Next Article in Journal
New Horizons in Chemical Functionalization of Endohedral Metallofullerenes
Next Article in Special Issue
The Budesonide-Hydroxypropyl-β-Cyclodextrin Complex Attenuates ROS Generation, IL-8 Release and Cell Death Induced by Oxidant and Inflammatory Stress. Study on A549 and A-THP-1 Cells
Previous Article in Journal
Rindera graeca (Boraginaceae) Phytochemical Profile and Biological Activities
Previous Article in Special Issue
Understanding Surface Interaction and Inclusion Complexes between Piroxicam and Native or Crosslinked β-Cyclodextrins: The Role of Drug Concentration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of β-Cyclodextrin in the Textile Industry—Review

by
Fabricio Maestá Bezerra
1,*,
Manuel José Lis
2,*,
Helen Beraldo Firmino
3,
Joyce Gabriella Dias da Silva
4,
Rita de Cassia Siqueira Curto Valle
5,
José Alexandre Borges Valle
5,
Fabio Alexandre Pereira Scacchetti
1 and
André Luiz Tessaro
6
1
Textile Engineering (COENT), Universidade Tecnológica Federal do Paraná (UTFPR), Apucarana 86812-460, Paraná, Brazil
2
INTEXTER-UPC, Terrassa, 0822 Barcelona, Spain
3
Postgraduate Program in Materials Science & Engineering (PPGCEM), Universidade Tecnológica Federal do Paraná (UTFPR), Apucarana 86812-460, Paraná, Brazil
4
Postgraduate Program in Environmental Engineering (PPGEA), Universidade Tecnológica Federal do Paraná (UTFPR), Apucarana 86812-460, Paraná, Brazil
5
Department of Textile Engineering, Universidade Federal de Santa Catarina (UFSC), Blumenau 89036-002, Santa Catarina, Brazil
6
Chemistry graduation (COLIQ), Universidade Tecnológica Federal do Paraná (UTFPR), Apucarana 86812-460, Paraná, Brazil
*
Authors to whom correspondence should be addressed.
Molecules 2020, 25(16), 3624; https://doi.org/10.3390/molecules25163624
Submission received: 30 June 2020 / Revised: 6 August 2020 / Accepted: 8 August 2020 / Published: 9 August 2020
(This article belongs to the Special Issue Cyclodextrin Chemistry and Toxicology)

Abstract

:
β-Cyclodextrin (β-CD) is an oligosaccharide composed of seven units of D-(+)-glucopyranose joined by α-1,4 bonds, which is obtained from starch. Its singular trunk conical shape organization, with a well-defined cavity, provides an adequate environment for several types of molecules to be included. Complexation changes the properties of the guest molecules and can increase their stability and bioavailability, protecting against degradation, and reducing their volatility. Thanks to its versatility, biocompatibility, and biodegradability, β-CD is widespread in many research and industrial applications. In this review, we summarize the role of β-CD and its derivatives in the textile industry. First, we present some general physicochemical characteristics, followed by its application in the areas of dyeing, finishing, and wastewater treatment. The review covers the role of β-CD as an auxiliary agent in dyeing, and as a matrix for dye adsorption until chemical modifications are applied as a finishing agent. Finally, new perspectives about its use in textiles, such as in smart materials for microbial control, are presented.

Graphical Abstract

1. Introduction

Since the first publication on cyclodextrins (CDs) in 1891, and the first patent in 1953, many technological advances have occurred, and the application of CDs has expanded [1]. According to Szejtli [2], over the years, CDs have been used in many diverse areas, and are identified, among all the receptor molecules, as the most important.
This scenario is no different in the textile sector, which constantly seeks technological innovation, especially in the dyeing, finishing, and water treatment sectors. With the market and consumers increasingly demanding environmental improvements, the development of new features combined with green processes has become a constant challenge [3]. Among the various materials that can be used for this purpose CDs stand out; they are oligosaccharides made up of glucose units that are organized in a conical trunk shape, providing a well-defined cavity for the formation of host–guest complexes with a series of molecules [4]. This versatility allows complexation with drugs, dyes, insecticides, essential oils, cosmetics, and other compounds [5,6,7,8,9], allowing this class of molecules to assume a leading role in the textile industry.
For the period from 1948 until today, since the term cyclodextrin started to be used as a research topic, 46,989 research papers have been reported by SCOPUS, and this number is continually increasing, as shown in Figure 1. This growth became significant in 1996, when the terms cyclodextrin and textile were combined and used as a research topic. These data were downloaded on 6 June 2020.
Furthermore, due to the presence of numerous hydroxyl groups either in the interior or exterior, CDs are susceptible to the addition of new functional groups, which may yield new properties and functionalities. Additionally, CDs have a set of outstanding characteristics, such as high biodegradability, high biocompatibility, and approval by Food and Drug and Administration (FDA), which makes them human and environmental-friendly [10]. Therefore, this review presents an overview of the use of cyclodextrins, especially beta CD and its derivatives, in the textile field. Although some general physicochemical characteristics are presented, the scope of the work is focused on the application of CDs in the areas of dyeing, finishing, and wastewater treatment.

2. General Characteristics of Cyclodextrins

Initially known as Schardinger dextrins [11], the widespread use of CDs as hosts in supramolecular chemistry is relatively recent. Because they are natural products, CDs are biocompatible and accepted in biological applications; therefore, there is a growing interest in them both scientifically and industrially [12]. The optimization of methods for obtaining and applying CDs is, as a result, constantly evolving [7].
CDs are obtained through the enzymatic degradation of potatoes, corn and rice starch, which gives a mixture of linear, branched, or cyclic dextrins [13]. Initially, the cyclization reaction of the starch glucopyranose linear chains occurs by the enzyme cyclomaltodextrin-glucanotransferase (CGTase) [14], produced for example by Bacilus firmus. This step results in a mixture of α-CD, β-CD and γ-CD, composed of six, seven and eight units of D-(+)-glucopyranose, respectively, joined by α-1,4 bonds [15].
Subsequently, the separation and purification of these three CDs are required [5,16]. Among the methods used for this purpose, the most simple and widely used to isolate α-, β- and γ-CD is selective precipitation, forming inclusion complexes with an appropriate guest molecule—for example, α, β and γ-CD crystallize with 1-decanol, toluene, and cyclohexadec-8-en-1-one, respectively [7]. However, separation has a relatively high cost, making the entire synthesis process expensive. Fortunately, over time, research intertwined with the production of CGTase has evolved and allowed the isolation of α, β and γ-CGTase, increasing yield and consequently decreasing the production costs of the CDs [7].
As a structural consequence of the glucose units connecting through α-1,4 bonds, CDs occur as conical trunk shaped structures which are capable of solubilizing and encapsulating hydrophobic molecules in an aqueous environment [4,17,18].
The structure of CDs consists of primary hydroxyl groups (C6) located at the end of the rings, and secondary hydroxyls (C2 and C3) located at the outer edge of the rings. Ether type oxygen and polar hydrogen groups (C3 and C5) are present inside the trunk of the CDs. While the external hydroxyls are responsible for the relative solubility of CDs in water and micro-heterogeneous environments, the glycosidic oxygen bridges and, consequently, their pairs of non-binding electrons facing the interior of the cavity give this region, in addition to its Lewis basic character, hydrophobicity [18,19], making it capable of complexing nonpolar molecules [20]. The chirality caused by the five chiral carbons of the D-glucose unit associated with the rigidity of the macrocycle due to the intramolecular hydrogen interactions between the 2- and 3-hydroxyl groups are fundamental characteristics of the chemistry of CDs. Table 1 lists the physical properties of natural CDs.
The data presented in Table 1 indicate an apparent regularity in some properties, however, irregularities have been observed regarding the degradation temperature and solubility. Szejtli [2] has suggested that the lower solubility of β-CD is associated with intramolecular hydrogen bonds occurring at the edge. Although it has the lowest solubility, β-CD and its derivatives are the most used due to factors such as simplicity in obtaining it, lower price, reduced sensitivity and irritability to skin, and the absence of mutagenic effects [22].
The limitations imposed by the reduced solubility combined with the expressive attractiveness cause CD derivatives to be synthesized industrially. The CD derivatives that are most industrially produced include methylated β-CD, heptakis (2,6-dimethyl)-β-CD, heptakis (2,3,6-trimethyl)-β-CD, hydroxypropyl-β-CD, peracetylated β-CD, sulfobutyl ether-CD, and sulfated CD [10,23,24]. All have greater solubility in water compared to natural CDs, expanding the spectrum of applications in the controlled release of drugs, increasing blood solubility and bioavailability of medicines and textile deodorants, and assisting in polymerization [5,21,25]. Studies on their toxicology, mutagenicity, teratogenicity, and carcinogenicity have been carried out, and have shown negative results [7,26,27]. CDs also have hemolytic activity in vitro; β-CD has the highest and γ-CD the lowest activity [27,28].
The industrial applications of CDs are very diverse; they have been used in the pharmaceutical industry, in agriculture, in the textile area, in food technology, in chemical and biological analysis, in environmental protection, and in cosmetics [6,9,29,30].
CDs play a significant role in the textile industry, as they can be used to remove surfactants from washed textile materials [31], as leveling agents in dyeing [32,33,34], in textile finishing [35,36,37,38,39], and in wastewater treatment [40,41,42,43].

Host-Guest Complex Formation

A striking feature of CDs is that they can form inclusion complexes with a variety of organic or inorganic compounds, allowing for the subsequent controlled release of these active compounds [44,45]. The fundamental factor for the guest molecule to be able to form a complex with the CD (host molecule) is its suitability within the cavity, which can be integral or partial [38,46,47]. Note that in the complexation, no covalent chemical bonds occur between the guest-molecule, nor is the compound closed within the macromolecular structure, which makes this type of complexation unique in terms of behavior as a modeler for the release of compounds [21,37].
Thus, the appropriate choice of the CD to be used for the possible formation of a complex is of great value. For small molecules, it is easier to form stable complexes with α-CD and β-CD, due to the compatibility of the volume of the guest molecule and the size of the CD cavity (Table 1). In the case of γ-CD, if the guest molecule is too small, the fit becomes unfavorable due to the much larger size of the cavity [19,48].
The general trend of CD complexation thermodynamics can be understood based on the concept of size; that is, by the analysis of the size and shape of the included molecule, and critical factors in the van der Waals interactions. Therefore, due to the fact that the cavity diameter of α-CD is much smaller than that of β-CD, and because the van der Waals forces are dependent on the distance between the molecules, it is expected that the forces induced by the complexation of extended chain molecules will be greater for α-CD than for β-CD [49].
As long as the fit-size requirements are satisfied, a number of other factors contribute to the complexation thermodynamics of the guest molecule in CDs. Considering only the aqueous environment, the following can be mentioned: (i) the entry of the hydrophobic portion of the guest-molecule into the CD cavity, (ii) the dehydration of the guest molecule and the exclusion of water molecules from the interior of the cavity, (iii) interactions of the hydrogen bonds between specific groups of the guest molecule and the OH of the receptor, and (iv) changes in conformation and/or stress reduction [49]. Although the preference for inclusion is of the hydrophobic portion (i), since charged species and hydrophilic groups are located in the bulk, certain groups with a hydrophilic character, such as phenolic OH, penetrate the cavity [50] and interact (iii) with the receptor.
According to Venturi et al. [48], after complexation in an aqueous environment, the new chemical environment experienced by the guest molecule causes changes in its chemical reactivity. In numerous cases, an increase in stability, reduction in volatility, stabilization against light, heat and oxidation [47,51], solubility of the guest molecule in the solution, increase in the speed of dissolution [52,53], and bioavailability [54,55] were observed. However, depending on the experimental condition and type of CD, the inclusion can be deleterious for the guest, for example, enhancing the chlorpromazine photodegradation as observed by Wang et al. [56].
In terms of the stoichiometry of the inclusion complex, the four most common types of complexes are considered in CDs: guest molecules with a 1:1, 1:2, 2:1 and 2:2 ratio [57]. However, Pinho et al. [10] point out that the most common cases of complexation are 1:1 and 1:2. These configurations are dependent on the size and structural aspect of the guest-molecule in relation to the cavity of the CDs, allowing the formation of stable inclusion complexes [58].
However, Rama et al. [59] highlight that the chemical composition of the guest molecule, as well as its solubility, ionization state, and molecular mass, in addition to the conditions of the medium, such as the pH, temperature, solvent used, and other parameters, influence the process. The choice of the appropriate medium, working temperature, pH, and other factors will determine the best conditions for the interaction between the CD and the guest molecule [60,61]. Voncina et al. [62] highlight that an increase in temperature in the dyeing of polyacrylonitrile with cationic dyes using β-CD improves complexation, which reaffirms the importance of these parameters in the process. Other determining factors are related to the type of cyclodextrin used and the method of preparation: physical mixing [63], kneading [64], atomization [65], lyophilization [66], or coprecipitation [67].
The mechanism of the formation of inclusion complexes can be divided into several steps; an illustration is shown in Figure 2. In the complexation of a substance in aqueous solution, the ends of the isolated CD cavity are opened in such a way that the guest molecule can enter the CD ring from both sides. There is, in principle, the absence of the guest molecule, and the slightly non-polar cavity, which acts as a host, is occupied by water molecules that are energetically unfavorable, as seen in Figure 2a. Given the nature of the polar–non-polar interaction, they can be easily replaced by a guest molecule that is less polar than water [14,68] (Figure 2b).
The molecules interact with each other as they are influenced by forces arising from the characteristics that are specific to each substance. Then, a complex phenomenon of molecular interaction occurs, since each interaction corresponds to a set of distinct forces [48]. Complexation is characterized by the absence of formation and the breaking of covalent bonds [69]. The driving force of the process is the increase in the entropy caused by the exit of water molecules present in the cavity and their consequent freedom [21]. Other forces also contribute to the maintenance of the complex, such as the release of the ring tension (especially for α-cyclodextrin), van der Waals interactions, hydrogen bonds, and changes in the surface tension of the solvent used as a medium for complexation [37,70].

