Next Article in Journal
Polyvinyl Alcohol Polymer Functionalized Graphene Oxide Decorated with Gadolinium Oxide for Sequestration of Radionuclides from Aqueous Medium: Characterization, Mechanism, and Environmental Feasibility Studies
Next Article in Special Issue
Chemical Recycling of Vacuum-Infused Thermoplastic Acrylate-Based Composites Reinforced by Basalt Fabrics
Previous Article in Journal
Development of Natural Polysaccharide–Based Nanoparticles of Berberine to Enhance Oral Bioavailability: Formulation, Optimization, Ex Vivo, and In Vivo Assessment
Previous Article in Special Issue
Risk Analysis on PMMA Recycling Economics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Possibility Routes for Textile Recycling Technology

by
Damayanti Damayanti
1,2,
Latasya Adelia Wulandari
2,
Adhanto Bagaskoro
2,
Aditya Rianjanu
3 and
Ho-Shing Wu
1,*
1
Department of Chemical Engineering and Materials Science, Yuan Ze University, 135 Yuan-Tung Road, Chung-Li, Taoyuan 32003, Taiwan
2
Department of Chemical Engineering, Institut Teknologi Sumatera, Jl. Terusan Ryacudu, Way Huwi, Kec. Jati Agung, Lampung Selatan 35365, Indonesia
3
Department of Materials Engineering, Institut Teknologi Sumatera, Jl. Terusan Ryacudu, Way Huwi, Kec. Jati Agung, Lampung Selatan 35365, Indonesia
*
Author to whom correspondence should be addressed.
Polymers 2021, 13(21), 3834; https://doi.org/10.3390/polym13213834
Submission received: 9 October 2021 / Revised: 29 October 2021 / Accepted: 4 November 2021 / Published: 6 November 2021

Abstract

:
The fashion industry contributes to a significant environmental issue due to the increasing production and needs of the industry. The proactive efforts toward developing a more sustainable process via textile recycling has become the preferable solution. This urgent and important need to develop cheap and efficient recycling methods for textile waste has led to the research community’s development of various recycling methods. The textile waste recycling process can be categorized into chemical and mechanical recycling methods. This paper provides an overview of the state of the art regarding different types of textile recycling technologies along with their current challenges and limitations. The critical parameters determining recycling performance are summarized and discussed and focus on the current challenges in mechanical and chemical recycling (pyrolysis, enzymatic hydrolysis, hydrothermal, ammonolysis, and glycolysis). Textile waste has been demonstrated to be re-spun into yarn (re-woven or knitted) by spinning carded yarn and mixed shoddy through mechanical recycling. On the other hand, it is difficult to recycle some textiles by means of enzymatic hydrolysis; high product yield has been shown under mild temperatures. Furthermore, the emergence of existing technology such as the internet of things (IoT) being implemented to enable efficient textile waste sorting and identification is also discussed. Moreover, we provide an outlook as to upcoming technological developments that will contribute to facilitating the circular economy, allowing for a more sustainable textile recycling process.

Graphical Abstract

1. Introduction

The improvement of living standards around the world due to economic development could be affected by the textile industry. The textile industry faces huge challenges related to environmental issues. In 2018, the total global textile production was 105 million metric tons [1,2,3,4]. In total, up to 64% of the textile fibers that are produced are produced from the petrochemical industry. The rest of the fibers, 36%, are shared by cotton, 24%, cellulosic fibers, 6%, and wool 1% and natural fibers. With its present fast fashion business model, which is defined by mass production, variety, agility, and affordability, the apparel industry has made a significant contribution to the amount and rate of trash generation [5]. The volume of textile waste that is primarily being disposed of in landfills or incinerators is continuing to significantly increase [6]. Less than 1% of the material used to produce clothing was recycled into new clothing [7].
Textile waste can be categorized into pre-consumer and post-consumer waste. Furthermore, there are several materials that can be used to produce fabric: (i) natural materials; the natural resources used to produce it, e.g., wool and silk (protein fibers); cotton, linen, and hemp (cellulosic); (ii) regenerated material: produced from natural polymers through processing, such as rayon, viscose, cuprammonium, and acetate (semi-synthesis fiber); and (iii) synthetic material: materials that come from the petrochemical industry, e.g., polyester and nylon [8]. Natural materials are environmentally friendly. Synthetic fibers are petrochemical-based products that are produced from materials that are not sustainable and that require enormous amounts of energy to be produced [9]. Textile waste can be recycled and utilized in secondhand stores to create value-added items. The circularity approach is currently being used in the textile reuse and recycling process, which is characterized by changing from the old linear paradigm of “take, make, and trash” to a circular model, where fiber, fabric, or garments can be used to their maximum potential [6,10]. Currently, recycling technologies have five problems: (i) lack of commercially viable recycling technologies for a low-grade textile fraction; (ii) lack of mainstreamed, up-scaled processes and know-how to separate fiber types from the mixed blends and composite structures; (iii) costly recovery process; (iv) the fact that the recycling end-market is dominated by low-quality materials and blends; (v) and costly logistics and the low availability of textile recycling plants on both local and regional levels [11]. In addition, textile recycling techniques should be pass the Life Cycle Assessment (LCA). The LCA suggests a critical area in the textile supply chain, specifically, it emphasizes recycling techniques and transportation distances between the textile recycling plant and waste collection areas due to the amount of energy that is consumed during transport [12,13,14,15].
Due to the enormous waste and pollution caused by the phenomena of fast fashion, the textile sector faces tremendous environmental and resource challenges [16]. The transition to a circular economy (CE) is already underway, and it offers an effective option for resource scarcity and waste disposal issues [17]. Sorting for textile material identification and separation takes place before textile waste is recycled due to the complex material in textile fibers [18]. Moreover, in the future, the large-scale separation and sorting of textile waste, regardless of the strategy that is selected, quick and reliable methods to accurately quantify the polyester content in cotton, polyester, and other textile material compound mixes are urgently needed [19]. Specific materials should be recycled, e.g., medical textiles, which are all fiber-based products and structures that are utilized for emergency treatment and clinical, surgical, and hygienic purposes [20]. The process of recycling textile waste can be easily achieved by Internet of Things (IoT) technology. IoT can assist in collecting the data provided by various sensors, including smart meters, in connecting stakeholders across the value chain [21]. Furthermore, IoT needs data collection due to the fact that smart sensor and RFID technologies require big data [22]. Fatimah has been introduced to modern technologies such as Indonesia’s IoT-based sustainable CE approach; waste management has also benefited from IoT technology [23].
Identifying textile materials is the most challenging task in the textile recycling process because of the complicated structures of fabrics and textiles. Cellulose polyester mixtures are among the most popular multi-component garments on the marketplace and can be used to customize material attributes such as moisture absorption, wrinkle resistance, and wearing comfort [24]. There are some promising methods that can be used to analyze the complex components of different textiles. For example, applying near-infrared (NIR) spectroscopy and chemometrics for textile analysis was reported over two decades ago [25]. The cotton component, which comprises mixed cotton and polyester, was determined by genetic algorithms [25]. In addition, nuclear magnetic resonance (NMR) has been used to predict the proportion of cellulose and polyester in textile blends [24]. Nevertheless, solid-state NMR can show lower sample throughput and efficiency. Still, it allows for more precise measurements that are not influenced by factors such as particle size, dyes, or surface structure [26]. On the other hand, the nature of the sample surface, such as particle size, brightness, color, moisture content, and coating chemicals, has a significant impact on the NIR signals that are produced [27].
Polyester and cellulose performed very differently in terms of behavior, especially when considering differences such as thermal stability and hydrophilicity. The diversity of properties in these textiles allows them to be reprocessed and separated in different ways, such as through depolymerization, re-polymerization, and spinning [24]. Therefore, solving these problems requires a comprehensive analytical toolkit for the qualification and identification of textile materials [28]. Textile can be recycled through mechanical, thermal, and chemical methods. Textile recycling mainly entails the reprocessing of post- or pre-consumer waste fabric in order for it to be reused in new textile or non-textile products. The three types of textile recycling routes are the dissolution of natural fibers, the chemical degradation of polymeric fibers, mechanical treatment (pretreatment), and thermal treatment [29].
A research gap exists in the field of textile recycling technology, and it is possible for textile recycling technologies to be implemented based on the CE principle. Current technologies such as the IoT and big data can be used to enable the efficient collection of textile waste, identification, and recycling. Still, there are some challenges to recycling textile waste due to different materials that are added to fabric and textiles to make fashion items. Therefore, the strategies that can be used to recycle textile fibers, such as mechanical recycling, chemical recycling (pyrolysis, enzymatic hydrolysis, hydrothermal, ammonolysis, gasification, glycolysis), and decolorization technology in textile recyclizing, are reviewed to minimize the textile waste that ends up being incinerated or in landfills.

