0; T 294–298 K; CT—2–1440 min. Background electrolyte concentration—0.01 M NaNO3
LM Pectin —0.86 mmol/g
HM Pectin —0.87 mmol/g
[115]
Orange wastePhosphateMetal-loaded orange waste bio-sorbent:
La(III)-loaded, Ce(III)-loaded and Fe (III)-loaded BS
BAT—25 mg of La(III)- and Ce(III)-loaded BS and 60 mg of Fe(III)-loaded BS in 15 mL phosphate solution: (Phosphate) = 20–40 mg/L; pH = 7.5 for La(III)/Ce(III)-loaded BS and 3.0 for Fe(III)-loaded BS experiments. SRs = 140 rpm, 24 h, 303 K
DAT—Glass column length—20 cm, diameter—0.8 cm; loaded with 150 mg of wet metal-loaded BS, (Phosphate) = 20–40 mg/L; FR = 7 mL/h
Phosphate adsorption by
M—loaded BS (% removal) in BAT
La(III)-loaded BS —13.84 mgP/g (98.5%)
Ce(III)-loaded BS —14.0 mgP/g (98.8%)
Fe (III)-loaded BS (99% removal)
13.63 mgP/g adsorbent in DAT
[116]
Orange wasteNickel, Cobalt, Cadmium, ZincWash; Dry, 323 K, 72 h. Ball mill—P.S. 0.1–0.2 mm; pore size—30.5 Å, BET surface area—128.7 m2/g
BPT—Different treatments, viz., isopropyl alcohol, alkali
saponification, acid oxidation to yield OP, PA, SNa, Sam/SCa, SOA, SCA, SPA; BAT—0.025 g BS in 15 mL MS; [M2+] = 0.001–0.01 M, CT—3 h
SPA—Ni2+—1.28 mol/Kg (95% increase)
SPA—Co2+ —1.23 mol/Kg (178% increase)
SCA—Cd2+—1.13 mol/Kg (60% increase)
SOA—Zn2+—1.23 mol/Kg (130% increase) in comparison to raw orange peel (OP)
Zn2+ → SCA>SNa>SOA>SPA>Sam>SCa>OP
Cd2+ → SOA>SCA>SPA>SNa>Sam>SCa>OP
Co2+/Ni2+
SPA>SCA>SOA>SNa>Sam>SCa>OP
[85]
Orange peelLead, Zinc, CopperWash; Dry, 333.15 K; Grind: P.S. ≤ 0.45 mm –(OP); BET—0.828 m2/g
BPT—100 g dried OP + 500 mL EtOH + 0.8 M NaOH + 0.8 M CaCl2; soak for 20 h, filter, wash until neutral pH—SCOP; BET—1.496 m2/g
BAT—0.1 g BS (OP and SCOP) in 25 mL MS; [Pb2+] = 200 mg/L; [Zn2+] = 50 mg/L; [Cu2+] = 50 mg/L; SRs—120 rpm; CT—0–12 h; T 298 K
Adsorption capacity SCOP/OP (mg/g)
Cu2+ → 70.73/44.28
Pb2+ → 209.8/113.5
Zn2+ → 56.18/21.25
Maximum adsorption was
found at pH 5.5
Pb2+(99.4%) > Cu2+(93.7%) > Zn2+(86.6%)
[117]
Orange PeelsChromiumWash; Dry, 353 K; Grind, PS ≤ 200 µm
BPT—100 g OP + 1 L of 0.1 M NaOH. Soak for 48 h; shake at 120 rpm, filter, wash, dry at 353 K—MOP (BET—0.8311 m2/g)
BAT—0.2–5.0 g MOP in 50 mL MS; [Cr(VI)] = 100 mg/L; SRs—120 rpm; T 298 K; CT = 30 min–4 h; pH 1–8.0
OP → 97.07 mg/g (39.9%)