3. Application of Cyclodextrins in the Textile Area

The wide range of applications of CDs has attracted the attention of many industries; however, according to Venturini et al. [48], initiatives for the industrial application of CDs have not been widely considered for three reasons: their scarce quantity and high prices, incomplete toxicological studies, and the fact that the knowledge obtained about CDs was not yet broad enough to envision their use in industry. The 1970s and 1980s were of fundamental importance for the diffusion of CDs in industry. Several studies have been successful in the production of CDs and their derivatives, and reliable tests have reduced doubts about their toxicity [2]. Their introduction into textile-related studies took on increasing relevance from the 1990s, according to SCOPUS data.
Bhaskara-Amrit et al. [31] emphasize that CDs have a very important role in textile processing and innovation; their use provides immediate opportunities for the development of environmentally friendly products and eco-textiles, in addition to having great potential in various applications. Cyclodextrins can be applied in the areas of spinning [71], pretreatment [72], dyeing [62,68,73], finishing [44,74,75,76,77,78], and dye removal [40,79,80,81], with dyeing, finishing, and water treatment being the most applicable in the textile area registered so far.

3.1. Dyeing Process

The use of cyclodextrins in the dyeing process can include their use as a dyeing aid, forming a complex with the dye [33,82], or as a chemical modification of the surface [83,84]. Figure 3 shows these two processes.
CDs can form a variety of inclusion complexes with textile dyes (Table 2), and this inclusion changes their properties, directly influencing the quality of the dyeing. Therefore, cyclodextrins can be used as auxiliaries in the dyeing process.

3.1.1. Cyclodextrin as an Auxiliary Agent in Dyeing

The introduction of new auxiliaries in the textile industry is feasible when they are used in low concentrations, are biodegradable, and do not affect the quality of the effluent. In addition to being biodegradable, CDs do not cause problems in textile effluents, and they improve the biodegradability of many toxic organic substances [17,68]. Their use is due to their formation of complexes with dyes, and they can be used to improve the uniformity of dyes and washing processes [31,82,86]. However, for the cyclodextrin to act as an auxiliary, the formation of the CD:dye complex is essential; if it is not formed, there will be no improvement in the quality of the dyeing [75,95].
In the dyeing of polyester fibers, dyes of the dispersed type are used for the coloring of the substrate [100]. These dyes are poorly soluble in water and generally require the use of surfactants [73], which are products from non-renewable sources that cause environmental problems and must be replaced to reduce damage [85]. Therefore, cyclodextrins are an alternative to these products that can maintain the quality of the coloring of the textile article [68].
Carpignano et al. [73] conducted studies on the application of β-CD with dispersed type dyes and polyester, and stated that the presence of β-CD positively affects color uniformity, intensity, and bath exhaustion when compared to dyeing using commercial surfactants. The insertion of cyclodextrins into the dyeing process decreases the amount of free dye molecules [33], causing the dyeing rate to decrease and favoring leveling [91]. This is due to the fact that the complex (CD:dye) has a molar mass greater than that of free dye, hindering its diffusion into the fiber, thus favoring the dye delay mechanism, which causes better leveling [101].
Another important synthetic fiber in the textile area is polyamide. This fiber presents, at the ends of its chains, carboxylic and amine groups, which gives it a substantivity for several classes of anionic dyes [102]. Commercially, acid dye is the most used due to the dye–fiber interaction in the acid medium, the leveling results, and the achieved colors [103]. Dispersed dye is seldom used due to its low adsorption and the possibility of a barre effect; therefore, in order to be able to use dispersed dyes for the dyeing of polyamide, it is necessary to improve the leveling and adsorption of this dye by the fiber. This can be achieved when using cyclodextrins as an auxiliary agent in dyeing [32].
Ferreira et al. [87] studied the dyeing of a polyamide 6 microfiber using dispersed dye complexed with cyclodextrins, and found that the complex changes the dyeing kinetics, improving its distribution in the fiber. There are also changes related to the thermodynamics of dyeing, since the dyeing also proved to be more intense, with greater adsorption of the dye by the fiber related to the increase in the dispersibility of the dye in the aqueous phase [88]. Similar results were found by Savarino et al. [89] when they dyed polyamide 6 with dispersed dyes, showing changes in the kinetic and thermodynamic phases of the dyeing. This indicates that cyclodextrins can replace additives from non-renewable sources and improve the dyeing and the effluent generated.
With regards to natural fibers, cotton is one of the most important textile fibers [104] and, in its dyeing, the use of reactive and direct dyes stands out. Reactive dye has structure groups that covalently bond with the fiber, improving the solidity; however, they have low affinity, requiring high amounts of electrolytes for good dyeing to occur [105]. Direct dye, on the other hand, has a high affinity for the cotton fiber [106] and it is often necessary to use retarding agents, such as alkaline salts, to prevent stains on the fabric and thus achieve better leveling [107].
The use of cyclodextrins can help to solve these dyeing problems. Parlat et al. [91] dyed cotton with reactive dye using cyclodextrins as an auxiliary. In this case, as a result of the complexation of the reactive dye, there was a good diffusion of the dye into the fiber, increasing its uniformity and color intensity.
In the works of Cireli et al. [82], the insertion of CDs occurred in the process of dyeing cotton with direct dye. The CDs acted as a retarding agent, forming complexes with the dye molecules, causing the dyeing speed to decrease, which improved the leveling.
Other works performed dyeing using β-CD, such as those of Voncina et al. [62], who dyed polyacrylonitrile with cationic dye and observed an improvement in color intensity and exhaustion when compared to the use of quaternary ammonium. Shibusawa et al. [92], who dyed cellulose acetate with dispersed dye, found that the complex formed between CD: dye changed the speed at which the chemical balance of the process was achieved, making it slower.
In general, cyclodextrins inserted as an auxiliary affect both the properties of the dyes and the dyeing kinetics, allowing improvements in exhaustion, uniformity, and in the quality of the effluent water. However, it is worth mentioning that this is only achieved when inclusion of the dye in the cavity of the CD is achieved.

3.1.2. Dyeing Chemical Modification

Some textile fibers present difficulty in dyeing due to the terminal groups present in their chains, causing some dyes to fail to create interactions, as is the case with polypropylene fibers [99] and vinylon fibers [95]. Other fibers present selectivity for dyes, such as cotton, which is not dyed by acidic and dispersed dyes [108]. However, promoting the modification of the surface of these fibers can cause new possibilities for the interactions between the dye and the fibers [109].
Cyclodextrins are polymers that can cause this chemical modification through incorporation into the fiber [99]. This incorporation can be seen as a pre-treatment for the dyeing or as a finishing, depending on the actions taken after modification. In this section, only the modifications for dyeing will be addressed and, in the next, finishing will be explored.
With cyclodextrins incorporated into the fabric, new groups and pores through which the dyes can fix become available. One fiber that presents difficulty in dyeing is cellulose acetate fiber, due to its compact structure, low content of polar groups, and hydrophobicity [110]. These factors make it difficult for dyes to diffuse in the fiber. To obtain better results in the dyeing process, Raslan et al. [75] treated the cellulose acetate fabric (38.5% acetyl) with monochlorotriazinyl-β-cyclodextrin (MCT-β-CD) using the padding technique to improve its dyeability. As a result, they were able to perform dyeing at a low temperature, improving the color intensity, and they also increased the diffusion of the dye within the fiber by about 70%.
In the case of polyester fibers, some authors have performed the process of acetylation [83] or coating [97] to modify the surface with CDs. This results in an improvement in the solidity of the dyeing [98], in addition to the possibility of dyeing with other classes of dyes. Zhang et al. [97], after performing the modification of polyester fiber, dyed this fabric with cationic dye. The fabric showed a gain in hydrophilicity, a reduction in the dyeing temperature to 70 °C, and interaction between the crosslinking carboxylate groups and the cationic dye, in addition to its complexation by the CDs.
Another work that used the modification of the polyester surface with cyclodextrins was carried out by Chen et al. [94]. In this work, the modification enabled a 47% increase in the color intensity in the stamping process, a fact associated with the greater sharpness and depth achieved by the dyes. In addition, the CDs, when chemically bonded to the fabric, can act as an anti-migration agent, because during the drying or curing of polyester fabrics dispersed dyes tend to migrate to the fabric surface and the CDs act as a dye sequestrant, consequently preventing this dyeing defect [83].
In the case of the modification of cotton fiber with cyclodextrins, several routes are possible, but the most used is esterification using citric acid or 1,2,3,4-butane tetra-carboxylic acid (BTCA) [111] as crosslinking. These changes will be covered in more detail in the next section. Rehan et al. [96] carried out the modification of cotton fiber with CDs and citric acid to perform dyeing with acid dye. These dyes present low affinity for the dyeing of cellulosic fiber [108]. After the modification, the authors realized that the dye was adsorbed by the cyclodextrins, which allowed the dyeing to achieve satisfactory solidity.
In general, the modification of the fiber surface through the insertion of cyclodextrins increases the adsorption of dye and allows a greater variability of dye classes in fibers that have no affinity, often achieving better color standards in multi-fiber items [97,98] and improving the efficiency of the dyeing process for fibers that require greater use of auxiliaries to achieve the proper color standard.

3.2. Textile Finishing

In the area of textile finishing, cyclodextrins can have many applications; they are able to absorb unpleasant odors, and act as an encapsulation agent for essential oils [38,76,78,112,113], vitamins [114], hormones [77] and biocides [6,115] in order to preserve compounds and/or control their release, as shown in Figure 4. The loading of active ingredients allows the incorporation of specific and desired functions into textile materials, which may act differently under particular uses, such as in medicine [116], cosmetics [117], and engineering [118].
In numerous cases, the complexation of active ingredients by CDs improves their physicochemical properties, controls their release, maintains bioavailability, increases shelf life, provides storage conditions, reduces environmental toxicity, increases chemical stability, protects against oxidation, and favors resistance to repeated washing [6,7,114,119,120].
In order to make it possible to incorporate these active molecules into the textile substrate, there is a need to fix the CDs in the fiber. Several methods have been proposed for the permanent fixation of CDs into textile fibers, and in some cases, there is a need for a first step—the modification of the cyclodextrins—so that they can be incorporated into the fabric. The selection of the best method for fixing CDs into a textile substrate depends on different factors, the main ones being reactivity of the cyclodextrins to the final application, and the type of fiber [23,121].

3.2.1. Preparation of Cyclodextrins

Cyclodextrins are capable of forming complexes with a wide range of molecules, but they cannot form a direct covalent bond with textile materials; therefore, some cyclodextrin derivatives have been synthesized with reactive groups to allow them to chemically bond to various substrates [122], as shown in Figure 5.
One of the most common reactive derivatives of cyclodextrins is MCT-β-CD, as seen in Figure 5a, synthesized through the reaction between cyanuric chloride and β-cyclodextrin [123]. MCT-β-CD is the most interesting derivative used on cellulosic substrates due to the simple bonding process in relatively mild conditions. The monochlorotriazine groups incorporated into the CDs react by a nucleophilic substitution mechanism, and form covalent bonds with the hydroxyl groups of the cellulose [124]. Another product that can be synthesized from MCT-β-CD is the cyclodextrin polymer (6A-O-triazine-β-cyclodextrin), produced by polycondensation using β-CD and cyanuric chloride [125].
Formation occurs due to nucleophilic substitution, in which the hydroxyl groups of the CDs react with the chlorine contained in the cyanuric chloride, and thus form the β-CD copolymer [125]. From the formation of this compound it is possible to create interactions with the hydroxyl groups present in the textile fibers; this occurs by substitution.
The modification of CDs can also be performed using itaconic acid (Figure 5b) containing carboxyl and vinyl groups. This bifunctional compound can be linked to the CDs via an esterification reaction, and its vinyl group can perform polymerization by free radicals [5,122]. Itaconic anhydride is obtained from itaconic acid at 110 °C in the presence of sodium hypophosphite [122]. From the modification of the CDs, the end containing the itaconic anhydride is able to bond with the textile fibers through covalent reactions.
Another CD modification for incorporation in textiles can be carried out via a reaction with acryloyl derivative (Figure 5c). The CDs are dissolved in dimethylformamide (DMF), mixed by stirring with triethanolamine (TEA), and reacted with acryloyl chloride dissolved in DMF, forming an acryloyl ester derivative [126]. The compound has a vinyl group on the side chain that is able to react with hydroxyl groups, and can be incorporated into the fibers [23,124].
In addition to the reaction through the incorporation of new chemical groups into the CDs, to make them more reactive hydroxyl groups can be oxidized, as can be seen in Figure 5d. The hydroxyl groups in the polysaccharides can be oxidized by a laccase/2,2,6,6-tetramethylpiperidine-1-oxyl enzyme catalyzed to convert the hydroxyl groups of the CDs into aldehyde groups that are capable of reacting with the amino groups of polyamide, silk, and wool [127].