2. Circular Economy of Textile Recycling

A CE is a concept where material production can be redesigned. It utilizes resources to produce, use, and dispose of in favor of as much reuse and recycling as possible [8]. The CE approach has an environmental, industrial economic, and ecological basis [30,31,32]. There are some barriers impeding the implementation of CE in the fashion industry due to the various fashion styles, aesthetics, and the roles of shoppers. The regulations and policies regarding the environment and sustainability of CE need to be considered [19]. It has also been determined that the textile business also has a detrimental impact on society because it consumes enormous resources such as oil, carbon, and water [33,34]. Apart from that, it eliminates the over-generation of waste and obtains the total value of products by focusing on the CE through giving products a longer life and through reusing materials [8]. The CE is driven by three main methods and approaches (reduce, reuse, and recycle); all traditional waste managements are practiced [35].
Fabric waste reduction includes all of the manufacturing stages (including a minimum of feedstock) and various stages of use and consumption. The concept of the reuse stage is that textile materials are easily recyclable or repurposed for different types of applications. In addition, increasing reuse and repurposing will reduce the amount of textile materials that are needed for production [18,36]. Figure 1 shows this critical point in the textile CE. It is vital to have a CE because it emphasizes seeking scientific solutions to complete the loop. Waste is reduced at the source and is recycled back into the economy for reuse rather than production and consumption being stopped due to garbage disposal [37]. The CE cycles includes recycling, reuse, and remanufacturing. Textile reuse involves various methods for extending the useful life of textile items by redistributing them to a new buyer [38]. Nevertheless, the secondhand clothing markets will end up exporting these secondhand textiles to emerging market countries. On the other hand, CE emphasizes local consumer reuse, and the retailers must contribute to this circular system. This system can be applied to introduce new customer segments and new potential for reuse models [39].
Remanufacturing is used in the fashion and textile industry to describe the process of reconstructing worn clothes to create new garments by considering the quality of the textile product or the customer value [40,41]. In the early 21st century, the consideration of textile remanufacturing gained popularity because fashion designers and entrepreneurs were more aware of the sustainability of the post-consumer raw materials and textile waste [42,43]. The reasonable prices and the remanufacturing of the textile industry in some developing economies, such as in China, have been more competitive [44]. Furthermore, by using an eco-efficient value creation (EVC) model integrated with analysis cost, eco-cost, and retail value, remanufactured goods have a lower eco-cost due to their reduced materials and pollution. An EVC has a positive effect on the cost-benefit [45].
Due to the textile sector, recycling is the principal environmental polluter, and textiles have complicated the manufacturing process. Furthermore, the microfibers in textile waste harm the oceans, and the textile industry contributes to this pollution. Globally, textile waste increases, and recycling could help reduce new material waste [14]. Mechanical and chemical recycling are two primary types of recycling technologies. Mechanical recycling via melt-extrusion is a technique that is used to extract fibers from waste materials, allowing them to then be spun into yarns. Therefore, to obtain great yarn strength and fineness, this technique must be applied to virgin fibers.
On the contrary, chemical recycling occurs when the polymers of textile waste are depolymerized into small monomers and/or are re-polymerized to make new fibers. The chemical recycling of cellulose fabric such as cotton or viscose also uses the dissolution approach. At the same time, organic solvents apply ionic liquids to dissolve cellulose polymers [45,46]. In addition, the valorization process offers another potential textile recycling method. Li et al. developed the term “textile waste valorization”, which is related to the combination of “textile waste recycling and hydrolysis process”. This process degrades the complex polymers found in textiles in order for them to become various value-added products that are supported by microorganisms such as enzymes. This kind of method is inspired the CE approach, which aims provide multiple solutions for recycling textile materials [47].
Textile material recycling affects a variety of substances and contributes significantly to current social obligations. The recycling process allows businesses to make a more significant profit by avoiding textile waste being dumped in landfills. Meanwhile, the textile recycling contributes a positive impact on social and environmental issues, e.g., gives job opportunities to workers who are unemployed or underemployed, philanthropic commitments, and the expansion of reusable textile to the countries around the world that need proper cloth [48,49].
There are many challenges when transitioning the textile industry to a circular economy: (i) the market for recycled textile products is still limited; (ii) there is a lack profitable circular business models on the market; (iii) there are challenges in having supply chain partners work together to create innovations; and (iv) the quality of recycling of textiles of a relatively low grade is pricey in the short-term, with low economic benefits [50]. There are some fundamental strategies for recycling business models (RMBs); for instance, recycling process, partnership, and scale up the business models (BMs). Furthermore, approach RMBs are becoming more practicable. Collaboration between various stakeholders and across a variety of industry sectors is needed [51]. In addition, the participation of waste companies and educational institutions in investigating several options for closing material loops and identifying waste streams is a crucial change agent to facilitating the textile CE transformation [52].