AOP → 139.0 mg/g (41.4%)
Maximum adsorption was
found at pH 2.0 and BS dose of 4 g/L in 180 min
[118]
Citrus lemonCobaltWash; Dry, 353 K, 24 h; Grind
BPT—Thermal activation in air at 773 K, 1 h. Wash; dry, 373 K, 24 h. PS: BS 150–200
BAT—10 g/l BS; [Co2+] = 0–1000 mg/L; CT—10 h; SRs—200 rpm; pH 6.0
22 mg/g adsorbent[119]
Orange waste
Peel (OP),
Bagasse (OB),
Peel-bagasse (OPB)
LeadWash; Dry, 333.15 K; Grind, Sieve—BS—100-mesh
BPT—1 g BS + 20 m 0.1 M NaOH, agitation—2 h; Wash, dry—328 K, 24 h. 1 g modified BS + 8.3 mL 1.2 M citric acid; agitation—30 min; filter; dry—328 K, 24 h; heat—393 K, 90 min; Wash; dry-328 K, 24 h
BAT—0.5 g BS (OP, OB, OPB)/modified BS (OMP, OMB, OMPB) in 50 mL
MS; [Pb2+] = 700 mg/L; pH 2.0–6.0; T = 303 K; CT—10–1440 min
Highest adsorption capacity shown by
O-MP—84.53 mg/g
OP—55.52 mg/g
OB—46.90 mg/g
OMB—80.19 mg/g
OPB—32.55 mg/g
OMPB—73.37 mg/g
[120]
Pomelo PeelsCopperWash; Dry, 343 K, 2–3 h. Grind, Sieve—(Pomelo Peel, PP)
BPT—PP+ hot acidified water, pH 1.5, T = 388–393 K, 60 min; Filter. The solid peel residue after filtration is de-pectinated pomelo peel or DPP. Washed and dried at 343 K, 2–3 h; Grind: PS—0.42 mm. Filtrate contains pectin, precipitated by 95% EtOH.
BAT—0.5 g BS (PP, DPP) + 100 mL MS; [Cu2+] = 25 mg/L; pH = 2.0–6.0; SRs—150 rpm; T 298 K; CT—180 min
Pomelo peel (PP)—19.7 mg/g
De-pectinated pomelo peel (DPP)—21.1 mg/g
[121]
Lemon PeelCadmiumWash; Dry—323 K; Grind: PS—0.5–1.0 mm—native peel (NP)
BPT—(a) Protonation—10 g NP + 1.0 L of 0.1 M HCl, stir—6 h, 120 rpm; T 298 K; Filter, wash; Dry—323 K—protonated peels (PrP)
(b) 10 g NP + 400 mL diluted HCl; pH 1.5 at 358 K; Stir—150 rpm—2 h; T 358 K. Filter: Filtrate is coagulated by 95% EtOH. Wash with 60%, 80% and 95% EtOH with a retention time of 30 min in each washing; vacuum dry—323 K—peel pectic acid (PP)
(c) The solid residue left after pectic acid extraction is washed to remove all soluble sugars; vacuum dry—323 K—De-pectinated peels (DPP)
BAT—50 mg BS (NP, PrP, PP, DPP) in 50 mL MS; [Co2+] = 100 mg/L (1.78 mequiv./L); pH 5.0; SRs—120 rpm; T 298 K; CT—180 min
Native peel (NP)→ 1.92 mequiv./g
Protonated peels (PrP) → 2.44 mequiv./g
De-pectinated peels (DPP)→ 1.75 mequiv./g
Peel pectic acid (PP)→ 2.86 mequiv./g