3.2.2. Grafting of Cyclodextrins onto Textile Substrates

The most common procedure in the application of cyclodextrin into textiles is esterification, which can be done using modified cyclodextrins (Figure 5), or through a reaction using dimethylol urea [128], citric acid [111], BTCA [78,129], or other products.
Esterification can be defined as a nucleophilic substitution reaction of the acyl group catalyzed by a mineral acid, involving a carboxylic acid and an alcohol [130]. From there, a proton transitions from one oxygen to another, resulting in a second tetrahedral intermediate, and converts the -OH group into a leaving group, culminating in the loss of a proton that regenerates the acid catalyst, originating the ester [131].
Figure 6 shows the procedure for incorporating MCT-β-CD into cellulosic fiber. The interaction occurs due to the availability of the chlorine group present in MCT-β-CD and the hydroxyl group of cellulose, thus representing a second order nucleophilic substitution reaction [132].
MCT-β-CD is fixed on cellulosic fibers in alkaline conditions and, due to the covalent bond between the cellulosic chain and MCT-β-CD, the durability of β-CD in textile products is excellent [23,133].
Ibrahim et al. [134] also used MCT-β-CD for the functionalization of wool by a method of fixation in foularding. Due to the presence of -OH groups in the protein, it is also possible to perform nucleophilic substitution. As with polyamide fabrics and polyester/cotton blends, this β-CD derivative has also been grafted, making the fabric antibacterial and a receptor for drugs and essential oils, in addition to improving thermal stability and dyeability [128,130].
The MCT compound was also used to make polyester a functional fabric, made from alkaline hydrolysis, which created reactive hydroxyl groups on the surface of the polyester fibers able to react with MCT-β-CD covalently [39]. From the interaction with the cyclodextrins, the modified polyester can adsorb bioactive molecules [112].
Cyclodextrin compounds treated with itaconic anhydride can bind to cellulosic and polyamide fibers. In the case of cellulosic fibers, the fabric must be treated with a mixture of nitric acid (1%) and cholic ammonium nitrate to generate free radicals and, after drying, the cotton is treated with a derivative of CD itaconate, which is able to covalently bond to cellulosic fibers [5,122].
In addition to the processes using modified cyclodextrins, esterification between cyclodextrins and textile fibers can be achieved. In this case, the esterification reaction requires a crosslinking agent such as citric acid, BTCA, or other polycarboxylic acids [135]. The disadvantage of using citric acid is the yellowing of the cellulosic fabric in the curing phase [136]. This process includes two steps; in the first, a cyclic anhydride is formed between two groups of adjacent carboxylic acids and, in the second, the esterification reaction occurs between the acid anhydrides previously formed and the hydroxyl groups of the macromolecules of the fiber and of the cyclodextrins, to form ester bonds [23].
Figure 7 illustrates the bonding between CDs, through BTCA as a crosslinking agent, and -OH groups of fibers.
For the esterification reaction to occur, both sodium hypophosphite and the cure are used as catalysts [38]. The same process can be performed on other fibers that have -OH groups, such as cellulose, silk, polyamide, and wool [78].
Regarding the insertion of cyclodextrins into polyester fibers, they can be functionalized by forming a network of CDs that cover the fiber, forming a reticulated coating between β-CD and BTCA through a polyesterification reaction [38,137].
As shown in Figure 5d, the hydroxyl groups of the CD can be oxidized by enzymes, converting them into aldehyde groups, which are able to react with the amine groups of the wool fibers through a Schiff-based reaction [127]. Figure 8 shows this reaction.
In this way, the application of CDs in fibrous polymers occurs. The substrate undergoes a change at the surface that can transform it, in the future, into functionalized fabrics after the complexation of the bioactive molecules by the CDs present on the surface of the materials.
Table 3 shows some studies that used cyclodextrin for the functionalization of finished textiles.

3.3. New Trends in Textile Finishes Using Cyclodextrins

The use of citric acid as a reticulating agent was also a strategy adopted by Castriciano et al. [154] to design polypropylene fabric finished with hydroxypropyl β-CD. After complexation with tetra-anionic 5,10,15,20-tetrakis(4-sulfonatophenyl)-21H,23H-porphine (TPPS), the textile device was evaluated as a biocidal agent via antimicrobial Photodynamic Therapy (aPDT)—an alternative treatment to overcome the drug resistance associated with the indiscriminate use of antibiotics. The base of aPDT is the irradiation of a photosensitizer (PS) in the presence of oxygen, to generate reactive oxygen species (ROS) which attack the microorganisms at the target site (Figure 9). The PP-CD/TPPS fabric, containing 0.022 ± 0.0019 mg cm−2 of the TPPS, was capable of photokilling 99.98% of Gram-positive S. aureus, with low adhesion of bacteria to the textile. The aPDT approach was also used by Yao et al. [155] to develop biocidal materials based on beta cyclodextrins modified with hyaluronic acid (HA) for coating purposes. After the inclusion of PS methylene blue (MB), HA-CD/MB was tested against S. aureus, eradicating 99% of the bacteria at 0.53 ± 0.06 μg cm−2. The use of aPDT in textile finishing may represent a new class of smart textiles with high anti-microorganism potential.

3.4. Cyclodextrins in Textile Effluent Treatment

Cyclodextrins, in addition to being used as additives for the dyeing process when seeking improvements in washing, color intensity, and leveling, and as a functionalization agent, can be used to remove dyes and auxiliaries present in industrial effluents [40,156]. In the wastewater from the dyeing process, the presence of several types of dyes, surfactants, and salts can be an issue [157]. The dyes used in dyeing are compounds that are stable to oxidizing agents and light, have a complex structure, are non-biodegradable, and are highly soluble in water. Therefore, they are difficult to remove and can easily enter the ecosystem, affecting flora and fauna [79,158,159,160,161,162].
Various technologies for the treatment of water from the textile industry are used, such as photocatalytic oxidation [163], electrochemical oxidation [164], membrane separation [165], coagulation/flocculation [166], ozonation [167], and biological treatment [168], among others; however, there are restrictions regarding these processes, due to the high energy consumption and sludge generation. Thus, the search for processes that can eliminate residues from the dyeing and finishing processes is essential to alleviate major environmental problems. Lin et al. [169] and Crini et al. [170] point out that, among the different treatment systems, adsorption should be highlighted. It has been increasingly used, mainly due to its adaptability, easy operation, and low cost.
Among the adsorbents used, cyclodextrins are seen as a promising product [40] due to the high reactivity of the hydroxyl groups present in CDs for the adsorption process [171]. In addition, other advantages are related to its biodegradability, non-toxicity, availability [172], and the possibility of them interacting with the hydrophobic chain of surfactants, keeping them within its cavity [173]. In this way, they can form an insoluble CD:dye:surfactant system that can be removed from the water [79]. In general, Crini et al. [170] showed that the use of cyclodextrins as a dye adsorbent can be carried out by two methods, shown in Figure 10.
In the first method, cyclodextrins are incorporated into an insoluble matrix (nanoparticles, composites, nanotubes, and others), while in the second, CDs form an insoluble polymer capable of adsorbing the dyes. Table 4 shows some studies that have used cyclodextrin for the removal of textile dyes.

3.4.1. Cyclodextrin Matrix

A material used for the adsorption of dyes present in effluents must have a high adsorption capacity, ease of regeneration, mechanical resistance, and ability to adsorb a variety of dyes [182]. This last characteristic is often neglected, with experiments being carried out on solutions that contain only one type of dye; however, Debnath et al. [161] showed that most wastewater contains a mixture of different dyes, which affect the behavior of the adsorption system differently to a single dye system.
Therefore, cyclodextrins, due to their well-defined structure, can guarantee high reactivity for the adsorption of various dyes [171], and this can be improved if they are inserted into other adsorbent materials. These hybrid materials have a high adsorption capacity due to large specific surfaces and pore volume [43].
One of the techniques used for the production of promising adsorbent materials is electrospinning [183]. Abd-Elhamid et al. [162] produced nanocomposites using polyacrylic as an incorporation matrix and graphene oxide and cyclodextrin as adsorbent materials. According to the author, the nanocomposite is easy to prepare and has a high sorption capacity and is easy to remove from water. The combination of cyclodextrins and graphene to obtain a hybrid adsorbent material was also used by Liu et al. [176], who, in this case, also used poly(acrylic acid). This nanocomposite showed efficiency in pollutant adsorption, water dispersibility due to the hydrophilicity of the polymer, ease of regeneration, and a small loss of adsorption capacity.
Cyclodextrins can also be used in the production of biosorbents, together with chitosan. Chitosan is a compound rich in hydroxyl and amino groups, which allows interactions with organic and inorganic compounds [184,185]. However, chitosan, if not modified, can dissolve in acidic solutions because of the protonation of amino acids, hindering the adsorption of dyes [186]. To avoid such a problem, chitosan can be crosslinked with carboxylic acids and, to improve this biosorbent, Zhao et al. [187] chemically incorporated cyclodextrins into chitosan by means of esterification using citric acid, obtaining a biosorbent with a high capacity for adsorption of reactive dyes from textile effluents.
Chen et al. [157] showed that some researchers have grafted β-CD into insoluble solids, such as zeolite, activated carbon, silica gel and magnetic materials, obtaining good adsorption results. These characteristics show that adsorbent materials with cyclodextrin incorporation have great potential for applications in wastewater treatment, due to their large amount of hydroxyl groups, hydrophobic cavity, and interactions with organic and inorganic compounds.

3.4.2. Cyclodextrin Polymers

The synthesis of cyclodextrin polymers, especially those that are insoluble in water, has aroused growing interest given their applications in water treatment. Among the various methods of obtaining them, deprotonation stands out, in which the hydroxyl anion can be used in SN2 type polymerization reactions, direct dehydration in the presence of appropriate diodes and diacids, and condensation in the presence of a series of linkers [188]. In addition to polymerization, some studies have used β-CD for the development of organic-inorganic hybrid systems for the removal of dyes, such as magnetic CD polymers [40,189] and Halloysite−Cyclodextrin Nanosponges [190].
Crini et al. [81], using epichlorohydrin as a crosslinking agent for obtaining β-CD polymers, evaluated their efficiency in removing various dyes (acid blue (AB25), basic blue (BB3), reactive blue (RB19), dispersive blue (DB3) and direct red (DR81)). The capability to remove dyes by these polymers followed the order AB25 > RB19 > DB3 > DR81 >> BB3, with AB25 being close to 100% removed. The same author also prepared β-CD/carboxy methylcellulose polymers using the same crosslinking agent for the removal of Basic Blue 3, Basic Violet 3 and Basic Violet 10. Kinetic and equilibrium studies suggested that the process occurs by chemisorption, with an adsorptive capacity of 53.2, 42.4 and 35.8 mg of dye per gram of polymer for BV 10, BB 3 and BV 3, respectively [160].
Pellicer et al. [171] also used epichlorohydrin as a crosslinking agent to synthesize polymers of β-CD and HP-β-CD, which were used to remove the azo dye Direct Red 83:1. The adsorption capacity of the polymer synthesized from β-CD was approximately six times greater than that obtained using HP-β-CD. Ozmen and Yilmaz [191] used β-CD polymer, prepared using 4-4-methylene-bis-phenyldiisocyanate (MDI), to remove Congo red dye. The authors observed 80% removal after one hour of contact in solution at pH 5.8. The same authors, using MDI and hexamethylene diisocyanate (HMDI) with crosslinking agents, synthesized β-CD polymers and evaluated their adsorptive capacities against the azo dyes Evans Blue and Chicago Sky Blue. At pH 2, the polymers showed around 50% removal.
Jiang et al. [192] synthesized a new polymer of β-CD for the removal of methylene blue. The strategy used by the authors was the use of tetrafluoroterephtalonitrile (TFPN) as a crosslinking agent, which, after being hydrolyzed, generates sites of carboxylic acids that interact electrostatically with the MB at the appropriate pH. A maximum adsorption capacity of 672 mg/g of the polymer was observed and, even after four cycles of adsorption/desorption, the capacity of the material remained high. The same group of researchers used a similar strategy for the synthesis of β-CD polymers, however, the nitrile groups of TFPN were modified with ethanolamine. This strategy enabled the selective removal of MO in a mixture of MO and MB. The polymer also showed a high adsorptive capacity for MO (602 mg/g) and Congo red (1085 mg/g). Recently, the selective removal of the anionic dye Orange G in a mixture with methylene blue has also been carried out by modifying the TFPN nitriles to form amide groups [41]. An innovative strategy using molecularly imprinted polymers (MPI) from chitosan and β-CD was used for the selective separation of Remazol Red 3BS in a trichromatic mixture. This new polymer also showed a high adsorption capacity after four cycles of use [168].
Some multifunctional CD polymers have also been developed for the simultaneous removal of dyes and other contaminants (bisphenol and heavy metals). Zhou et al. [42] synthesized a polymer of β-CD using citric acid as a crosslinking agent, which, after esterification, was grafted with 2-dimethylamino ethyl methacrylate monomer (DMAEMA) for the polymerization reaction. This elegant strategy allows modulating of the zeta potential of the adsorbent with the pH, enabling its electrostatic interaction with anionic (MO) or cationic (MB) dye. Simultaneously, the material can adsorb Bisphenol A inside the CD, and its interaction with the CD is unchanged between pH 2 to 10. The adsorption capacity at equilibrium for Bisphenol A was 79.0 mg/g, while the adsorption capacity of MO and MB was, respectively, 165.8 and 335.5 mg/g.
Zhao et al. [193] presented an elegant strategy for the treatment of industrial wastewater by means of a bifunctional adsorbent, consisting of a polymer of ethylene diamine tetra-acetic acid and β-CD (EDTA-β-CD). This bifunctional agent can simultaneously remove metals and dyes from wastewater, since β-CD has the ability to include dyes while EDTA becomes a site for metals. In experiments with binary systems containing Cu2+ and dyes (methylene blue, safranin O or crystal violet), the authors observed an increase in the adsorption capacity of the metal, but no significant change in the adsorption of the dyes, compared to experiments in systems with the isolated metal. The increase in the adsorption of the metal in binary Cu2+-dye systems was attributed to the presence of the complexed dye in the CD, which provides extra groups containing nitrogen that become new sites for the adsorption of metals.
Despite the efficiency of the CD-based polymer in removing dyes and other agents in the textile process, some important points should be highlighted. Most of the studies presented above still need to be applied at a high scale level (in a real industrial system). Another important issue that should be emphasized is the regenerability of the CDs, making the process more ecofriendly and viable, with a lower cost.