3. Starting with Municipal Waste to Textile Trash in Applied IoT

Sustainability is also required in the waste management sector. It starts with the manufacturing process to the disposal phase. A loop feedback system for sustainable waste management comprises several parts, such as process activity and waste diversity [23,53]. The direct regulations of the European Union’s CE action plan will prohibit the disposal of unsold textile products and a restriction on single-use products, and according to the Recycling and Waste Reduction Act 2020 with replaced the product stewardship Act 2011 in Australia, there are textile regulations that are related to the ambition of separate textile collection by 2025. A new French legislation has established a command-and-control prohibition on the destruction of various unsold products, including fashion items. This legislation requires that new products use a minimum amount of recycled material [54].
Furthermore, to support the current regulation and policy of waste textiles, digital innovations have been adopted. Recently, digital innovation and industry 4.0 have positively influenced CE transformation, such as data-driven product lifecycle analysis [55]. Industrial revolution 4.0 is achieving waste treatment sustainability, with waste treatment becoming more practicable, dependable, clear, efficient, and optimal due to the application of IoT, digitalization, and information and communication technology (ICT). ICT and IoT implementation can minimize the time and resources needed to improve waste management effectiveness and to move toward more sustainable and intelligent systems [56]. In addition, these methods will be selected as the best treatment technologies for the waste material in terms of technical, economic, and environmental feasibility. IoT and ICT can be combined to gather, analyze, and share data [23].
Digitalization can be improved in product/raw materials tracing and tracking by providing real-time data on product availability, location, and condition. At the same time, sensors and digital platforms can enhance the product life and can create CE solutions more effective and efficient [55]. IoT-enabled manufacturing is related to advanced concepts. Traditional manufacturing resources can be transformed into smart manufacturing objects (SMOs) that have the ability to interact, sense, and interconnect with one another to perform manufacturing logic [57]. The fundamental issue with these technologies is the sharing and collecting of real-time data, including wireless communication standards and radio frequency identification (RFID). Physical production flows, including material transportation, and associated information flows, such as the visibility and traceability of different manufacturing operations, can be easily connected utilizing RFID technology [58,59].
Cloud computing technology is a data management solution, and it can be used as a replacement for local servers [60]. It has resulted in a massive scale of data sharing among several users and has become the latest data management standard [61]. These processes and resources must be handled intelligently in cloud manufacturing technology because the product life cycle includes the start of the manufacturing process, and includes design, manufacture, and maintenance [62]. Resources can be controlled and managed automatically by applying various IoT technologies, making digital sharing easier. Underpinning technologies are typically service-oriented, and cloud manufacturing complements the conceptual framework of cloud manufacturing [63].
The design approach for the sorting and identification of textile waste is based on the IoT in Figure 2. Waste materials are sorted through a funnel and a single metallic sensor. The metallic sensor detects any metal compounds that are going through the funnel. The metal materials go into metal storage. If the object does not contain any non-metal materials, then it is placed on the sensing platform. The capacitive sensors sense the object while it is on the sensing platform based on its moisture content. Then, the materials that have moved through capacitive sensors are divided into two sections: plastic bottles and textiles. Furthermore, the sorted textile waste will undergo several sorting and identification processes due to the complex mixture of materials [64].
On the other hand, manual sorting can be applied. Mechanical machines will transport the materials to a manual sorting platform by conveyors for preliminary sorting, which will sort the hazardous materials, bottles, large glasses, cotton, and textiles to be recycled. Each material will be separated based on its characteristics and will then be delivered to each storage bin. Each storage bin is equipped with a weight level sensor that sends the general information into the database. The database will receive and store all of the data related to the system’s waste workflow. According to the performance of the trash that is collected and the existing processing technology, the database will suggest treatment options [23].
Yu et al. developed a hybridized intelligent framework (AIHIF) based on artificial intelligence (AI) for automated recycling to improve trash management. By combining machine learning and graph theory, the system enhances waste-collecting over a short distance. Furthermore, AI design technology can be supported through various approaches that can be adapted to specific interest groups, collecting their data and increasing ecological planning and urban management production, precision, and efficiency [65]. Enevo has already developed recycling systems and innovative waste treatment technologies with the Enevo One software. The software contains ultrasonic wireless sensors. Waste collection optimization has been performed not only by cloud service in the background but also by the web interface used for Enevo One and the sensors themselves. Enevo’s sensors are compatible with more than 100 container types. The following waste types are currently supported: mixed, textile, glass, metal, and waste electrical and electronic equipment (WEEE) [66]. In addition, Stack4things was created by the mobile and distributed systems lab at the University of Messina in Italy. The OpenStack add-on controls sensing and actuation resources, including remote control, virtualization, and network overlays. The main objective is applied through cloud computing to manage data and changes to a software as a service (SAaaS) perspective, in which developers or end-users are provided with actual or virtualized sensing and actuation resources [67].