PP > PrP > NP > DPP
[122]
Orange peels, Lemon Peels,
Lemon-based pectin peels (PP)
CadmiumNative orange and lemon peels → Wash,
Dry—311–313 K, 12 h; Grind
BPT—Protonation—Lemon-based pectin peels are treated with 0.1 N HNO3, 6 h; Dry for 12 h, 311–313 K; Wash, Dry; Grind—
PS: 0.7–0.9 mm—Protonated pectin peels (PPP)
BAT—0.1 g BS (NOP, NLP, PPP) in 50 mL MS; [Co2+] = 10–700 mg/L; Shake—6 h; pH 3.0/5.0
0.7–1.2 mequiv./g (39–67 mg/g)[123]
Citrus Pectin formsLeadDifferent forms of citrus pectins:
GA oligomers—Large DP, medium DM and small DP size class GA oligomers (galacturonic acid)
PME demethylated pectin (DM—50–80%)
Pectin from peel residue
Non-calcium-sensitive pectin
Lyophilized BS from single-state fermentation of citrus peel (hydrolysis)
BAT—50 mg BS in 50 mL MS; [Pb2+] = 0.5–1 g/l; pH 4.5; SRs—120 rpm; T 298 K; CT—6 h
Medium DP size class GA—380 mg/g, small DP size class GA—360 mg/g
Large DP size class GA—300 mg/g
PME demethylated pectin—220–270 mg/g
Pectin from peel residue- 140 mg/g
Non-calcium-sensitive pectin (NCSP)—200 mg/g
Lyophilized BS from fermented citrus peel—100 mg/g
[89]
Orange PeelsCopper
Cadmium
Wash; Dry—343 K, 24 h; Grind: PS—0.45 mm—native peel (NP)
BPT—(a) 60 g NP + 300 mL 1% NaOH + 300 mL EtOH; RT; 24 h. Wash; Dry, 343 K, 24 h, DPOP (De-pigmented orange peels)
(b) 30 g DPOP + 1 L 1% merceptoacetic acid, 12 h; Wash;
Dry—343 K, 12 h. Grind—PS ≥ 0.45 mm
BAT—50 mg BS in 10 mL MS; [M2+] = 0.05–1 g/l; SRs—120 rpm; CT—1.5 h;
T 298 K, pH—5.0–7.0
Cu2+—70.67 mg/g
Cd2+—136.05 mg/g
[86]
Lemon PeelCobaltWash; Dry—333.15 K, 24 h. Grind: PS= 1 mm—native peel (NP)
BPT—10 NP + 100 mL 2% IPA, 0.1 N NaOH, 0.1 N HCl, 0.1 N H2SO4, 0.1 N HNO3; 4 h, 303 K, Wash; Dry—333.15 K, 24 h
BAT—0.1 g BS in 50 mL MS; [Co2+] = 100 mg/L, T 303 K, SRs—150 rpm; CT—6 h
Native Peels—20.83 mg/g
Modified Peels—35.7 mg/g
[124]
D-limoneneMercuryDirect reaction between sulfur and D-limonene at (a) 443 K, 1 h; (b) 453 K, 50 mm Hg, 4 h; (c) 373 K < 1 mm Hg, 5 h55% removal[125]
Table 4. Citrus waste reuse as bio-sorbents for the removal of poisonous dyes from wastewater.
Table 4. Citrus waste reuse as bio-sorbents for the removal of poisonous dyes from wastewater.
AdsorbentDyeProcessing MethodMaximum Adsorption Capacity (Per Gram Adsorbent)Ref.