4. Final Considerations and Future Perspectives

The increasing use of CDs in the textile industry is the result, among other factors, of the versatility of these cyclic molecules and the benefits of their use across the productive chain of this sector. Their unique ability to form an inclusion complex with a wide variety of molecules allows their use in several sectors. CDs are able to include dyes, repellents, insecticides, essential oils, caffeine, vitamins, drugs and surfactants, among other substances. Although they are used in the spinning and pretreatment areas, it is in the dyeing, finishing, and water treatment processes that β-CD and its derivatives have the greatest applicability.
The advantages of using CDs in dyeing include changes in bath exhaustion, color uniformity, less effluent treatment, dye savings, and the fact that they are biodegradable. They can be used as a dyeing aid, or as a surface modifying agent that absorbs more dye.
With regard to finishing, different types can be made with CDs, expanding the range of applications for these textiles and giving rise to a new class of materials called functional or intelligent textiles.
It is foreseeable that the use of CDs will continue to expand to keep up with the demands for differentiated products, and fill the gap that still exists in the literature around their application in the textile area, aiming at the optimization of the processes and viable results for industrial use.
This functionalization of CDs in substrates opens the door for the development of new products, such as medical textiles. With the new reality caused by the SARS-CoV-2 pandemic, the development of antiviral textiles is on the rise, and many of these new materials could be generated from technologies that use CDs. Furthermore, the transposition of new medical treatment technologies into textile materials from CDs is already a reality. An example is the use of β-CD for the development of textiles aiming at the photodynamic inactivation of microorganisms.
Finally, the capacity of CDs to adsorb and separate pollutants (dyes, metals, surfactants, etc.) from industrial waste is important with regards to environmentally sustainable industrial processes. In addition to adaptability and ease of operation, their biodegradability and lack of toxicity make CDs stand out in different areas.
Without a doubt, the use of CDs in basic and applied research around the development of new materials is fundamental, and should be the focus of many future studies seeking sustainable alternatives in the textile area.

Author Contributions

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

Funding

This research received no external funding.