4. Sorting and Identification of Textile Waste

The dilemma for textile recycling is sorting and identifying textile materials due to the complex structure of textile polymers. In addition, the majority of textile materials comes from fiber, and the main challenges are sorting and recycling due to the complicated fabric of the textile. The fabric waste is commonly mixed with other materials, such as buttons, zippers, or other decorations [68]. There are various ways to identify textile materials, including iso-standardized quantification methods based on distinct dissolution behavior (ISO 1833-1, for example) and morphological variations found using microscopy [69]. Figure 3 shows the source and mechanism of textile recycling, a common source of textile waste from consumers, and unsellable textile waste from the textile industry. In addition, the sorting and collecting of textile waste is not able to be achieved manually, which would improve material quality assessment [70].
The thermal properties of textile materials are essential qualities of these materials and are regarded as the most crucial consideration when describing the comfort properties of apparel [71]. In addition, several different thermal properties have been demonstrated in numerous research studies, such as in textile fiber polymers, which are semi-crystalline, whose thermal conductivity is affected by the molecular structure, crystallization level, and molecular chain mobility of amorphous regions, etc. [72]. One of the methods that can be used to analyze the thermal properties of textile waste is by thermogravimetric analysis (TGA). The data resulting from TGA are related to the thermal stability and degradation degree from the fibers. TGA is a simple method that can be used to obtain data on the thermal behavior of a textile before that textile waste is pyrolyzed in the industry [73]. Therefore, some real-time technology is required to develop the textile material recognition rate.
Table 1 indicates the recognition rate using different types of analyzers. A single type of textile waste polymer is easier to convert into small molecules with a higher value than when other types of waste polymers are presents. Recently, near-infrared (NIR) technology has been applied to sort textile waste. NIR spectroscopy can be implemented automatically by convolutional neural network machine learning (CNN) followed by wavelength recognition that has been transformed through image classification. Then, the surface area of the spectrum is converted and modified by Textile Recycling Net (Tr-Net), and the images are sorted using the softmax classifier, allowing the qualitative of textile compositions to be obtained [74,75].
Wilts et al. developed a new AI technology to sort textile waste from municipal waste. This robot with AI could be applied with a recognition ability of more than 95%. On the other hand, textile waste recovery was not achieved, with only 13% of textile waste being recovered due to the complicated structure of materials [76]. Xiong designed a cheap near-infrared device to sort the seven types of plastics using machine learning. Machine learning could identify the plastic types with 90–99% certainty [77]. Berghmans et al. studied infrared spectrometer techniques to identify the post-consumer and industrial carpet waste. The carpet waste materials came from polyamide-6, polyamide-66, polyester, polypropylene, and wool. The complex of the carpet materials can be determined by analyzing the radiation passing through the fibers. The range of wavelengths near and mid-infrared was (800–2500 mm) and (2500–25,000 mm), respectively [78].
Riba et al. used an ATR-FTIR spectrometer (Attenuated Total Reflection–Fourier Transform Infrared) to sort and classify textiles based on the type of polymer materials in that textile. Mathematical modeling was applied to more accurately predict the data from the IR spectrum. The recognition rate percentage achieved by the mathematic algorithms (principal component analysis, canonical variate analysis, k-nearest neighbors algorithm (k-NN algorithm)) could detect synthetic and natural fibers, such as polyamide, viscose, polyester silk, linen, cotton, and wool, with an accuracy of up to 100% [86]. Furthermore, Peets et al. evaluated the capability of reflectance-FTIR (r-FTIR) to recognize textile fibers using an FTIR micro-spectrometer. In reflectance mode, the FTIR micro-spectrometer several reflectance spectra from textile fibers can be collected, and these spectra contain enough characteristics to create classification models and to identify unknown textile samples. Nevertheless, r-FTIR is a good approach for analyzing many textile samples quickly, easily, non-destructively, and non-invasively [87]. Morgado et al. identified PE, PET, PP, and PS using micro-ATR-FTIR spectra with known uncertainty [88].
The following three primary characteristics are reflected in the recognition and classification accuracy: (1) when physical properties, such as fabric color, yarn thickness, diameter, orientation, and uneven light, are considered, the model remains robust; (2) a model with fewer parameters should use the transfer learning technique, making it more computationally efficient; (3) the suggested model does not use handcrafted features, instead relying on a fully automated end-to-end architecture to extract and classify features [82].

5. Mechanical Recycling of Textile

Mechanical textile recycling has been developed since the industrial revolution began. Furthermore, mechanical recycling is one of the easiest recycling methods can be used with a low-cost budget. Textile waste is classified by the type of material and the color achieved through mechanical deconstruction; for instance, wool converts into useable yarn (re-woven or knitted) when carded yarn is spun with mixed shoddy [89,90]. Mechanical recycling can be categorized into several different methods based on the degree of breakdown that is undergone by the recovered material, such as fiber, fabric, polymer, and monomer recycling. The mechanical recycling applications in the wiper, fiber material, and prespun fiber industries could produce new fabrics for products [91].
On the other hand, there are few mechanical recycling options for post-consumer textile waste fabrics due to the variety and amount of fiber elements in one textile waste garment; mixed textiles cannot be mechanically recycled effectively. The shredding technique reduces the length of post-consumer textile waste garment fibers. The resulting short fibers are of a lower quality and strength than virgin cotton fibers, and they are frequently combined with virgin fibers to improve quality [11,13,92,93].
The metallic elements are removed when post-consumer textiles are converted to wipers. Textiles from the post- and pre-consumer markets can also be used as a raw material for filler products, including flocking, insulation, and nonwovens as well as shoddy fibers for recycled yarns [94]. A defector is developed with the mechanical ripping and carding processes during the production process. The first step of the defector process requires the feedstock to be shredded by the rotation crusher and sorted based on the type of fiber. Secondly, the fiber is crushed through correct fraction sizes and processes with the card oil. Furthermore, the defibering process is a three-phase mechanical ripping carding process. The materials are needled and folded to create a fiber blanket. In the needling process, other materials could be added to the product, e.g., plastic material. Then, the fiber blanket is trimmed to the correct size and is packed [95]. On the other hand, some textile materials cannot be recycled by mechanical process due to the low quality and difficulty of reprocessing these materials through the melting process to achieve high-value products, for instance, cotton or a mixture of synthesis fibers [90]. In addition, mechanical shredding harms natural fibers more than synthetic fibers in blended fabrics such as polyester–cotton. Chemical recycling becomes an alternative to recycle both fiber types [96].
Figure 4 shows an open and closed-loop garment recycling. The term “open-loop recycling” refers to methods in which fabrics or textiles are used to make new products [52]. Besides that, recycling wool through open-loop recycling can be used to mechanically extract the garments back into a fibrous form, and these materials can be used as raw materials. Non-woven fabrics are frequently made from these materials by garneting, carding, or air laying webs, which are then mechanically, thermally, or chemically bonded. Furthermore, hot water converges with the fibers when the melt-blown process is applied, followed by spun lace. The purpose of spun lace is to entangle fibers. The method utilizes extremely tiny high-velocity water jets rather than sharp needles when the textile is going through the drying process [97]. Besides that, the melt blowing process can be combined with polymer melt extrusion passed by the attenuation of the extrudate and orifices under extremely high-velocity air jets to generate nonwoven fibers in just one step [98].
In addition, Tsai described a technology to develop an electret via saturating a nonwoven textile with a liquid and then removing the fluid. This technology contributed (i) a sheet made from a thermoplastic fiber nonwoven web; (ii) the ability to place the sheet underneath a roller: water jets were added in these steps to acquire a saturated sheet, and (iii) a suction process that is able to remove the water at the saturated sheet that then allows it to dry [99]. The applications of open-loop recycling include insulators for mattresses and sound absorption [97]. Nylon waste, such as fishing nets (nylon 6) and carpets, is the common raw material for that is used in mechanical recycling. Nylon materials are recycled via reshaping and melting to create new fibers with appropriate strength and length [100].
Closed-loop textile recycling refers to processes in which the recycled material is identical to the provided material. Both pre-and post-consumer mechanically recycled materials can be considered to be closed-loop recycled since the waste material or fiber re-enters the apparel generation chain. The fineness, length, polymer quality, and shading of the new fiber properties can determine the quality and to create the most appropriate final product [48]. A closed loop conducts the recycling of wool fibers. Using traditional techniques, recycled wool yarns can be turned into very high-value apparel products. Woolen materials are mainly formed from long fibers that have been carefully treated, allowing for mechanical recycling to covert the wool into fibers. Wool recycling follows the same methods as mechanical cotton recycling [101]. The Cardato brand coat from M&S Shops is manufactured using closed-loop recycled wool. The recycled wool waste should ideally be compatible with the same carding, spinning, and fabric production processes used to process raw wool during the production of woolen yarn [101,102,103]. The post-consumer wool recycling can be applied. This depends on effective strategies to reduce fiber breakage and to maximize residual fibers length following the mechanical pulling method being available [104].
Klaus-Nietrost et al. studied the recycling of cellulose-based textiles with a minimum of one synthetic plastic. The regeneration process was applied to acquire a cellulosic-molded body. In addition, the cellulosic-molded body was composed of a crucial substance in the form of lyocell—fibers or viscose—fibers. Chemical fibers, viscose fibers, and regenerated fibers are all terms that are used to describe different types of fibers. They can be produced using the viscose technique, which is a wet-spinning method [105,106]. The cellulosic material is soaked in N-methyl morpholine-N-oxide (NMMO) and is pressed with a spinneret. It coagulates and generates a new fiber [107]. When recycled fibers are applied in cellulose production, the main issue is the purity of the raw materials, such as old-fashioned textile materials, and more significantly, the contamination of synthetic polymer materials. Nevertheless, the synthetic polymer can be denoted as being combined with cellulose materials [105].