Citrus reticulataAcid Yellow-73Wash; Sun dry—7 days; Grind; Sieve through 50 ASTM mesh
BPT-Soak—10% formaldehyde; air-dried—3 days; Oven-dried—353 K, 2 h
BAT—1.0 g BS in 50 mL dye solution; (Dye) = 20 ppm; pH = 3.0; T = 323 K; SRs—100 rpm; CT—65 min;
96.46 mg/g−1L−1[75]
Citrus sinensisCongo Red,
Rhodamine B,
Procion orange
Wash; Sun dry—7 days; Grind—PS= 75–500 µm
BAT—250 mg BS in 50 mL Congo Red dye solution; (Congo Red) = 60 mg/L; CT—20–90 min, SRs—140 rpm; T 302 K; pH 5.0
500 mg BS in 50 mL Rhodhamine B and Procion orange dye solutions; (Dye) = 10 mg/L; CT—20–90 min, SRs—140 rpm; T 302 K; pH 3.0
Congo Red—22.4 mg/g; pH = 5.0 (76.6%)
Procion orange—1.3 mg/g; pH = 3.0 (49%)
Rhodamine B—3.22 mg/g; pH = 3.0 (38.43%)
[126]
Grapefruit peelsMethylene BlueWash; Sun dry—2 days; Grind—PS > 90 µm
BPT- Carbonization—Treat with (a) BS_ 88% orthophosphoric acid (1: 3 ratio), (b) ZnCl2, or (c) 98% H2SO4
Heat at 723–823 K for 0.75–1.5 h, wash with NH4OH and H2O to neutral pH, Dry—12 h; Charred citrus peel (CCF); PS= 135 µm
BAT—0.30–1.0 g CCF in 200 mL MB dye solution; (MB) = 20–100 mg/L; T 303 K; CT—8 h; pH 3.0–10.0
99.08% removal[72]
Orange PeelDirect Yellow-12Wash; Dry—423 K, 5 h; Grind
BPT—Carbonization—3 kg dried orange peel + 2.5 L 98% H2SO4, Stand—2 h; Boil—3 h; Add to ice-cold water, filter; Dry—453 K, 2 h; immerse in 5.0 L of 5% NaHCO3, wash to neutral pH; Dry—423 K, 3 h; Grind ≤ 0.200 mm
BAT—0.5 g in 100 mL dye solution; (Dye) = 75 mg/L; T 300 K; CT—2 h; pH 1.5–11.2; SRs—200 rpm
96% removal[127]
Grapefruit PeelsCrystal VioletWash; Dry—423 K, 5 h; Grind
BAT—BS = 0.1–3 g/L of dye solution; (Dye) = 5–600 mg/L; pH 6.0; SRs—100 rpm; T 303 K; CT—60 min
96% removal in 60 min.
Maximum adsorption capacity = 254.16 mg/g
[128]
Pomelo PeelMethylene Blue (Cationic Dye); Acid Blue (Anionic Dye)Wash; Air dry; Grind-PS: 1.0–2.0 mm
BPT—Microwave modification: BS + 1:1.25 by wt. NaOH. Microwave heating at 2.45 GHz, 800 W, 5 min; Wash with 0.1 M DI until neutral pH
BAT—0.20 g Modified BS in 200 mL dye solution; (Dye) = 50–500 mg/L; SRs—120 rpm;
T 303 K; CT—until equilibrium
Methylene Blue—501.1 mg/g
Acid Blue—444.45 mg/g
[129]
Pomelo PeelCongo RedWash; Dry—313 K, 48 h; Ball Mill-PS= 0.840 mm
BAT—1.0–3.0 g BS in 1 L dye solution; (Dye) = 20–120 mg/L; T 276.15–333.15 K,
pH 6.0–8.7; CT- 24 h
0.75–1.08 mg/g[130]
Citrus medica,
Citrus aurentifolia,
Citrus documana
Reactive Red 2 (Red M5B)Wash; Dry—373–393 K, 24 h; Crush
BPT—Carbonization—Heat at 773 K, N2; Liquid phase oxidation with 1 M HNO3; Wash, Dry—423 K, 12 h
BAT—3 g BS in 100 mL dye solution; (Dye) = 20 mg/L; T 298 K; CT = 5–90 min; SRs—120 rpm; pH 3.0–10.0