Acknowledgments

INTEXTER-UPC, UTFPR-AP and National Council for Scientific and Technological Development (CNPq). The author José Alexandre Borges Valle are grateful to CAPES-PRINT, project number 88887.310560/2018-00.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Villiers, A. Sur la fermentation de la fécule par l’action du ferment butyrique. C. R. Acad. Sci. 1891, 112, 536–537. [Google Scholar]
  2. Szejtli, J. Introduction and General Overview of Cyclodextrin Chemistry. Chem. Rev. 1998, 98, 1743–1754. [Google Scholar] [CrossRef] [PubMed]
  3. Alonso, C.; Martí, M.; Barba, C.; Lis, M.; Rubio, L.; Coderch, L. Skin penetration and antioxidant effect of cosmeto-textiles with gallic acid. J. Photochem. Photobiol. B Boil. 2016, 156, 50–55. [Google Scholar] [CrossRef]
  4. Rasheed, A. Cyclodextrins as Drug Carrier Molecule: A Review. Sci. Pharm. 2008, 76, 567–598. [Google Scholar] [CrossRef]
  5. Radu, C.-D.; Parteni, O.; Ochiuz, L. Applications of cyclodextrins in medical textiles—Review. J. Control. Release 2016, 224, 146–157. [Google Scholar] [CrossRef]
  6. Nardello-Rataj, V.; Leclercq, L. Encapsulation of biocides by cyclodextrins: Toward synergistic effects against pathogens. Beilstein J. Org. Chem. 2014, 10, 2603–2622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Leclercq, L. 17 - Smart medical textiles based on cyclodextrins for curative or preventive patient care. In Active Coatings for Smart Textiles; Hu, J., Ed.; Woodhead Publishing Series in Textiles; Woodhead Publishing: Cambridge, UK, 2016; pp. 391–427. ISBN 978-0-08-100263-6. [Google Scholar]
  8. Crini, G. Review: A History of Cyclodextrins. Chem. Rev. 2014, 114, 10940–10975. [Google Scholar] [CrossRef]
  9. Buschmann, H.-J.; Schollmeyer, E. Applications of cyclodextrins in cosmetic products: A review. J. Cosmet. Sci. 2002, 53, 185–191. [Google Scholar]
  10. Pinho, E.; Grootveld, M.; Soares, G.; Henriques, M. Cyclodextrins as encapsulation agents for plant bioactive compounds. Carbohydr. Polym. 2014, 101, 121–135. [Google Scholar] [CrossRef] [Green Version]
  11. French, D. The Schardinger Dextrins. In Advances in Carbohydrate Chemistry; Wolfrom, M.L., Tipson, R.S., Eds.; Academic Press: New York, NY, USA, 1957; Volume 12, pp. 189–260. [Google Scholar]
  12. Fernández, M.A.; Silva, O.F.; Vico, R.V.; de Rossi, R.H. Complex systems that incorporate cyclodextrins to get materials for some specific applications. Carbohydr. Res. 2019, 480, 12–34. [Google Scholar] [CrossRef]
  13. Loftsson, T.; Duchêne, D. Cyclodextrins and their pharmaceutical applications. Int. J. Pharm. 2007, 329, 1–11. [Google Scholar] [CrossRef] [PubMed]
  14. Matioli, G. CICLODEXTRINAS E SUAS APLICAÇÕES EM: Alimentos, Fármacos, Cosméticos, Agricultura, Biotecnologia, Química Analítica e Produtos Gerais|Eduem—Editora da UEM, 1st ed.; Eduem: Maringá, Brazil, 2000; ISBN 85-85545-46-1. [Google Scholar]
  15. Challa, R.; Ahuja, A.; Ali, J.; Khar, R.K. Cyclodextrins in drug delivery: An updated review. AAPS PharmSciTech 2005, 6, E329–E357. [Google Scholar] [CrossRef] [PubMed]
  16. Schöffer, J.D.N.; Klein, M.P.; Rodrigues, R.C.; Hertz, P.F. Continuous production of β-cyclodextrin from starch by highly stable cyclodextrin glycosyltransferase immobilized on chitosan. Carbohydr. Polym. 2013, 98, 1311–1316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Andreaus, J.; Dalmolin, M.C.; Junior, I.B.D.O.; Barcellos, I.O. Application of cyclodextrins in textile processes. Química Nova 2010, 33, 929–937. [Google Scholar] [CrossRef] [Green Version]
  18. Amiri, S.; Amiri, S. Cyclodextrins: Properties and Industrial Applications; John Wiley&Sons: Hoboken, NJ, USA, 2017; ISBN 978-1-119-24760-9. [Google Scholar]
  19. Harata, K. Structural Aspects of Stereodifferentiation in the Solid State. Chem. Rev. 1998, 98, 1803–1828. [Google Scholar] [CrossRef]
  20. Faisal, Z.; Kunsági-Máté, S.; Lemli, B.; Szente, L.; Bergmann, D.; Humpf, H.-U.; Poór, M. Interaction of Dihydrocitrinone with Native and Chemically Modified Cyclodextrins. Molecules 2019, 24, 1328. [Google Scholar] [CrossRef] [Green Version]
  21. Del Valle, E.M.M. Cyclodextrins and their uses: A review. Process. Biochem. 2004, 39, 1033–1046. [Google Scholar] [CrossRef]
  22. Singh, N.; Yadav, M.; Khanna, S.; Sahu, O. Sustainable fragrance cum antimicrobial finishing on cotton: Indigenous essential oil. Sustain. Chem. Pharm. 2017, 5, 22–29. [Google Scholar] [CrossRef]
  23. Shabbir, M.; Ahmed, S.; Sheikh, J.N. Frontiers of Textile Materials: Polymers, Nanomaterials, Enzymes, and Advanced Modification Techniques; Scrivener Publishing: Beverly, MA, USA, 2020; ISBN 978-1-119-62036-5. [Google Scholar]
  24. Semeraro, P.; Rizzi, V.; Fini, P.; Matera, S.; Cosma, P.; Franco, E.; García, R.; Ferrándiz, M.; Núñez, E.; Gabaldón, J.A.; et al. Interaction between industrial textile dyes and cyclodextrins. Dye. Pigment. 2015, 119, 84–94. [Google Scholar] [CrossRef]
  25. Gao, S.; Liu, Y.; Jiang, J.; Ji, Q.; Fu, Y.; Zhao, L.; Li, C.; Ye, F. Physicochemical properties and fungicidal activity of inclusion complexes of fungicide chlorothalonil with β-cyclodextrin and hydroxypropyl-β-cyclodextrin. J. Mol. Liq. 2019, 293, 111513. [Google Scholar] [CrossRef]
  26. Saenger, W. Cyclodextrin Inclusion Compounds in Research and Industry. Angew. Chem. Int. Ed. 1980, 19, 344–362. [Google Scholar] [CrossRef]
  27. Stella, V.J.; He, Q. Cyclodextrins. Toxicol. Pathol. 2008, 36, 30–42. [Google Scholar] [CrossRef] [PubMed]
  28. Irie, T.; Uekama, K. Pharmaceutical Applications of Cyclodextrins. III. Toxicological Issues and Safety Evaluation. J. Pharm. Sci. 1997, 86, 147–162. [Google Scholar] [CrossRef] [PubMed]
  29. Cunha-Filho, M.S.S.D.; SÁ-BARRETO, L.C.L. Utilização de ciclodextrinas na formação de complexos de inclusão de interesse farmacêutico. Revista de Ciências Farmacêuticas Básica e Aplicada 2007, 28, 1–9. [Google Scholar]
  30. Bilensoy, E. Cyclodextrins in Pharmaceutics, Cosmetics, and Biomedicine: Current and Future Industrial Applications; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2011. [Google Scholar]
  31. Bhaskara-Amrit, U.R.; Agrawal, P.B.; Warmoeskerken, M.M.C.G. Applications of β-cyclodextrins in textiles. Autex Res. J. 2011, 11, 94–101. [Google Scholar]
  32. Savarino, P.; Viscardi, G.; Quagliotto, P.; Montoneri, E.; Barni, E. Reactivity and effects of cyclodextrins in textile dyeing. Dye. Pigment. 1999, 42, 143–147. [Google Scholar] [CrossRef]
  33. Bezerra, F.M.; Moraes, F.F.D.; Santos, W.L.F.; Santos, J.C. Emprego de β-ciclodextrina como auxiliar no tingimento de fibras têxteis. Química Têxt. 2013, 27, 12–21. [Google Scholar]
  34. Grigoriu, A.; Luca, C.; Grigoriu, A. Cyclodextrins Applications in the Textile Industry. Cellul. Chem. Technol. 2007, 1, 103–112. [Google Scholar]
  35. Crupi, V.; Ficarra, R.; Guardo, M.; Majolino, D.; Stancanelli, R.; Venuti, V. UV–vis and FTIR–ATR spectroscopic techniques to study the inclusion complexes of genistein with β-cyclodextrins. J. Pharm. Biomed. Anal. 2007, 44, 110–117. [Google Scholar] [CrossRef]
  36. Abdel-Halim, E.S.; Abdel-Mohdy, F.A.; Fouda, M.M.G.; El-Sawy, S.M.; Hamdy, I.A.; Al-Deyab, S.S. Antimicrobial activity of monochlorotriazinyl-β-cyclodextrin/chlorohexidin diacetate finished cotton fabrics. Carbohydr. Polym. 2011, 86, 1389–1394. [Google Scholar] [CrossRef]
  37. Marques, H.M.C. A review on cyclodextrin encapsulation of essential oils and volatiles. Flavour Fragr. J. 2010, 25, 313–326. [Google Scholar] [CrossRef]
  38. Lis, M.J.; Carmona, O.G.; Carmona, C.G.; Bezerra, F.M. Inclusion Complexes of Citronella Oil with β-Cyclodextrin for Controlled Release in Biofunctional Textiles. Polymers 2018, 10, 1324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Popescu, V.; Vasluianu, E.; Popescu, G. Quantitative analysis of the multifunctional finishing of cotton fabric with non-formaldehyde agents. Carbohydr. Polym. 2014, 111, 870–882. [Google Scholar] [CrossRef] [PubMed]
  40. Hu, X.; Hu, Y.; Xu, G.; Li, M.; Zhu, Y.; Jiang, L.; Tu, Y.; Zhu, X.; Xie, X.; Li, A. Green synthesis of a magnetic β-cyclodextrin polymer for rapid removal of organic micro-pollutants and heavy metals from dyeing wastewater. Environ. Res. 2020, 180, 108796. [Google Scholar] [CrossRef] [PubMed]
  41. Xu, M.-Y.; Jiang, H.-L.; Xie, Z.-W.; Li, Z.-T.; Xu, D.; He, F.-A. Highly efficient selective adsorption of anionic dyes by modified β-cyclodextrin polymers. J. Taiwan Inst. Chem. Eng. 2020, 108, 114–128. [Google Scholar] [CrossRef]
  42. Zhou, Y.; Hu, Y.; Huang, W.; Cheng, G.; Cui, C.; Lu, J. A novel amphoteric β-cyclodextrin-based adsorbent for simultaneous removal of cationic/anionic dyes and bisphenol A. Chem. Eng. J. 2018, 341, 47–57. [Google Scholar] [CrossRef]
  43. Teng, M.; Li, F.; Zhang, B.; Taha, A.A. Electrospun cyclodextrin-functionalized mesoporous polyvinyl alcohol/SiO2 nanofiber membranes as a highly efficient adsorbent for indigo carmine dye. Colloids Surf. A: Physicochem Eng. Asp. 2011, 385, 229–234. [Google Scholar] [CrossRef]
  44. Nostro, P.L.; Fratoni, L.; Ridi, F.; Baglioni, P. Surface treatments on Tencel fabric: Grafting with β-cyclodextrin. J. Appl. Polym. Sci. 2003, 88, 706–715. [Google Scholar] [CrossRef]
  45. Perchyonok, V.T.; Oberholzer, T. Cyclodextrins as Oral Drug Carrier Molecular Devices: Origins, Reasons and In-vitro Model Applications. Curr. Org. Chem. 2012, 16, 2365–2378. [Google Scholar] [CrossRef]
  46. Acartürk, F.; Çelebi, N. Cyclodextrins as Bioavailability Enhancers. In Cyclodextrins in Pharmaceutics, Cosmetics, and Biomedicine; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2011; pp. 45–64. [Google Scholar]
  47. Andersen, F.M.; Bundgaard, H. Inclusion complexation of metronidazole benzoate with β-cyclodextrin and its depression of anhydrate-hydrate transition in aqueous suspensions. Int. J. Pharm. 1984, 19, 189–197. [Google Scholar] [CrossRef]
  48. Venturini, C.D.G.; Nicolini, J.; Machado, C.; Machado, V.G. Propriedades e aplicações recentes das ciclodextrinas. Química Nova 2008, 31, 360–368. [Google Scholar] [CrossRef] [Green Version]
  49. Rekharsky, M.V.; Inoue, Y. Complexation Thermodynamics of Cyclodextrins. Chem. Rev. 1998, 98, 1875–1918. [Google Scholar] [CrossRef] [PubMed]
  50. Ross, P.D.; Rekharsky, M.V. Thermodynamics of hydrogen bond and hydrophobic interactions in cyclodextrin complexes. Biophys. J. 1996, 71, 2144–2154. [Google Scholar] [CrossRef] [Green Version]
  51. Loftsson, T.; Björnsdóttir, S.; Pálsdóttir, G.; Bodor, N. The effects of 2-hydroxypropyl-β-cyclodextrin on the solubility and stability of chlorambucil and melphalan in aqueous solution. Int. J. Pharm. 1989, 57, 63–72. [Google Scholar] [CrossRef]
  52. Pitha, J.; Milecki, J.; Fales, H.; Pannell, L.; Uekama, K. Hydroxypropyl-β-cyclodextrin: Preparation and characterization; effects on solubility of drugs. Int. J. Pharm. 1986, 29, 73–82. [Google Scholar] [CrossRef]
  53. Loftsson, T. Cyclodextrins and the Biopharmaceutics Classification System of Drugs. J. Incl. Phenom. 2002, 44, 63–67. [Google Scholar] [CrossRef]
  54. Chow, D.D.; Karara, A.H. Characterization, dissolution and bioavailability in rats of ibuprofen-β-cyclodextrin complex system. Int. J. Pharm. 1986, 28, 95–101. [Google Scholar] [CrossRef]
  55. Vila-Jato, J.L.; Blanco, J.; Torres, J.J. Biopharmaceutical aspects of the tolbutamide-beta-cyclodextrin inclusion compound. Farm. Prat. 1988, 43, 37–45. [Google Scholar]
  56. Wang, Z.; Landy, D.; Sizun, C.; Cézard, C.; Solgadi, A.; Przybylski, C.; de Chaisemartin, L.; Herfindal, L.; Barratt, G.; Legrand, F.-X. Cyclodextrin complexation studies as the first step for repurposing of chlorpromazine. Int. J. Pharm. 2020, 584, 119391. [Google Scholar] [CrossRef]
  57. Kfoury, M.; Landy, D.; Fourmentin, S. Characterization of Cyclodextrin/Volatile Inclusion Complexes: A Review. Molecules 2018, 23, 1204. [Google Scholar] [CrossRef] [Green Version]
  58. Takahashi, K. Organic Reactions Mediated by Cyclodextrins. Chem. Rev. 1998, 98, 2013–2034. [Google Scholar] [CrossRef] [PubMed]
  59. Rama, A.C.R.; Veiga, F.; Figueiredo, I.V.; Sousa, A.; Caramona, M. Aspectos biofarmacêuticos da formulação de medicamentos para neonatos: Fundamentos da complexação de indometacina com hidroxipropil-beta-ciclodextrina para tratamento oral do fechamento do canal arterial. Revista Brasileira de Ciências Farmacêuticas 2005, 41, 281–299. [Google Scholar] [CrossRef]
  60. Charumanee, S.; Titwan, A.; Sirithunyalug, J.; Weiss-Greiler, P.; Wolschann, P.; Viernstein, H.; Okonogi, S. Thermodynamics of the encapsulation by cyclodextrins. J. Chem. Technol. Biotechnol. 2006, 81, 523–529. [Google Scholar] [CrossRef]
  61. Ameen, H.M.; Kunsági-Máté, S.; Bognár, B.; Szente, L.; Poór, M.; Lemli, B. Thermodynamic Characterization of the Interaction between the Antimicrobial Drug Sulfamethazine and Two Selected Cyclodextrins. Molecules 2019, 24, 4565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Vončina, B.; Vivod, V.; Jaušovec, D. β-Cyclodextrin as retarding reagent in polyacrylonitrile dyeing. Dye. Pigment. 2007, 74, 642–646. [Google Scholar] [CrossRef]
  63. Hedges, A.R. Industrial Applications of Cyclodextrins. Chem. Rev. 1998, 98, 2035–2044. [Google Scholar] [CrossRef]
  64. Cirri, M.; Maestrelli, F.; Orlandini, S.; Furlanetto, S.; Pinzauti, S.; Mura, P. Determination of stability constant values of flurbiprofen-cyclodextrin complexes using different techniques. J. Pharm. Biomed. Anal. 2005, 37, 995–1002. [Google Scholar] [CrossRef]
  65. Vozone, C.M.; Marques, H.M.C. Complexation of Budesonide in Cyclodextrins and Particle Aerodynamic Characterization of the Complex Solid Form for Dry Powder Inhalation. J. Incl. Phenom. 2002, 44, 111–116. [Google Scholar] [CrossRef]
  66. Cao, F.; Guo, J.; Ping, Q. The physicochemical characteristics of freeze-dried scutellarin-cyclodextrin tetracomponent complexes. Drug Dev. Ind. Pharm. 2005, 31, 747–756. [Google Scholar] [CrossRef]
  67. Miro, A.; Ungaro, F.; Quaglia, F. Cyclodextrins as Smart Excipients in Polymeric Drug Delivery Systems. In Cyclodextrins in Pharmaceutics, Cosmetics, and Biomedicine; John Wiley & Sons, Ltd.: New Jersey, NY, USA, 2011; pp. 65–89. [Google Scholar]
  68. Buschmann, H.-J.; Denter, U.; Knittel, D.; Schollmeyer, E. The Use of Cyclodextrins in Textile Processes—An Overview. J. Text. Inst. 1998, 89, 554–561. [Google Scholar] [CrossRef]
  69. Loftsson, T.; Brewster, M.E. Pharmaceutical Applications of Cyclodextrins. 1. Drug Solubilization and Stabilization. J. Pharm. Sci. 1996, 85, 1017–1025. [Google Scholar] [CrossRef] [PubMed]
  70. Alzate-Sánchez, D.M.; Smith, B.J.; Alsbaiee, A.; Hinestroza, J.P.; Dichtel, W.R. Cotton Fabric Functionalized with a β-Cyclodextrin Polymer Captures Organic Pollutants from Contaminated Air and Water. Chem. Mater. 2016, 28, 8340–8346. [Google Scholar] [CrossRef]
  71. Tonelli, A.E. Cyclodextrins as a means to nanostructure and functionalize polymers. J. Incl. Phenom. Macrocycl. Chem. 2008, 60, 197–202. [Google Scholar] [CrossRef]
  72. Hodul, P.; Duris, M.; Králik, M. Inclusion complexes of B-cyclodextrin with non-ionic surfactants in textile preparation process. Vlakna a Text. 1996, 3, 16–19. [Google Scholar]
  73. Carpignano, R.; Parlati, S.; Piccinini, P.; Savarino, P.; Giorgi, M.R.D.; Fochi, R. Use of β-cyclodextrin in the dyeing of polyester with low environmental impact. Color. Technol. 2010, 126, 201–208. [Google Scholar] [CrossRef]
  74. Sricharussin, W.; Sopajaree, C.; Maneerung, T.; Sangsuriya, N. Modification of cotton fabrics with β-cyclodextrin derivative for aroma finishing. J. Text. Inst. 2009, 100, 682–687. [Google Scholar] [CrossRef]
  75. Raslan, W.M.; El-Aref, A.T.; Bendak, A. Modification of cellulose acetate fabric with cyclodextrin to improve its dyeability. J. Appl. Polym. Sci. 2009, 112, 3192–3198. [Google Scholar] [CrossRef]
  76. Scacchetti, F.A.P.; Pinto, E.; Soares, G.M.B. Functionalization and characterization of cotton with phase change materials and thyme oil encapsulated in beta-cyclodextrins. Prog. Org. Coatings 2017, 107, 64–74. [Google Scholar] [CrossRef] [Green Version]
  77. Mihailiasa, M.; Caldera, F.; Li, J.; Peila, R.; Ferri, A.; Trotta, F. Preparation of functionalized cotton fabrics by means of melatonin loaded β-cyclodextrin nanosponges. Carbohydr. Polym. 2016, 142, 24–30. [Google Scholar] [CrossRef]
  78. Bezerra, F.M.; Carmona, O.G.; Carmona, C.G.; Plath, A.S.; Lis, M. Biofunctional wool using β-cyclodextrins as vehiculizer of citronella oil. Process. Biochem. 2019, 77, 151–158. [Google Scholar] [CrossRef]
  79. Chang, Y.; Dou, N.; Liu, M.; Jiang, M.; Men, J.; Cui, Y.; Li, R.; Zhu, Y. Efficient removal of anionic dyes from aqueous solution using CTAB and β-cyclodextrin-induced dye aggregation. J. Mol. Liq. 2020, 309, 113021. [Google Scholar] [CrossRef]