6. Chemical Recycling of Textile

Chemical treatment is seen as a promising textile recycling method. During this process, chemicals are added to degrade the complex textile polymer to create smaller polymer molecules. The “CRM”, or chemical recycling to monomer, is a popular nowadays that can be used to degrade polymers such as those in textile waste to monomers. Furthermore, a solution-purified polymer, an oligomer, a monomer, or a crude feedstock in gaseous or liquid form are all examples of textile chemical recycling products. Re-polymerization can be used to renew polymers such as oligomers and monomers [90,108]. Table 2 lists the industrial scale that is applied when recycling textiles through various chemical processes. Typical chemical processes are glycolysis, hydrolysis, and alcoholysis. In addition, thermo-chemical methods consist of hydrocracking, pyrolysis, and gasification, which are typically conducted in the presence of catalysts at high temperatures [109]. Meanwhile, biochemical processes use enzymes or microorganisms in recycling processes such as enzymatic hydrolysis [106].
The drawbacks and advantages of the various chemical textile recycling processes are shown in Table 3. Chemical polymer recycling for cellulose-based synthetics and mixtures of textile fibers has been studied. Textile fibers are destroyed by mechanical methods, such as shredding, and their destruction is followed by a chemical dissolution process. Hazardous solvents are often used; the polymer is retained while the fibers are regenerated and spun [119]. In addition, the dyes need to be removed from the textile waste, and they are removed using harmful chemicals such as bleach. Ionic liquid solvents have recently been used to reintroduce dyed post-consumer textile waste into a new life cycle and to reduce environmental effects [120]. Furthermore, the conversion of cotton to viscose is one of the most well-known examples of chemical textile recycling. Pure cotton fabric is depolymerized into a pulp, which is transformed to viscose. The production process is similar to that made from wood pulp. Industry leaders such as Lenzing and Birla Cellulose adapted the conversion of cotton to viscose technology, with the beginning of the process being the depolymerization of 100% cotton textile materials. Nevertheless, the yield of the synthesized polymer chains is lower than that of the pure wood pulp process due to decreased physical properties. Therefore the regenerated fibers must be combined with virgin viscose fibers to achieve a high yield [121].
Nylon 6 is a polyamide made from the seven-membered ring caprolactam, making it a valuable candidate for CRM. Since the early 1960s, researchers have been studying the depolymerization of nylon 6 back to a monomer, and CRM has been used for nylon 6 for decades [122]. The reactor is conducted by superheated steam and a catalyst to form a caprolactam distillate. Nylon 6 is made from crude caprolactam that has been distilled and depolymerized. In terms of purity, the caprolactam obtained is similar to virgin caprolactam. The repolymerized nylon 6 is spun into yarn and is tufted into carpet. The physical properties of the carpets produced using this technology are fairly similar to those of virgin caprolactam carpets [123,124,125]. In addition, textile waste can be used to produce geotextiles (thick ropes), which can be used to protect slopes from erosion and sliding. Natural fibers are dissolved with mixtures of 5% sodium hypochlorite (NaClO) and Schweizer’s reagent for cellulosic fibers as solvents. Then, the geotextiles are installed in conditions where the ground is sloped and where there is vegetation—this method reduces soil movement and pedestalling. [126,127]. Another application for textile waste that has been processed by chemical recycling is for use as packaging. The cellulose nanofiber textile waste has been used to enhance the strength performance of biopolymers. The combination of polylactic acid and cellulose nano-fibrillated fiber (CNF) for packaging applications has shown excellent strength reinforcement ability, tensile modulus, and elongation [128,129].
The chemical recycling of textiles usually necessitates integrated material flow and energy within the chemical industry. Nevertheless, some sorted waste products from municipal solid waste have more potential for small-scale feedstock production [130]. Figure 5 shows the possible pathways to recycling textile using chemical methods. Furthermore, Yousef et al. studied a long-term approach for recovering cotton from textile waste and regenerating it to become a new substance. The primary processes used to recover the cotton were leaching nitric acid, dissolving dimethyl sulfoxide, and diluting hydrochloric acid for bleaching. This recovery rate, carbon footprint, and economic performance of this technology were determined to be 93%, −1534 CO2 eq/ton, and 1466 USD/ton, respectively [73] if textile waste is used as a renewable resource for the production of thermal energies. Nunes et al. showed that cotton briquettes had a heating value of 16.80 MJ/kg and a cost of 0.006 EUR/kWh when used as a fuel [131].