C. medica → 87% (0.5800 mg/g)
C. aurentifolia → 85% (0.5667 mg/g)
C. documana → 91% (0.6067 mg/g)
[70]
Citrus limonumMethyl Orange,
Congo Red
Wash; Dry—373 K, 24 h; Grind
BPT—Heat at 773 K in air, 1 h; Wash; Dry—373 K, 24 h; PS-BSS 100–250
BAT—0.1 g BS in 10 mL dye solution; (Dye) = 0.3–0.45 mM; T 298 K; CT-until equilibrium; pH 5.5–6.5
Methyl Orange → 50.3 mg/g
Congo Red → 34.5 mg/g
[131]
Citrus sinensis
bagasse
Methylene BlueWash; Dry—333 K, 72 h; Grind-PS: 0.25–0.75 mm
BAT—0.1 g in 100 mL dye solution; (Dye) = 50 mg/L; CT—24 h; T 303 K; pH 7.0
96.4 mg/g[132]
Orange Peel, Lemon PeelMethylene BlueWash; Dry—353 K, 24 h; Grind-PS: <3.0 mm
BAT—0.25 g in 25 cm3 dye solution; (Dye) = 50–1000 mg/dm3; T = 298 K; pH 2.0–3.0
Orange Peel → 4.76–95.03 mg/g
Lemon Peel → 4.41–92.1 mg/g
[133]
Grapefruit PeelLeather Dye mixture: Sella Solid Blue, Special Violet, Derma Burdeaux, Sella Solid OrangeWash; Dry—333.15 K, 24 h; Grind-PS: <0.5 mm
BPT—1.5 g BS + 150 mL of 1 M H2O2; Stirring—110 rpm, 24 h; Dry; Grind
BAT—0.3–1.5 g BS in 50 mL dye solution; (Dye) = 100–400 mg/L; pH 5.5; T 298 K; SRs—120 rpm; CT = 24 h
Untreated Grapefruit peel BS →45%
Modified Grapefruit peel BS → 80%
Maximum capacity → 1.1003 meq/g
Maximum uptake → 37.427 mg/g
[134]
Orange peelCongo Red,
Methyl Orange
Wash; Sun dry—72 h; Grind
BPT—(a) BS + 1% NaOH, EtOH; Filter; Wash, Air dry —OP (removal of lignin and pigments)
(b) 10 g OP + 100 mL DI; Stir and heat at 353 K; 120 min; Cool to RT, add N-vinyl-2-pyrrolidone, stir for 5 min—NVP/OP copolymer
(c) Transfer to glass tubes and irradiate with Gamma source—radiation dose (10–50 kGy); dose rate—1.46 kGy/h. Cross-linked NVP/OP hydrogel; Wash; Dry in air
BAT—1:1 BS in 20 mL dye solution; (Dye) = 10–50 mg/L; T = 293–333 K; pH = 7.0 for Congo Red and 6.0 for Methyl Orange; CT (Congo Red) = 6000 min, (Methyl Orange) = 4000 min
Congo Red → 4.8–26 mg/g
Methyl Orange → 4.6–10 mg/g
[73]
Citrus grandisMethylene BlueWash; Dry—333.15 K, 48 h; Grind: PS= 0.5–1.0 mm
BAT—0.20 g BS in 200 mL dye solution; (Dye) = 50–500 mg/L; pH 7.0; T 303 K;
CT = 5.15 h; SRs—100 rpm
344.83 mg/g at 303 K[135]
Mosambi peelsErichrome
Black T
Wash; Sun dry; Grind; Dry—333 K, 24 h
BPT—BS + Concentrated H2SO4 (1:1)- 24 h; Dry—378 K, 12 h; Wash with NaHCO3; Dry—378 K, Mosambi peel activated carbon (MPAC)
93.8%[136]
Citrus sinensis L.Remazol Brilliant BlueWash, Dry—333 K, 24 h; Grind: Wash, Dry—333 K; PS = 44–1180 µm
BAT—300 mg BS in 30 mL dye solution; (Dye) = 30, 100, 250 mg/L; SRs = 150 rpm;
T 293–333 K; CT = 24 h
11.62 mg/g[137]
Citrus sinensisReactive Blue 19,
Reactive Blue 49
Wash; Dry; Grind—PS < 0.25 mm—BS