  80. Keskin, N.O.S.; Celebioglu, A.; Sarioglu, O.F.; Uyar, T.; Tekinay, T. Encapsulation of living bacteria in electrospun cyclodextrin ultrathin fibers for bioremediation of heavy metals and reactive dye from wastewater. Colloids Surf. B Biointerfaces 2018, 161, 169–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Crini, G. Studies on adsorption of dyes on beta-cyclodextrin polymer. Bioresour. Technol. 2003, 90, 193–198. [Google Scholar] [CrossRef]
  82. Cireli, A.; Yurdakul, B. Application of cyclodextrin to the textile dyeing and washing processes. J. Appl. Polym. Sci. 2006, 100, 208–218. [Google Scholar] [CrossRef]
  83. Park, J.S.; Kim, I.-S. Use of β-cyclodextrin in an antimigration coating for polyester fabric. Color. Technol. 2013, 129, 347–351. [Google Scholar] [CrossRef]
  84. Kacem, I.; Laurent, T.; Blanchemain, N.; Neut, C.; Chai, F.; Haulon, S.; Hildebrand, H.F.; Martel, B. Dyeing and antibacterial activation with methylene blue of a cyclodextrin modified polyester vascular graft. J. Biomed. Mater. Res. Part A 2014, 102, 2942–2951. [Google Scholar] [CrossRef]
  85. Dardeer, H.M.; El-sisi, A.A.; Emam, A.A.; Hilal, N.M. Synthesis, Application of a Novel Azo Dye and Its Inclusion Complex with Beta-cyclodextrin onto Polyester Fabric. Int. J. Text. Sci. 2017, 6, 79–87. [Google Scholar]
  86. Dutra, F.V.; Santos, K.R.M.D.; Bezerra, F.M. Complexação de corantes dispersos utilizando ß-ciclodextrina em tricromia para poliéster. Química Têxt. 2019, 43, 40–49. [Google Scholar]
  87. Ferreira, B.T.M.; Espinoza-Quiñones, F.R.; Borba, C.E.; Módenes, A.N.; Santos, W.L.F.; Bezerra, F.M. Use of the β-Cyclodextrin Additive as a Good Alternative for the Substitution of Environmentally Harmful Additives in Industrial Dyeing Processes. Fibers Polym. 2020, 21, 1266–1274. [Google Scholar] [CrossRef]
  88. Savarino, P.; Piccinini, P.; Montoneri, E.; Viscardi, G.; Quagliotto, P.; Barni, E. Effects of additives on the dyeing of nylon-6 with dyes containing hydrophobic and hydrophilic moieties. Dye. Pigment. 2000, 47, 177–188. [Google Scholar] [CrossRef]
  89. Savarino, P.; Parlati, S.; Buscaino, R.; Piccinini, P.; Degani, I.; Barni, E. Effects of additives on the dyeing of polyamide fibres. Part I: β-cyclodextrin. Dye. Pigment. 2004, 60, 223–232. [Google Scholar] [CrossRef]
  90. Savarino, P.; Parlati, S.; Buscaino, R.; Piccinini, P.; Barolo, C.; Montoneri, E. Effects of additives on the dyeing of polyamide fibres. Part II: Methyl-β-cyclodextrin. Dye. Pigment. 2006, 69, 7–12. [Google Scholar] [CrossRef]
  91. Parlati, S.; Gobetto, R.; Barolo, C.; Arrais, A.; Buscaino, R.; Medana, C.; Savarino, P. Preparation and application of a β-cyclodextrin-disperse/reactive dye complex. J. Incl. Phenom. Macrocycl. Chem. 2007, 57, 463–470. [Google Scholar] [CrossRef]
  92. Shibusawa, T.; Okamoto, J.; Abe, K.; Sakata, K.; Ito, Y. Inclusion of azo disperse dyes by cyclodextrins at dyeing temperature. Dye. Pigment. 1998, 36, 79–91. [Google Scholar] [CrossRef]
  93. Dutra, F.V.; Caruzi, B.B.; Kawasaki, I.; Silva, T.L.; Bezerra, F.M. Utilização de β-ciclodextrina no tingimento de lã com Extratode Urucum (BiXa orellana). Química Têxt. 2014, 37, 58–64. [Google Scholar]
  94. Chen, L.; Wang, C.; Tian, A.; Wu, M. An attempt of improving polyester inkjet printing performance by surface modification using β-cyclodextrin. Surf. Interface Anal. 2012, 44, 1324–1330. [Google Scholar] [CrossRef]
  95. Lu, M.; Liu, Y.P. Dyeing Kinetics of Vinylon Modified with β-Cyclodextrin. Fibres Text. East. Eur. 2011, 5, 88. [Google Scholar]
  96. Rehan, M.; Mahmoud, S.A.; Mashaly, H.M.; Youssef, B.M. β-Cyclodextrin assisted simultaneous preparation and dyeing acid dyes onto cotton fabric. React. Funct. Polym. 2020, 151, 104573. [Google Scholar] [CrossRef]
  97. Zhang, W.; Ji, X.; Wang, C.; Yin, Y. One-bath one-step low-temperature dyeing of polyester/cotton blended fabric with cationic dyes via β-cyclodextrin modification. Text. Res. J. 2019, 89, 1699–1711. [Google Scholar] [CrossRef]
  98. Ibrahim, N.A.; El-Zairy, E.M.R. Union disperse printing and UV-protecting of wool/polyester blend using a reactive β-cyclodextrin. Carbohydr. Polym. 2009, 76, 244–249. [Google Scholar] [CrossRef]
  99. Ghoul, Y.E.; Martel, B.; Achari, A.E.; Campagne, C.; Razafimahefa, L.; Vroman, I. Improved dyeability of polypropylene fabrics finished with β-cyclodextrin–citric acid polymer. Polym. J. 2010, 42, 804–811. [Google Scholar] [CrossRef] [Green Version]
  100. Burkinshaw, S.M.; Liu, K.; Salihu, G. The wash-off of dyeings using interstitial water Part 5: Residual dyebath and wash-off liquor generated during the application of disperse dyes and reactive dyes to polyester/cotton fabric. Dye. Pigment. 2019, 171, 106367. [Google Scholar] [CrossRef] [Green Version]
  101. Hou, A.; Chen, B.; Dai, J.; Zhang, K. Using supercritical carbon dioxide as solvent to replace water in polyethylene terephthalate (PET) fabric dyeing procedures. J. Clean. Prod. 2010, 18, 1009–1014. [Google Scholar] [CrossRef]
  102. Burkinshaw, S.M.; Lagonika, K. Sulphur dyes on nylon 6,6. Part 3. Preliminary studies of the nature of dye–fibre interaction. Dye. Pigment. 2006, 69, 185–191. [Google Scholar] [CrossRef]
  103. Burkinshaw, S.M.; Son, Y.-A. A comparison of the colour strength and fastness to repeated washing of acid dyes on standard and deep dyeable nylon 6,6. Dye. Pigment. 2006, 70, 156–163. [Google Scholar] [CrossRef]
  104. Xie, K.; Liu, H.; Wang, X. Surface modification of cellulose with triazine derivative to improve printability with reactive dyes. Carbohydr. Polym. 2009, 78, 538–542. [Google Scholar] [CrossRef]
  105. Burkinshaw, S.M.; Mignanelli, M.; Froehling, P.E.; Bide, M.J. The use of dendrimers to modify the dyeing behaviour of reactive dyes on cotton. Dye. Pigment. 2000, 47, 259–267. [Google Scholar] [CrossRef]
  106. Burkinshaw, S.M.; Salihu, G. The role of auxiliaries in the immersion dyeing of textile fibres part 2: Analysis of conventional models that describe the manner by which inorganic electrolytes promote direct dye uptake on cellulosic fibres. Dye. Pigment. 2019, 161, 531–545. [Google Scholar] [CrossRef] [Green Version]
  107. Burkinshaw, S.M.; Salihu, G. The role of auxiliaries in the immersion dyeing of textile fibres: Part 4 theoretical model to describe the role of liquor ratio in dyeing cellulosic fibres with direct dyes in the absence and presence of inorganic electrolyte. Dye. Pigment. 2019, 161, 565–580. [Google Scholar] [CrossRef] [Green Version]
  108. Hunger, K. Industrial Dyes: Chemistry, Properties, Applications; Wiley-VCH: Weinheim, Germany, 2003; ISBN 978-3-527-30426-4. [Google Scholar]
  109. Wang, H.; Lewis, D.M. Chemical modification of cotton to improve fibre dyeability. Color. Technol. 2002, 118, 159–168. [Google Scholar] [CrossRef]
  110. Udrescu, C.; Ferrero, F.; Periolatto, M. Ultrasound-assisted dyeing of cellulose acetate. Ultrason. Sonochem. 2014, 21, 1477–1481. [Google Scholar] [CrossRef] [PubMed]
  111. Martel, B.; Ruffin, D.; Weltrowski, M.; Lekchiri, Y.; Morcellet, M. Water-soluble polymers and gels from the polycondensation between cyclodextrins and poly(carboxylic acid)s: A study of the preparation parameters. J. Appl. Polym. Sci. 2005, 97, 433–442. [Google Scholar] [CrossRef]
  112. Arias, M.J.L.; Coderch, L.; Martí, M.; Alonso, C.; Carmona, O.G.; Carmona, C.G.; Maesta, F. Vehiculation of Active Principles as a Way to Create Smart and Biofunctional Textiles. Materials 2018, 11, 2152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Ciobanu, A.; Mallard, I.; Landy, D.; Brabie, G.; Nistor, D.; Fourmentin, S. Retention of aroma compounds from Mentha piperita essential oil by cyclodextrins and crosslinked cyclodextrin polymers. Food Chem. 2013, 138, 291–297. [Google Scholar] [CrossRef]
  114. Peila, R.; Migliavacca, G.; Aimone, F.; Ferri, A.; Sicardi, S. A comparison of analytical methods for the quantification of a reactive β-cyclodextrin fixed onto cotton yarns. Cellulose 2012, 19, 1097–1105. [Google Scholar] [CrossRef]
  115. Hedayati, N.; Montazer, M.; Mahmoudirad, M.; Toliyat, T. Ketoconazole and Ketoconazole/β-cyclodextrin performance on cotton wound dressing as fungal skin treatment. Carbohydr. Polym. 2020, 240, 116267. [Google Scholar] [CrossRef]
  116. El-Ghoul, Y. Biological and microbiological performance of new polymer-based chitosan and synthesized amino-cyclodextrin finished polypropylene abdominal wall prosthesis biomaterial. Text. Res. J. 2020, 0040517520926624. [Google Scholar] [CrossRef]
  117. McQueen, R.H.; Vaezafshar, S. Odor in textiles: A review of evaluation methods, fabric characteristics, and odor control technologies. Text. Res. J. 2019, 90, 1–17. [Google Scholar] [CrossRef]
  118. Kadam, V.; Kyratzis, I.L.; Truong, Y.B.; Wang, L.; Padhye, R. Air filter media functionalized with β-Cyclodextrin for efficient adsorption of volatile organic compounds. J. Appl. Polym. Sci. 2020, 137, 49228. [Google Scholar] [CrossRef]
  119. Azizi, N.; Ben Abdelkader, M.; Chevalier, Y.; Majdoub, M. New β-Cyclodextrin-Based Microcapsules for Textiles Uses. Fibers Polym. 2019, 20, 683–689. [Google Scholar] [CrossRef]
  120. Wang, C.X.; Chen, S.L. Aromachology and its application in the textile field. Fibres Text. East. Eur. 2005, 13, 41–44. [Google Scholar]
  121. Agrawal, P.B.; Warmoeskerken, M.M.C.G. Permanent fixation of β-cyclodextrin on cotton surface—An assessment between innovative and established approaches. J. Appl. Polym. Sci. 2012, 124, 4090–4097. [Google Scholar] [CrossRef]
  122. Nazi, M.; Malek, R.M.A.; Kotek, R. Modification of β-cyclodextrin with itaconic acid and application of the new derivative to cotton fabrics. Carbohydr. Polym. 2012, 88, 950–958. [Google Scholar] [CrossRef]
  123. Zhang, F.; Islam, M.S.; Berry, R.M.; Tam, K.C. β-Cyclodextrin-Functionalized Cellulose Nanocrystals and Their Interactions with Surfactants. ACS Omega 2019, 4, 2102–2110. [Google Scholar] [CrossRef] [PubMed]
  124. Reuscher, H.; Hirsenkorn, R. BETA W7 MCT—New ways in surface modification. J. Incl. Phenom. Macrocycl. Chem. 1996, 25, 191–196. [Google Scholar] [CrossRef]
  125. Shown, I.; Murthy, C.N. Grafting of cotton fiber by water-soluble cyclodextrin-based polymer. J. Appl. Polym. Sci. 2009, 111, 2056–2061. [Google Scholar] [CrossRef]
  126. Nichifor, M.; Constantin, M.; Mocanu, G.; Fundueanu, G.; Branisteanu, D.; Costuleanu, M.; Radu, C.D. New multifunctional textile biomaterials for the treatment of leg venous insufficiency. J. Mater. Sci. Mater. Med. 2009, 20, 975–982. [Google Scholar] [CrossRef]
  127. Yu, Y.; Wang, Q.; Yuan, J.; Fan, X.; Wang, P. A novel approach for grafting of β-cyclodextrin onto wool via laccase/TEMPO oxidation. Carbohydr. Polym. 2016, 153, 463–470. [Google Scholar] [CrossRef]
  128. Haji, A.; Mehrizi, M.K.; Akbarpour, R. Optimization of β-cyclodextrin grafting on wool fibers improved by plasma treatment and assessment of antibacterial activity of berberine finished fabric. J. Incl. Phenom. Macrocycl. Chem. 2015, 81, 121–133. [Google Scholar] [CrossRef]
  129. Montazer, M.; Jolaei, M.M. β-Cyclodextrin stabilized on three-dimensional polyester fabric with different crosslinking agents. J. Appl. Polym. Sci. 2010, 116, 210–217. [Google Scholar] [CrossRef]
  130. Abdel-Halim, E.S.; Abdel-Mohdy, F.A.; Al-Deyab, S.S.; El-Newehy, M.H. Chitosan and monochlorotriazinyl-β-cyclodextrin finishes improve antistatic properties of cotton/polyester blend and polyester fabrics. Carbohydr. Polym. 2010, 82, 202–208. [Google Scholar] [CrossRef]
  131. Shi, H.; Wang, Y.; Hua, R. Acid-catalyzed carboxylic acid esterification and ester hydrolysis mechanism: Acylium ion as a sharing active intermediate via a spontaneous trimolecular reaction based on density functional theory calculation and supported by electrospray ionization-mass spectrometry. Phys. Chem. Chem. Phys. 2015, 17, 30279–30291. [Google Scholar] [CrossRef]
  132. Radu, C.-D.; Salariu, M.; Avadanei, M.; Ghiciuc, C.; Foia, L.; Lupusoru, E.C.; Ferri, A.; Ulea, E.; Lipsa, F. Cotton-made cellulose support for anti-allergic pajamas. Carbohydr. Polym. 2013, 95, 479–486. [Google Scholar] [CrossRef] [PubMed]
  133. Khanna, S.; Chakraborty, J.N. Optimization of monochlorotriazine β-cyclodextrin grafting on cotton and assessment of release behavior of essential oils from functionalized fabric. Fash. Text. 2017, 4, 6. [Google Scholar] [CrossRef] [Green Version]
  134. Ibrahim, N.A.; Abdalla, W.A.; El-Zairy, E.M.R.; Khalil, H.M. Utilization of monochloro-triazine β-cyclodextrin for enhancing printability and functionality of wool. Carbohydr. Polym. 2013, 92, 1520–1529. [Google Scholar] [CrossRef] [PubMed]
  135. Martel, B.; Weltrowski, M.; Ruffin, D.; Morcellet, M. Polycarboxylic acids as crosslinking agents for grafting cyclodextrins onto cotton and wool fabrics: Study of the process parameters. J. Appl. Polym. Sci. 2002, 83, 1449–1456. [Google Scholar] [CrossRef]
  136. Liu, S. Bio-Functional Textiles. In Handbook of Medical Textiles; Textiles; Woodhead Publishing: Cambridge, UK, 2011; pp. 336–359. ISBN 978-0-85709-369-1. [Google Scholar]
  137. Blanchemain, N.; Karrout, Y.; Tabary, N.; Neut, C.; Bria, M.; Siepmann, J.; Hildebrand, H.F.; Martel, B. Methyl-β-cyclodextrin modified vascular prosthesis: Influence of the modification level on the drug delivery properties in different media. Acta Biomater. 2011, 7, 304–314. [Google Scholar] [CrossRef]
  138. Abdel-Halim, E.S.; Al-Deyab, S.S.; Alfaifi, A.Y.A. Cotton fabric finished with β-cyclodextrin: Inclusion ability toward antimicrobial agent. Carbohydr. Polym. 2014, 102, 550–556. [Google Scholar] [CrossRef]
  139. Bajpai, M.; Gupta, P.; Bajpai, S.K. Silver (I) ions loaded cyclodextrin-grafted-cotton fabric with excellent antimicrobial property. Fibers Polym. 2010, 11, 8–13. [Google Scholar] [CrossRef]
  140. Hebeish, A.; El-Shafei, A.; Sharaf, S.; Zaghloul, S. In situ formation of silver nanoparticles for multifunctional cotton containing cyclodextrin. Carbohydr. Polym. 2014, 103, 442–447. [Google Scholar] [CrossRef]
  141. Rukmani, A.; Sundrarajan, M. Inclusion of antibacterial agent thymol on β-cyclodextrin-grafted organic cotton. J. Ind. Text. 2011, 42, 132–144. [Google Scholar] [CrossRef]
  142. Selvam, S.; Gandhi, R.R.; Suresh, J.; Gowri, S.; Ravikumar, S.; Sundrarajan, M. Antibacterial effect of novel synthesized sulfated β-cyclodextrin crosslinked cotton fabric and its improved antibacterial activities with ZnO, TiO2 and Ag nanoparticles coating. Int. J. Pharm. 2012, 434, 366–374. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, J.; Cai, Z. Incorporation of the antibacterial agent, miconazole nitrate into a cellulosic fabric grafted with β-cyclodextrin. Carbohydr. Polym. 2008, 72, 695–700. [Google Scholar] [CrossRef]
  144. Cabrales, L.; Abidi, N.; Hammond, A.; Hamood, A. Cotton Fabric Functionalization with Cyclodextrins. Surfaces 2012, 6, 14. [Google Scholar]
  145. Khanna, S.; Sharma, S.; Chakraborty, J.N. Performance assessment of fragrance finished cotton with cyclodextrin assisted anchoring hosts. Fash. Text. 2015, 2, 19. [Google Scholar] [CrossRef] [Green Version]
  146. Abdel-Mohdy, F.A.; Fouda, M.M.G.; Rehan, M.F.; Aly, A.S. Repellency of controlled-release treated cotton fabrics based on cypermethrin and prallethrin. Carbohydr. Polym. 2008, 73, 92–97. [Google Scholar] [CrossRef]
  147. Radu, C.D.; Parteni, O.; Popa, M.; Muresan, I.E.; Ochiuz, L.; Bulgariu, L.; Munteanu, C.; Istrate, B.; Ulea, E. Comparative Study of a Drug Release from a Textile to Skin. J. Pharm. Drug Deliv. Res. 2016, 2015, 1–8. [Google Scholar] [CrossRef]
  148. El-Ghoul, Y.; Blanchemain, N.; Laurent, T.; Campagne, C.; El Achari, A.; Roudesli, S.; Morcellet, M.; Martel, B.; Hildebrand, H.F. Chemical, biological and microbiological evaluation of cyclodextrin finished polyamide inguinal meshes. Acta Biomater. 2008, 4, 1392–1400. [Google Scholar] [CrossRef]
  149. Scalia, S.; Tursilli, R.; Bianchi, A.; Nostro, P.L.; Bocci, E.; Ridi, F.; Baglioni, P. Incorporation of the sunscreen agent, octyl methoxycinnamate in a cellulosic fabric grafted with β-cyclodextrin. Int. J. Pharm. 2006, 308, 155–159. [Google Scholar] [CrossRef]
  150. Blanchemain, N.; Karrout, Y.; Tabary, N.; Bria, M.; Neut, C.; Hildebrand, H.F.; Siepmann, J.; Martel, B. Comparative study of vascular prostheses coated with polycyclodextrins for controlled ciprofloxacin release. Carbohydr. Polym. 2012, 90, 1695–1703. [Google Scholar] [CrossRef]
  151. Shlar, I.; Droby, S.; Rodov, V. Antimicrobial coatings on polyethylene terephthalate based on curcumin/cyclodextrin complex embedded in a multilayer polyelectrolyte architecture. Colloids Surf. B Biointerfaces 2018, 164, 379–387. [Google Scholar] [CrossRef] [PubMed]