6.1. Textile Recycling Using Pyrolysis

Textile recycling using pyrolysis is a promising technology that can be used to degrade the carbon-polymers of solid waste in order for them to become three pyrolysis products in the solid, liquid, and gas states [140,141,142]. Pyrolysis could be utilized in diverse textile materials that have not been sorted prior or in multi-material textile products that otherwise would have only been treated in waste-to-energy facilities [143]. Under oxygen-free circumstances, pyrolysis reallocates C/H/O elements from organic molecule waste into three-phase pyrolysates. Syngas (a combination of H2 and CO) produced by thermochemical processes was utilized as a direct fuel or as a raw material to produce other hydrocarbons and alcohols [144,145,146].
The pyrolysis process does not require pretreatment, and it is a potential method for treating polluted waste. These factors make pyrolysis a more efficient process than chemical methods (including bio-chemicals), which require many chemicals and frequently produce a lot of trash that must be re-landfilled and takes longer to operate on a large scale [147,148,149]. Table 4 shows a summary of the diversity yields of the solid, liquid, and gas products produced via textile waste pyrolysis. Commonly, the liquid products produced from textile waste are mono-polyaromatic and oxygen compounds containing hydrocarbons, such as alcohols, aldehydes, ketones, and carboxylic acids [150]. The production of oil and tar formation from the pyrolysis process depends on raw materials and operation conditions [151].
The aromatic oxygen and hydrocarbon compounds were found at 500 °C. Furthermore, when the temperature reaction was as high as 600 °C, alkylphenols were produced, and at a temperature greater than 800 °C, the yield of the oxygen compounds began to decrease slowly. The condensable compounds at 800 °C were predominantly aromatics with no substituent groups, benzene, or naphthalene. On the other hand, if the temperature reaction was about 850 °C, the aromatic hydrocarbon compounds of 3 and 4 rings were produced [152].
Char derived via textile waste pyrolysis can be applied as a primary filler or a hybrid filler (carbon nanotubes/nanoball (CNTs/CB)) or as graphene oxide (GO/CB) in concrete composite applications. To synthesize CB of 10–20 nm, the process is conducted in a pyrolysis reactor followed by milling and chemical treatments to modify the cotton textile waste (CTW). Then, synthetic CB (0.05 wt%) and hybrid fillers (CNTs/CB and graphene/CB: 50/50) are applied to improve the cement paste characteristics [153]. On the other hand, Jagdale et al. studied applications of recycled and reused carbon produced from cotton waste by the pyrolysis process in active electrode battery materials. The cotton-based carbon fiber electrode demonstrated outstanding cycling behavior and provided a high discharge capacity, with a voltage range of 0.02–1.2 V [154].
Furthermore, pyrolytic char derived from cotton waste can also be utilized as an adsorbent. A one-step low-temperature pyrolysis process produced char-based adsorbents made from CTW with different iron salts compounds. The Freundlich model fit the adsorption processes well when measuring Cr(VI) the adsorption performance of the chars, and char-FeCl3 had the highest adsorptive capacity of 70.39 mg/g. Consequently, this low-temperature pyrolysis method was cost-effective and straightforward to implement, with a high adsorption capacity for Cr(VI) removal [158]. The catalytic pyrolysis of nylon 6 waste fishing nets (WFNs) over a ZSM-5 zeolite catalyst was investigated by Eimontas et al. The activation energy using ZSM-5/WFNs was larger than 112 kJ/mol via the free kinetic analysis method. The major volatile organic compounds were alkyl C–H, carbonyl (C=O), and caprolactam [159].
Silva et al. studied colored cotton wastes to generate renewable aromatic hydrocarbons by the catalytic reforming of ash pyrolysis utilizing a pyrolysis raw material (colored cotton wastes). These materials have high energetic compactness, which is good for use thermochemical processes. Light oxygenated carbon chains with up to four carbons resulted from the ash pyrolysis. The partial deoxygenation and cracking reactions were increased by pyrolysis vapor reformation at 500 °C, which could be improved by catalytic reforming [160].
The catalytic pyrolysis over Ultrastable Y (USY) zeolite was investigated by Wang et al. This method produces benzene, toluene, and xylene (BTX) via Diels–Alder reactions by means of catalytic co-pyrolysis with specified plastic wastes and furanic substituents via the catalytic deconstruction of textile waste. The selectivity of co-pyrolysis to BTX for polyethylene and 20% co-flax waste was 81.6% and 80%, respectively [161]. In addition, the primary components of dryer lint were produced when the drying garments were cotton and polyester. The range of the activation energy (Ea) of lint pyrolysis was 167–204 kJ/mol, and the primary products for a heating rate of 25 °C/min were furan, 3-methyl- (4.78%), furfural (6.48%), isobutane (10.77%), and carbon dioxide (11.68%) [162].