BPT—(a) Immobilization: BS + sodium alginate (1:2). The resultant beads preserved in 0.02 M CaCl2 solution. Immobilized BS
(b) 1 g BS + 5% glacial acetic acid. Wash after 1 h; Dry—343 K, 24 h. Acetic acid-treated BS
BAT—0.5–1.5 g in 50 mL dye solution; (Dye) = 50–300 mg/mL; CT = 60–120 min;
pH 2.0; T 303 K; SRs—100 rpm
Reactive Blue 19
BS → 37.45 mg/g
Immobilized BS → 400.00 mg/g
Acetic acid-treated BS → 75.19 mg/g
Reactive Blue 49
BS → 135.16 mg/g
Immobilized BS → 80.00 mg/g
Acetic acid-treated BS → 232.56 mg/g
[66]
Citrus wasteMethylene BlueWash; Dry; Grind—383 K, 24 h; PS < 0.5 mm
BAT—0.70 g in 100 mL dye solution; (Dye) = 5–60 mg/L; T 300 K, CT = 180 min; SRs—120 rpm
3.2994 mg/g adsorbent at
(Dye) = 50 mg/L. Maximum removal percentage → 49.35% at 60 mg/L
[138]
Lime PeelRemazol Brilliant Blue RWash; Dry—T 378 K, 24 h; PS= 150 µm
BAT—1–9 g BS in 50 mL dye solution; (Dye) = 10–50 mg/L; SRs—120 rpm; CT = 24 h;
T 300 K
73–95.89% removal. Adsorption capacity of 7.29–9.58 mg/g
[139]
Pomelo PeelMalachite GreenWash; Dry—T 393 K, overnight; Grind
BPT—Carbonization—973 K, N2, 1 h; Char is soaked in KOH (1:1); Dry—423 K, overnight; Pyrolyze at 1073 K; N2 (150 cm3/min); At T = 1073 K, CO2 flow for 2 h; Cool to RT under N2 flux; Wash with 0.1 M HCl; Wash until neutral pH; BET—1357.21 m2/g
BAT—0.2 g BS in 100 mL dye solution; (Dye) = 50–400 mg/g; T 293–333 K; CT = 4 h; SRs = 120 rpm; pH 3.0–10.0
178.48 mg/g
Best result at pH = 8.0; T = 333.15 K
95.06% removal
[140]
Citrus reticulataIndigo Caramine DyeWash; Sun dry, 7 d;ays Oven dry—T 343 K, 4 h
BPT—(a) 100 g BS + 500 mL 20% formaldehyde, 3 h, FBS
(b) 100 g BS + 10% urea solution, 3 h, UBS
(a, b)—Filter, air dry; Oven dry, 343 K, 4 h; Grind: Sieve through 50-mesh ASTM = 297 µm
BAT—0.3–3.0 g BS in 50 mL dye solution; (Dye) = 25 mg/L; T 293–343 K; CT = 10–70 min; pH 1.0–10.0; SRs—100 rpm
Dried Peel (BS) → 5.90 mg/g
Formaldehyde-treated peel (FBS) → 14.79 mg/g
Urea-treated peel (UBS) → 71.07 mg/g
[74]
Orange PeelsAcid Violet 17Wash; Sun dry—4 days; Grind; PS: 53–500 µm
BAT—100–600 mg BS in 50 mL dye solution; (Dye) = 10 mg/L; pH 2.0–10.0; CT—80 min; T 303 K
19.88 mg/g; 87% removal at pH = 2.0 and
100% removal at pH = 6.27 at adsorbent dose of 600 mg in 50 mL of 10 mg/L dye solution
[141]
Citrus limetta PeelsMethylene BlueWash; Sun dry—4 days; Oven dry—363 K, 24 h; Grind; Sieve, 80 BSS mesh; Wash; Dry—T 378 K, 4 h; Grind; PS= 80–200 BSS
BAT—0.05 g in 25 mL dye solution; (Dye) = 25–250 mg/L; CT—3 h; pH 4.0
227.3 mg/g; 97–98% removal[48]
Citrus sinensis PeelsC.I. Direct Blue 77 dyeWash; Dry—T 378 K; Grind; PS = 75 µm
BAT—5–30 mg in 100 mL dye solution; (Dye) = 50 mg/L; pH 2.0–12.0; SRs = 125 rpm; CT = 60 min