  152. Martin, A.; Tabary, N.; Leclercq, L.; Junthip, J.; Degoutin, S.; Aubert-Viard, F.; Cazaux, F.; Lyskawa, J.; Janus, L.; Bria, M.; et al. Multilayered textile coating based on a β-cyclodextrin polyelectrolyte for the controlled release of drugs. Carbohydr. Polym. 2013, 93, 718–730. [Google Scholar] [CrossRef] [PubMed]
  153. Martel, B.; Morcellet, M.; Ruffin, D.; Vinet, F.; Weltrowski, L. Capture and Controlled Release of Fragrances by CD Finished Textiles. J. Incl. Phenom. 2002, 44, 439–442. [Google Scholar] [CrossRef]
  154. Castriciano, M.A.; Zagami, R.; Casaletto, M.P.; Martel, B.; Trapani, M.; Romeo, A.; Villari, V.; Sciortino, M.T.; Grasso, L.; Guglielmino, S.; et al. Poly(carboxylic acid)-Cyclodextrin/Anionic Porphyrin Finished Fabrics as Photosensitizer Releasers for Antimicrobial Photodynamic Therapy. Biomacromolecules 2017, 18, 1134–1144. [Google Scholar] [CrossRef]
  155. Yao, T.; Wang, J.; Xue, Y.; Yu, W.; Gao, Q.; Ferreira, L.; Ren, K.-F.; Ji, J. A photodynamic antibacterial spray-coating based on the host–guest immobilization of the photosensitizer methylene blue. J. Mater. Chem. B 2019, 7, 5089–5095. [Google Scholar] [CrossRef]
  156. Jiang, H.-L.; Xu, M.-Y.; Xie, Z.-W.; Hai, W.; Xie, X.-L.; He, F.-A. Selective adsorption of anionic dyes from aqueous solution by a novel β-cyclodextrin-based polymer. J. Mol. Struct. 2020, 1203, 127373. [Google Scholar] [CrossRef]
  157. Chen, J.; Liu, M.; Pu, Y.; Wang, C.; Han, J.; Jiang, M.; Liu, K. The preparation of thin-walled multi-cavities β-cyclodextrin polymer and its static and dynamic properties for dyes removal. J. Environ. Manag. 2019, 245, 105–113. [Google Scholar] [CrossRef]
  158. Afkhami, A.; Saber-Tehrani, M.; Bagheri, H. Modified maghemite nanoparticles as an efficient adsorbent for removing some cationic dyes from aqueous solution. Desalination 2010, 263, 240–248. [Google Scholar] [CrossRef]
  159. Zare, K.; Gupta, V.K.; Moradi, O.; Makhlouf, A.S.H.; Sillanpää, M.; Nadagouda, M.N.; Sadegh, H.; Shahryari-ghoshekandi, R.; Pal, A.; Wang, Z.; et al. A comparative study on the basis of adsorption capacity between CNTs and activated carbon as adsorbents for removal of noxious synthetic dyes: A review. J. Nanostruct. Chem. 2015, 5, 227–236. [Google Scholar] [CrossRef] [Green Version]
  160. Sansuk, S.; Srijaranai, S.; Srijaranai, S. A New Approach for Removing Anionic Organic Dyes from Wastewater Based on Electrostatically Driven Assembly. Environ. Sci. Technol. 2016, 50, 6477–6484. [Google Scholar] [CrossRef]
  161. Debnath, S.; Ballav, N.; Maity, A.; Pillay, K. Competitive adsorption of ternary dye mixture using pine cone powder modified with β-cyclodextrin. J. Mol. Liq. 2017, 225, 679–688. [Google Scholar] [CrossRef]
  162. Abd-Elhamid, A.I.; El-Aassar, M.R.; El Fawal, G.F.; Soliman, H.M.A. Fabrication of polyacrylonitrile/β-cyclodextrin/graphene oxide nanofibers composite as an efficient adsorbent for cationic dye. Environ. NanoTechnol. Monit. Manag. 2019, 11, 100207. [Google Scholar] [CrossRef]
  163. Batista, L.M.B.; dos Santos, A.J.; da Silva, D.R.; de Melo Alves, A.P.; Garcia-Segura, S.; Martínez-Huitle, C.A. Solar photocatalytic application of NbO2OH as alternative photocatalyst for water treatment. Sci. Total Environ. 2017, 596–597, 79–86. [Google Scholar] [CrossRef] [PubMed]
  164. Basha, C.A.; Selvakumar, K.V.; Prabhu, H.J.; Sivashanmugam, P.; Lee, C.W. Degradation studies for textile reactive dye by combined electrochemical, microbial and photocatalytic methods. Sep. Purif. Technol. 2011, 79, 303–309. [Google Scholar] [CrossRef]
  165. Chollom, M.N.; Rathilal, S.; Pillay, V.L.; Alfa, D. The applicability of nanofiltration for the treatment and reuse of textile reactive dye effluent. Water SA 2015, 41, 398–405. [Google Scholar] [CrossRef] [Green Version]
  166. Yeap, K.L.; Teng, T.T.; Poh, B.T.; Morad, N.; Lee, K.E. Preparation and characterization of coagulation/flocculation behavior of a novel inorganic–organic hybrid polymer for reactive and disperse dyes removal. Chem. Eng. J. 2014, 243, 305–314. [Google Scholar] [CrossRef]
  167. Gosavi, V.D.; Sharma, S. A General Review on Various Treatment Methods for Textile Wastewater. Available online: /paper/A-General-Review-on-Various-Treatment-Methods-for-Gosavi-Sharma/e8d649881280265b4bf146b25f6e5eb2ddb99afb (accessed on 14 June 2020).
  168. Ahmed, M.B.; Zhou, J.L.; Ngo, H.H.; Guo, W.; Thomaidis, N.S.; Xu, J. Progress in the biological and chemical treatment technologies for emerging contaminant removal from wastewater: A critical review. J. Hazard. Mater. 2017, 323, 274–298. [Google Scholar] [CrossRef]
  169. Lin, Q.; Gao, M.; Chang, J.; Ma, H. Adsorption properties of crosslinking carboxymethyl cellulose grafting dimethyldiallylammonium chloride for cationic and anionic dyes. Carbohydr. Polym. 2016, 151, 283–294. [Google Scholar] [CrossRef]
  170. Crini, G. Kinetic and equilibrium studies on the removal of cationic dyes from aqueous solution by adsorption onto a cyclodextrin polymer. Dye. Pigment. 2008, 77, 415–426. [Google Scholar] [CrossRef]
  171. Pellicer, J.A.; Rodríguez-López, M.I.; Fortea, M.I.; Lucas-Abellán, C.; Mercader-Ros, M.T.; López-Miranda, S.; Gómez-López, V.M.; Semeraro, P.; Cosma, P.; Fini, P.; et al. Adsorption Properties of β- and Hydroxypropyl-β-Cyclodextrins Cross-Linked with Epichlorohydrin in Aqueous Solution. A Sustainable Recycling Strategy in Textile Dyeing Process. Polymers 2019, 11, 252. [Google Scholar] [CrossRef] [Green Version]
  172. Crini, G. Recent developments in polysaccharide-based materials used as adsorbents in wastewater treatment. Prog. Polym. Sci. 2005, 30, 38–70. [Google Scholar] [CrossRef]
  173. Eli, W.; Chen, W.; Xue, Q. The association of anionic surfactants with β -cyclodextrin. An isothermal titration calorimeter study. J. Chem. Thermodyn. 1999, 31, 1283–1296. [Google Scholar] [CrossRef]
  174. Zhao, R.; Wang, Y.; Li, X.; Sun, B.; Jiang, Z.; Wang, C. Water-insoluble sericin/β-cyclodextrin/PVA composite electrospun nanofibers as effective adsorbents towards methylene blue. Colloids Surf. B Biointerfaces 2015, 136, 375–382. [Google Scholar] [CrossRef] [PubMed]
  175. Yilmaz, A.; Yilmaz, E.; Yilmaz, M.; Bartsch, R.A. Removal of azo dyes from aqueous solutions using calix[4]arene and β-cyclodextrin. Dye. Pigment. 2007, 74, 54–59. [Google Scholar] [CrossRef]
  176. Liu, J.; Liu, G.; Liu, W. Preparation of water-soluble β-cyclodextrin/poly(acrylic acid)/graphene oxide nanocomposites as new adsorbents to remove cationic dyes from aqueous solutions. Chem. Eng. J. 2014, 257, 299–308. [Google Scholar] [CrossRef]
  177. Mohammadi, A.; Veisi, P. High adsorption performance of β-cyclodextrin-functionalized multi-walled carbon nanotubes for the removal of organic dyes from water and industrial wastewater. J. Environ. Chem. Eng. 2018, 6, 4634–4643. [Google Scholar] [CrossRef]
  178. Kyzas, G.Z.; Lazaridis, N.K.; Bikiaris, D.N. Optimization of chitosan and β-cyclodextrin molecularly imprinted polymer synthesis for dye adsorption. Carbohydr. Polym. 2013, 91, 198–208. [Google Scholar] [CrossRef]
  179. Ozmen, E.Y.; Sezgin, M.; Yilmaz, A.; Yilmaz, M. Synthesis of β-cyclodextrin and starch based polymers for sorption of azo dyes from aqueous solutions. Bioresour. Technol. 2008, 99, 526–531. [Google Scholar] [CrossRef]
  180. Yilmaz, E.; Memon, S.; Yilmaz, M. Removal of direct azo dyes and aromatic amines from aqueous solutions using two β-cyclodextrin-based polymers. J. Hazard. Mater. 2010, 174, 592–597. [Google Scholar] [CrossRef]
  181. Li, X.; Zhou, M.; Jia, J.; Jia, Q. A water-insoluble viologen-based β-cyclodextrin polymer for selective adsorption toward anionic dyes. React. Funct. Polym. 2018, 126, 20–26. [Google Scholar] [CrossRef]
  182. Kekes, T.; Tzia, C. Adsorption of indigo carmine on functional chitosan and β-cyclodextrin/chitosan beads: Equilibrium, kinetics and mechanism studies. J. Environ. Manag. 2020, 262, 110372. [Google Scholar] [CrossRef] [PubMed]
  183. Chen, P.-Y.; Tung, S.-H. One-Step Electrospinning to Produce Nonsolvent-Induced Macroporous Fibers with Ultrahigh Oil Adsorption Capability. Macromolecules 2017, 50, 2528–2534. [Google Scholar] [CrossRef]
  184. Akinyeye, O.J.; Ibigbami, T.B.; Odeja, O. Effect of Chitosan Powder Prepared from Snail Shells to Remove Lead (II) Ion and Nickel (II) Ion from Aqueous Solution and Its Adsorption Isotherm Model. Am. J. Appl. Chem. 2016, 4, 146. [Google Scholar] [CrossRef] [Green Version]
  185. Chatterjee, S.; Lee, D.S.; Lee, M.W.; Woo, S.H. Nitrate removal from aqueous solutions by cross-linked chitosan beads conditioned with sodium bisulfate. J. Hazard. Mater. 2009, 166, 508–513. [Google Scholar] [CrossRef] [PubMed]
  186. Cestari, A.R.; Vieira, E.F.S.; Mota, J.A. The removal of an anionic red dye from aqueous solutions using chitosan beads—The role of experimental factors on adsorption using a full factorial design. J. Hazard. Mater. 2008, 160, 337–343. [Google Scholar] [CrossRef] [PubMed]
  187. Zhao, J.; Zou, Z.; Ren, R.; Sui, X.; Mao, Z.; Xu, H.; Zhong, Y.; Zhang, L.; Wang, B. Chitosan adsorbent reinforced with citric acid modified β-cyclodextrin for highly efficient removal of dyes from reactive dyeing effluents. Eur. Polym. J. 2018, 108, 212–218. [Google Scholar] [CrossRef]
  188. Krause, R.W.M.; Mamba, B.B.; Bambo, F.M.; Malefetse, T.J. Cyclodextrin polymers: Synthesis and Application in Water Treatment. In Cyclodextrins: Chemistry and Physics; Transworld Research Network: Kerala, India, 2010; ISBN 978-81-7895-430-1. [Google Scholar]
  189. Vahedi, S.; Tavakoli, O.; Khoobi, M.; Ansari, A.; Faramarzi, M.A. Application of novel magnetic β-cyclodextrin-anhydride polymer nano-adsorbent in cationic dye removal from aqueous solution. J. Taiwan Inst. Chem. Eng. 2017, 80, 452–463. [Google Scholar] [CrossRef]
  190. Massaro, M.; Colletti, C.G.; Lazzara, G.; Guernelli, S.; Noto, R.; Riela, S. Synthesis and Characterization of Halloysite–Cyclodextrin Nanosponges for Enhanced Dyes Adsorption. ACS Sustain. Chem. Eng. 2017, 5, 3346–3352. [Google Scholar] [CrossRef] [Green Version]
  191. Ozmen, E.Y.; Yilmaz, M. Use of β-cyclodextrin and starch based polymers for sorption of Congo red from aqueous solutions. J. Hazard. Mater. 2007, 148, 303–310. [Google Scholar] [CrossRef]
  192. Jiang, H.-L.; Lin, J.-C.; Hai, W.; Tan, H.-W.; Luo, Y.-W.; Xie, X.-L.; Cao, Y.; He, F.-A. A novel crosslinked β-cyclodextrin-based polymer for removing methylene blue from water with high efficiency. Colloids Surf. A Physicochem. Eng. Asp. 2019, 560, 59–68. [Google Scholar] [CrossRef]
  193. Zhao, F.; Repo, E.; Yin, D.; Meng, Y.; Jafari, S.; Sillanpää, M. EDTA-Cross-Linked β-Cyclodextrin: An Environmentally Friendly Bifunctional Adsorbent for Simultaneous Adsorption of Metals and Cationic Dyes. Environ. Sci. Technol. 2015, 49, 10570–10580. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Number of publications available from SCOPUS when cyclodextrin (CD); textile (TE); dyeing (DY); textile finishing (FI); and textile wastewater (WA) are selected as keywords.
Figure 1. Number of publications available from SCOPUS when cyclodextrin (CD); textile (TE); dyeing (DY); textile finishing (FI); and textile wastewater (WA) are selected as keywords.
Molecules 25 03624 g001
Figure 2. Complexation system. (a) Inclusion of water molecules in the cyclodextrin cavity; (b) complexation mechanism of the guest molecule in aqueous medium.
Figure 2. Complexation system. (a) Inclusion of water molecules in the cyclodextrin cavity; (b) complexation mechanism of the guest molecule in aqueous medium.
Molecules 25 03624 g002
Figure 3. Use of cyclodextrins in dyeing.
Figure 3. Use of cyclodextrins in dyeing.
Molecules 25 03624 g003
Figure 4. Use of cyclodextrins in textile finishing.
Figure 4. Use of cyclodextrins in textile finishing.
Molecules 25 03624 g004
Figure 5. Modification of β-CD by: (a) cyanuric acid; (b) itaconic anhydride (IAnh); (c) acryloyl chloride and (d) laccase/2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) enzyme.
Figure 5. Modification of β-CD by: (a) cyanuric acid; (b) itaconic anhydride (IAnh); (c) acryloyl chloride and (d) laccase/2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) enzyme.
Molecules 25 03624 g005
Figure 6. Nucleophilic substitution reaction of MCT-β-CD with cellulose.
Figure 6. Nucleophilic substitution reaction of MCT-β-CD with cellulose.
Molecules 25 03624 g006
Figure 7. Direct connection of the β-CD to the textile fiber via crosslinking.
Figure 7. Direct connection of the β-CD to the textile fiber via crosslinking.
Molecules 25 03624 g007
Figure 8. Functionalization of wool fibers with β-CD after oxidation.
Figure 8. Functionalization of wool fibers with β-CD after oxidation.
Molecules 25 03624 g008
Figure 9. Finishing textiles with photodynamic potential.
Figure 9. Finishing textiles with photodynamic potential.
Molecules 25 03624 g009
Figure 10. The role of β-CD as a dye adsorbent.
Figure 10. The role of β-CD as a dye adsorbent.
Molecules 25 03624 g010
Table 1. Some physicochemical properties of cyclodextrins [2,21].
Table 1. Some physicochemical properties of cyclodextrins [2,21].
Propertiesα-CDβ-CDγ-CD
Empirical formulaC36H60O30C42H70O35C48H80O40
Molecular weight (g/mol)97211351297
Glucopyranose units678
Cavity diameter (nm)0.47–0.570.60–0.780.83–0.95
Internal cavity volume (nm3)174026204720
Number of water molecules in the cavity61117
Aqueous solubility (g/L)129.518.4249.2
Temperature of degradation (°C)278298267
Table 2. Examples of dyeing CD applications, as a textile aid and as a chemical modifier.
Table 2. Examples of dyeing CD applications, as a textile aid and as a chemical modifier.
Application of CyclodextrinsFiberDyeReference
Auxiliary agentPolyesterDisperse[73]
Synthetic [85]
Disperse Orange 30, Disperse Red 167, Disperse Blue 79[86]
Methylene Blue[84]
Polyamide 6Disperse Red 60[87]
Disperse[88]
Synthetic[32,89,90]
Nylon, polyester and cotton Synthetic, reactive and disperse dye[91]
Cellulose AcetateAzo disperse[92]
PolyacrylicBasic Blue 4[62]
CottonDirect[82]
WoolNatural (Bixa orellana)[93]
Chemical modificationPolyesterPigment inks carbon black, magenta, yellow and cyan[94]
Disperse Red 60, Disperse Yellow, Disperse Blue 56, Disperse Red 343[83]
Cellulose AcetateDisperse Red 60 and 82[75]
Vinylon fibreReactive Red 2[95]
CottonAcid[96]
Cotton and cotton/polyesterBasic Red 14, Basic Blue 3, Basic Yellow 24 and 13[97]
Polyester/WoolDisperse Red 54 and 167, Disperse blue 183[98]
PolypropyleneDisperse, acid and reactive[99]
Table 3. Studies that used cyclodextrin to graft finishes in textiles.
Table 3. Studies that used cyclodextrin to graft finishes in textiles.
FiberEffectActive MoleculeReference
CottonAntimicrobialOctenidine dihydrochloride[138]
Silver[139,140]
Phenolic compounds[76,141]
Ketoconazole[115]
ZnO, TiO2 and Ag nanoparticles[142]
Miconazole nitrate[143]
Triclosan[144]
Fragrance, antimicrobialEssential Oils[74,145]
Insect repellentCypermethrin and Prallethrin[146]
Nocturnal regulation of sleep and antioxidant propertiesMelatonin[77]
For coetaneous affectionsHydrocortisone acetate[147]
PolyamidePerfume, moisturize and UV-protect.2-ethoxynaphtalene (neroline)[119]
AntibioticsCiprofloxacin[148]
TencelSunscreenOctyl methoxycinnamate[149]
Fragrance, antimicrobial and insect repellent Vanillin, benzoic acid andIodine, N,N-diethyl-m-toluamide and dimethyl-phthalate[44]
PolyesterAntibioticsCiprofloxacin[150]
AntimicrobialCurcumin[151]
4-tert-butylbenzoic acid[152]
WoolInsect repellentCitronella essential oil[78]
Cotton and PolyesterInsect repellentCitronella essential oil[38]
Cotton, wool and polyesterFragranceβ-citronellol, camphor, menthol, cis-jasmone and benzyl acetate[153]
Table 4. Use of cyclodextrins as a removal agent in the textile process.
Table 4. Use of cyclodextrins as a removal agent in the textile process.
MethodDyeReference
Cyclodextrin incorporated into a matrixCrystal Violet[162]
Reactive Black 5[80]
Methylene Blue[174]
Methyl Orange[175]
Safranin O, Brilliant Green and Methylene Blue[161]
Methylene Blue and Safranine T[176]
Methylene Blue, Acid Blue 113, Methyl Orange and Disperse Red 1[177]
Remazol Red 3BS, Remazol Blue RN, Remazol Yellow gelb 3RS 133[178]
Methyl Blue[169]
β-cyclodextrin polymerAcid Blue 25, Reactive Blue 19, Disperse Blue 3, Basic Blue 3 and Direct Red 81[81]
Basic Blue 3, Basic Violet 3 and Basic Violet 10[170]
Direct Violent 51, Methyl Orange, And Tropaeolin 000[179]
Congo Red and Methylene Blue[41]
Evans Blue, Chicago Sky Blue, Benzidine, P- hloroaniline [180]
Methylene Blue And Methyl Orange[42]
Congo Red, Methylene Blue, Methylene Orange[156]
Basic Orange 2, Rhodamine B, Methylene blue trihydrate, and Bisphenol A[40]
Methyl orange, Congo Red, Rhodamine B[181]
Direct Red 83:1[171]