6.2. Textile Recycling Using Enzymatic Hydrolysis

Hydrolysis is a chemical breakdown process that occurs due to a reaction with water. Generally, the hydrolysis process requires a pretreatment process using acids, alkaline, and ionic liquid. Pretreatment is crucial in hydrolysis because it can affect the yield percentage [163]. Damayanti and Wu have already described recycled PET hydrolysis [109]. Before the recycling pf PET fibers from textile waste takes place, the pretreatment process is required. Pretreatment methods reduce the complexity of the structures of the textile materials by eliminating unwanted contaminants and improving hydrophilicity. PET fibers can be converted into raw materials through the following processes or reactions [164]:
The pretreatment process for textile waste materials that uses bases such as sodium, potassium, calcium, and ammonium hydroxides is known as alkali pretreatment. Alkali pretreatment will improve lignin solubilization and decreases cellulose crystallinity by enhancing cellulose digestibility. As a result, alkali pretreatment produces a high glucose yield. The reduced formation of fermentation inhibitors is a primary benefit of alkali pretreatment [165,166]. Furthermore, pretreatment using ionic liquids is a promising method due to the use of molecules that are more environmentally friendly that can dissolve cellulose at moderate temperatures without degrading the cellulose or the solvent. Solvent recovery can be applied through a thermal treatment with lower pressure [167]. The ionic liquid from organic solvents such as N-methylmorpholine-N-oxide (NMMO) and 1-allyl- 3-methyl-imidazolium chloride ([AMIM]Cl) was effective in pretreating cotton waste under enzymatic hydrolysis [168].
On the other hand, acid pretreatments such as sulfuric and phosphoric acid have been examined by researchers, even in the presence of cellulose solvents [169,170]. By breaking the polymeric structures in hemicellulose, acid pretreatments can hydrolyze the polymeric structures into monomers by increasing the availability of cellulose, hence boosting biodegradability [171]. There are some challenges such as the formation of side products (furfural), the cost of acid being relatively expensive, and the requirement of corrosion-resistant equipment [172].
Cotton waste from the textile industry has been discovered to be a viable feedstock for cellulosic ethanol production. Cotton waste can pre-treated and acid hydrolyzed to convert cellulose to reducing sugars. These findings are promising for ethanol production [173]. Table 5 lists the diverse range of textiles that can be used to generate glucose via hydrolysis. Furthermore, Pd-Au/SiO2 bimetallic catalysts are effective in oxidizing glucose generated from cotton waste. These experiments were conducted under the correct temperature and pressure operating conditions for acidic hydrolysis [174]. The recovery of polyester and sugar compounds from textile waste via enzymatic hydrolysis was investigated by Li et al. The temperature of the pretreatment process was −20 °C for 6 h, with the concentration of NaOH/urea of 7%/12%. The glucose recovery was 98.3%, with a cellulase dosage of 20 FPU/g [47].
Cellulases act as catalysts in hydrolysis reactions. In theory, these enzymes are unaffected by the reaction equilibrium and can be used again and again. Unfortunately, because these enzymes are massive organic compounds, some decomposition processes that occur over time must be considered [179]. Furthermore, these enzymes can break down material structures to change their physicochemical properties and biodistribution [180,181,182]. Enzymatic reactions involving various selections of materials can be initiated by certain enzymes [180,183]. As a result, the enzymatic breakdown of certain portions of these materials may be selective [184]. A commercial cellulase mixture typically contains three types of enzymes, each of which performs a different role in the catalytic process: (i) Endoglucanases degrade the cellulose chain at varying stages along the chain by increasing the number of accessible end parts, (ii) Exoglucanases degrade cellobiose units from both ends of the chain, and (iii) the depolymerization of the disaccharide cellobiose into monosaccharide units, e.g., glucose is carried out by beta-glucosidases. These three types of enzymes have generated excellent degradation performance in previously reported experiements [185].
Fermentation is an enzymatic degradation procedure for recovering textile waste such as cotton, polyester, nylon, and silk. For bioethanol production, five fermentation techniques can be implemented to degrade textile polymers: (i) Simultaneous saccharification and fermentation (SSF) converts sugars to ethanol [186]. Many researchers have used engineered microbes and have cultured them in biomass sugar solutions, such as glucose, xylose, mannose, galactose, and arabinose. The most commonly utilized microbes for ethanol production include Saccharomyces cerevisiae, Zymomonas mobilis, and Aspergillus niger [187]; (ii) separated hydrolysis fermentation (SHF) is a process where hydrolysis and fermentation are separated to produce bioethanol. SHF is a typical approach in which hydrolysis takes place first, followed by the fermentation process. This method allows the hydrolysis process to generate monosaccharide sugar initially, ensuring that sugar is available when fermentation starts [186,188,189]; (iii) semi-simultaneous saccharification and fermentation (SSnF): the excess sugar that accumulates during hydrolysis reduces enzyme activity, slowing down the process [186,190,191]. The short pre-hydrolytic method is used before the SSF process in SSnF. In addition, the yield of bioethanol production is be slightly higher than it would be if traditional SSF procedures were used [191]; (iv) consolidated bioprocessing (CBP) is required to degrade refractory biomass materials into solubilized sugars. Because CBP combines three processes (enzyme synthesis, saccharification, and fermentation), it is a viable technique for successful biofuel production. This technique can lower the cost of the reactor and enzymes, two main roadblocks to low-cost biomass processing [192,193]; (v) submerged fermentation (SMF) is one of the fermentation techniques that focuses on enhancing cellulase production. It is a widely accepted fermentation process for the synthesis of industrial enzymes because it is easy to regulate all of the factors such as pH, temperature, and operational approaches [186,187].
Table 6 lists ethanol production from various textile wastes using enzymatic hydrolysis and microbe fermentation. Jeihanipour and Taherzadeh studied ethanol production from the cotton from waste fabric. The alkali pretreatment was conducted in the concentration of NaOH (12 wt%) at 0 °C for 3 h, which was then fermented to ethanol via Saccharomyces cerevisiae, with an ethanol yield up to 99.1% [194]. Sanchis-Sebastiá et al. investigated green liquor as an alkaline solution in the processing of textiles to connect textile recycling with pulp mills. The textile recycling efficiency was predicted be 70% when using Cellic CTec 2 as a hydrolysis enzyme. Nevertheless, the sodium reduction in the pulping process would be less than 15% [195]. On the other hand, Sebastia et al. studied the cotton fiber recycling process when it was conducted at 130 °C via acid hydrolysis. Sulfuric acid could depolymerize the structure of cotton to obtain glucose. The yield was over 90% when a two-step method was used [196].

6.3. Textile Recycling Using the Hydrothermal Method

The hydrothermal method is a decomposition process using chemical crystallization engineering techniques at high temperatures and pressures, with water as the main constituent of the reaction. The hydrothermal method is one of the most promising alternative methods for degrading carbon-polymer waste and organic components into three product phases: liquid, solid, and gas phases. It is conducted in an autoclave reactor in which organic acids can catalyze [200]. The hydrothermal process does not require pretreatment and is a method that utilizes water and can be classified into five groups based on the temperature range that is used: (i) hot water extraction, (ii) pressurized hot water extraction, (iii) hot liquid water treatment, (iv) hydrothermal carbonization, and (v) hydrothermal liquefaction. The method requires temperatures that are less than 280 °C [201]. Table 7 lists a summary of hydrothermal products produced via textile recycling. The hydrothermal conditions control the powder formation and particle size. The drawbacks of the hydrothermal process are the high temperatures, high pressure, and long reaction time that are required for the material manufacturing process [202].
Hongthong et al. studied the conversion of nylon 6 fishing net waste in the hydrothermal liquefaction of the macroalgae Fucus serratus. The hydrothermal liquefaction of macroalgae was performed for 10 min with a temperature reaction of up to 350 °C. The scope of nylon polymers included nylon 6 and nylon 6,6. A bio-crude yield of up to 17 wt% could be with a combination of 50% nylon 6 and F serratus. Furthermore, nylon 6 fully degraded in the process that produced the molecule caprolactam [203]. On the other hand, Xu et al. determined that CTW was catalyzed using surfactants during hydrothermal carbonization. The application of Span 80 and sodium dodecyl benzenesulfonate enhanced the conversion of CTW into bio-oil. Nevertheless, Span 80 was better for the formation of pseudo-lignin, which gave the produced hydro-chars a higher energy density and improved their fuel quality and combustion performance. The majority of bio-oil yields are aromatic chemicals, acids, hydrocarbons, ketones, and esters, with yields of up to 12.79%, 11.3%, 25.4%, 23.09%, and 14.88%, respectively. In addition, cyclopentenone substances (3-methyl-2-hydroxy-2-cyclopentene-1-one and 3-ethyl-2-hydroxy-2-cyclopentene-1-one) were also found [135]. H2SO4 successfully degraded CTW through hydrolysis [204].
When recycling cotton-polyester textile waste with an organic acid catalyst, the hydrothermal process includes two steps: (i) splitting the mixture of polyester–cotton fiber waste into pieces and (ii) dispersing the mixture an organic acid catalyst in an aqueous phase to produce products. The reaction temperature was up to 140 °C under high pressure. The recycling yield of the obtained polyester fiber aggregate was 99%, and the recycling yield of the cotton fiber fragments was 81% [213]. The cotton was dissolved easily in a citric acid solution, and it was degraded in a strong acid solution. On the other hand, the cotton was unaffected by strong alkaline solutions [211]. Qi et al. used FeCl3 as a catalyst to increase cotton textile waste conversion to hydrochars, achieving highly efficient hydrophobicity by lowering the hydrothermal carbonization temperature. In addition, the simultaneous impacts of FeCl3 and hydrothermal carbonization produced a side reaction pathway by allowing more furfural compounds as derivatives [134].