59% removal; 9.43 mg/g[113]
Orange peelMethylene BlueWash; Dry—343 K, 5 h; Grind; PS= 75 µm, OP (dried orange peel)
BPT—60 g OP + 250 mL of 0.1 M NaOH, 24 h; Filter; Wash; Dry—MOP (Modified OP)
BAT—0.05 g in 25 mL dye solution; (Dye) = 35 mg/L; SRs = 160 rpm; pH 4.0; CT = 30 min; T 298–318 K
OP → 14.164 mg/g
MOP → 18.282 mg/g
[142]
Table 5. Thermodynamics and kinetics studies of various dyes and ions on citrus peel as adsorbents.
Table 5. Thermodynamics and kinetics studies of various dyes and ions on citrus peel as adsorbents.
Citrus Peel as AdsorbentAdsorbing
Substance
T
(Kelvin)
ΔG
(kJ mol1)
ΔH
(kJ mol1)
ΔS
(J mol1K1)
Isotherm FittedKinetics
Model
Ref.
Lemon peelCobalt298, 318−37.47, −38.56−21.254.61Langmuir
model
Pseudo-second-order[119]
Citrus sinensis
(Musambi peel)
Methylene Blue dye (MB)303, 308, 313, 318, 3230.76, 2.40, 2.90,
3.90, 4.40
−51.9 −0.18 Langmuir
model
Pseudo-second-order[132]
Citrus sinensis (Orange peel)Brilliant Blue dye (BB)303, 313,
323, 333
−14.70, −16.11,
−17.59, −18.32
22.82124.20Langmuir
and Freundlich isotherms
--[137]
Citrus sinensis (Orange peel)Arsenic293, 298,
303, 308,
313, 318
−30.18, −32.42,
−33.35, −34.48,
−35.06, −35.66
30.0−0.21Langmuir iothermPseudo-second-order[102,121]
Citrus grandis
(Pomelo peel)
Malachite Green dye (MB)303, 318,
333
−21.55, −22.89,
−24.22
9.160.15Langmuir isothermPseudo-second-order[140]
Citrus maxima
Pomelo peel
Copper298, 308,
318
−5.38, −4.19, −3.49−32.180.09Langmuir isothermPseudo-second-order[121]
Citrus limetta (Musambi peel)Methylene Blue dye (MB)293, 303,
313, 323
−7.87, −9.38,
−10.49, −12.41
35.13146.70Langmuir isothermPseudo-second-order[48]
Citrus limetta (Musambi peel)Fluoride298, 308,
318
−0.69, −3.45,
−5.61
72.58246.22Langmuir isothermPseudo-second-order[56]
Citrus limetta (Musambi peel)Chromium303,
313,
318,
323,
328
−420.21,
−725.16,
−726.24,
−2044.33,
−1151.51
1.914 × 10−357.44D–R isothermPseudo-second-order[152]
rGO-Kaffir Lime Peel ExtractMethylene Blue dye (MB)303, 313,
333
−5.98, −6.81,
−8.47
19.1582.93Langmuir isothermPseudo-second-order[151]
Magnetite Orange peel Crystal Violet303, 313,
323, 333
−5.92,−6.95, −8.02, −9.3528.02111.60Langmuir isothermPseudo-second-order model[153]
Lemon peels impregnated with phosphoric acidErythrosine-B (EB)
Rhodamine-B (RB)
298
298
−3.19
−2.97
27.43
24.41
102.76
91.87
Langmuir model
Pseudo-second-order model[154]
Orange (citrus sinensis) peels
by acid activation
Methylene Blue dye (MB)313, 323,
333
−20.5, −12.9,
−15.4
67.60281.70Langmuir modelPseudo-second-order model[155]
Table 6. Applications of RSM in the bio-sorption process.