Share and Cite

MDPI and ACS Style

Bezerra, F.M.; Lis, M.J.; Firmino, H.B.; Dias da Silva, J.G.; Curto Valle, R.d.C.S.; Borges Valle, J.A.; Scacchetti, F.A.P.; Tessaro, A.L. The Role of β-Cyclodextrin in the Textile Industry—Review. Molecules 2020, 25, 3624. https://doi.org/10.3390/molecules25163624

AMA Style

Bezerra FM, Lis MJ, Firmino HB, Dias da Silva JG, Curto Valle RdCS, Borges Valle JA, Scacchetti FAP, Tessaro AL. The Role of β-Cyclodextrin in the Textile Industry—Review. Molecules. 2020; 25(16):3624. https://doi.org/10.3390/molecules25163624

Chicago/Turabian Style

Bezerra, Fabricio Maestá, Manuel José Lis, Helen Beraldo Firmino, Joyce Gabriella Dias da Silva, Rita de Cassia Siqueira Curto Valle, José Alexandre Borges Valle, Fabio Alexandre Pereira Scacchetti, and André Luiz Tessaro. 2020. "The Role of β-Cyclodextrin in the Textile Industry—Review" Molecules 25, no. 16: 3624. https://doi.org/10.3390/molecules25163624

Article Metrics

Back to TopTop