Table 6. Applications of RSM in the bio-sorption process.
Object of Experimental DesignIndependent VariablesResponse(s)RemarksRef.
Parameters optimized for the preparation of adsorbent from citrus fruit peelWeight ratio (citrus fruit peel to activating
agent), temperature of carbonization
Operating parameters for carbonizing of citrus peelOptimized conditions for carbonization of citrus fruit peel are: weight ratio of peel to activating agent (3:1) at temperature of 798 K, and time of carbonization was 0.75 h[72]
Citrus fruit peel used in the removal of methylene blue (MB) dyeInitial concentration of MB, weight of CCFP and pH Percentage removal of MB 99.6225% MB removal at pH 3.64, weight of CCFP. Initial concentration of MB kept constant at 0.65 g and 20 mg/L. Prepared adsorbent is superior in terms of its porosity.[72]
Citrus limetta peel dust used for removal
of Cr(VI)
Initial concentration and pH of solutionCr(VI)
adsorption by musambi peel
Initial concentration 6.75, pH 4.29, dose 0.27 g/100 mL and contact time 56.40 min[152]
Adsorption of Brilliant Green (BG) dye by adsorbent prepared from Citrus limetta peelTemperature, pH, adsorbent dosage and contact timePercentage removal efficiency of BGThe model validations as optimum levels of the process parameters to obtain the maximum adsorption of dye of 85.17% at 313 K, pH 9, at an adsorbent dose of 3.5 g/L of aqueous dye solution and contact time of 240 min [160]
Adsorption of Eosin Y by the activated carbon (WCAC) prepared from waste citrus peelConcentration of Eosin Y, temperature and the adsorbent doseAdsorption of Eosin Y Maximum dye uptake of 59.3 mg/g at the dye concentration of 50 mg/L, temperature 333 K and the adsorbent dose of 0.1056 g[161]
Adsorption of antibiotic Trimethoprim studied by activated carbon prepared from waste citrus peel (WCAC)Concentration of solution, pH, temperature and adsorbent dose Adsorption efficiency of trimethoprim by WCACMaximum adsorption amount of TMP by WCAC calculated was 144.9 mg/g at 293 K[162]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mahato, N.; Agarwal, P.; Mohapatra, D.; Sinha, M.; Dhyani, A.; Pathak, B.; Tripathi, M.K.; Angaiah, S. Biotransformation of Citrus Waste-II: Bio-Sorbent Materials for Removal of Dyes, Heavy Metals and Toxic Chemicals from Polluted Water. Processes 2021, 9, 1544. https://doi.org/10.3390/pr9091544

AMA Style

Mahato N, Agarwal P, Mohapatra D, Sinha M, Dhyani A, Pathak B, Tripathi MK, Angaiah S. Biotransformation of Citrus Waste-II: Bio-Sorbent Materials for Removal of Dyes, Heavy Metals and Toxic Chemicals from Polluted Water. Processes. 2021; 9(9):1544. https://doi.org/10.3390/pr9091544

Chicago/Turabian Style

Mahato, Neelima, Pooja Agarwal, Debananda Mohapatra, Mukty Sinha, Archana Dhyani, Brajesh Pathak, Manwendra K. Tripathi, and Subramania Angaiah. 2021. "Biotransformation of Citrus Waste-II: Bio-Sorbent Materials for Removal of Dyes, Heavy Metals and Toxic Chemicals from Polluted Water" Processes 9, no. 9: 1544. https://doi.org/10.3390/pr9091544

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