6.4. Textile Recycling Using Ammonolysis

The ammonolysis process is the primary depolymerization method of nylon 6,6 carpet waste. Mckinney studied the reactions of nylon 6,6 and nylon 6,6/nylon 6 mixtures with ammonia or ammonium phosphate as a catalyst. The temperature and pressure of this process were 300–350 °C and 68 atm, respectively. The major products resulting from the depolymerization of nylon 6,6 were 5-cyanovaleramide, hexamethylenediamine, and adiponitrile. Furthermore, 6-aminocapronitrile, ε-caprolactam, and 6-aminocaproamide were the primary products of nylon 6 [214]. The depolymerization mechanism of the nylon 6,6 and nylon 6 mixture was described by Kalfas et al. The amine chain would have broken down, which would then be followed by the dehydration process and the ring addition and ring-opening reactions for the presence of cyclic lactams in nylon 6 [139].

6.5. Textile Recycling Using Gasification

Textile recycling by means of gasification processes is conducted under temperatures up to 400–1000 °C. The reactions occur under low-oxygen or oxygen-free conditions [152,215]. Compared to pyrolysis, gasification is more complicated since chemical reactions occur on the substance [216]. The gasification approach generates gases with high carbon and hydrogen concentrations, similar to the pyrolysis process. Nevertheless, the gaseous fraction in gasification is dramatically higher than that in pyrolysis [216]. The increasing pressure can lead to higher yield, calorific value, CO, and H2 concentrations when using gasification [217]. In contrast, the syngas output increases when the operating temperatures rise, whereas the char yield lowers [218]. Gasifiers can be clarified based on the reactor bed and flow type. Fixed bed and fluidized bed gasifiers are the most common gasification process. On the contrary, the spouted bed gasifier is a unique case of a fluidization type. Low gas flow rates can manage coarser solid particles, although segregation is hampered by a unique hydraulic structure [219].
Synthesis gas, often known as syngas, is the desired end product in gasification, and includes compounds such as hydrogen, carbon dioxide, carbon monoxide, and methane (CH4). The by-products of ethylene and ash can be formed [151]. Vela et al. studied the steam gasification of mixed textile waste using a fluidized bed bench-scale reactor at the reaction temperature up to 850 °C. The yields of CH4 with pure polyester, cotton, and a mixed (50% polyester and 50% cotton) were 7.0%, 3.4%, and 10.7%, respectively [220]. Furthermore, the textile wastes that were used were cotton, wool, and polyester fibers with activation energies of 144.49, 79.37, and 178.15 kJ mol−1, respectively. Char gasification took place in the temperature range of 800–1000 °C, with the primary product being CO2 [221]. On the other hand, textile waste activation energies (cotton, woolen, polyester fiber) up to 89.0 kJ/mol were achieved with a 5 wt% Fe2O3 catalyst [222].

6.6. Textile Recycling Using Glycolysis

Glycolysis is a degradation process that can be used to convert the large molecules of a textile to small molecules. This process is widely used to recycle textile materials such as PET fiber and polyurethane. PET fiber is the most widely used textile fabric in the textile industry due to its outstanding mechanical characteristics and inexpensiveness. Consequently, the generation of PET waste is also very large [223,224]. The large volume of PET fiber waste has wreaked havoc on the ecosystem. It is essential to find a sustainable and economical technique for recycling PET fiber waste to protect our resources and the environment. The glycolysis process has many advantages: it has a short reaction time and uses minimal energy. Nevertheless, PET fiber of glycolysis is slowly processed without a catalyst [225,226].
For degradation without a catalyst, it is challenging to glycolyze PET waste. A lot of effort focuses on developing a high-performance and eco-friendly catalyst to degrade the macromolecules of textile fibers to bis(2-hydroxyethyl) terephthalate (BHET) monomers and the other chemicals [227]. The recycling of polyester textiles with Mg-Al double oxides as a catalyst was studied by Guo et al. The yield of BHET was over 82 mol% using a Mg-Al double oxide catalyst with wet mixing or co-precipitation. Furthermore, the recycled PET (r-PET) fibers that were formed from the re-polymerization process had similar spinnability and mechanical characteristics to virgin PET fibers [228].
The depolymerization process of nylon 6 via the glycolysis process was investigated by Hommez et al. [229]. The glycolysis process was conducted at 250 °C in the presence of phosphoric acid. The free carboxylic acid and carboxylic acid were esterified with ethylene glycol; the major products were caprolactam, N-(5-hydroxy-3-oxa-pentyl)-caprolactam, N,N’-ethylene-di(caprolactam), and N-(2-hydroxyethyl)-caprolactam, and linear oligomers were reported. Huczkowski et al. investigated the glycolysis of nylon 6 in boiling ethylene glycol with and without a catalyst, resulting in oligoamides containing amino- and hydroxyl end-groups [230,231]. Holland and Hay presented the thermal degradation of nylon 6 and nylon 6,6 to produce a crosslinking formation and non-volatile char [232]. Kim et al. studied the glycolysis of nylon 6,6 with ethylene glycol to obtain major products from the bis(β-hydroxyethyl)adipate and β-hydroxyethylester group [233].
The glycolysis of polyester fibers is traditionally conducted in a boiling ethylene glycol solution under atmospheric pressure with several metal catalysts, e.g., zinc acetate (Zn(Ac)2). Zinc acetate can be used as a catalyst in the glycolysis process. The operation conditions for this process are 1 h and 96 °C. The PET fiber conversion and monomer yield of BHET achieved 100% and 80%, respectively. The oligomers were broken down quickly into dimers, which caused an instability process, resulting in a low molecular weight [234]. On the other hand, the glycolysis process of PET fibers to produce BHET with acetic acid, lithium hydroxide, sodium sulfate, and potassium sulfate resulted in yields of up to 64.42, 63.50, 65.72, and 64.42%, respectively [235].

6.7. Decolorization Technology in Textile Recycling

Decolorization technology is a critical problem in the high-quality chemical recycling and recovery of textile wastes. Only 1% of textile wastes, mainly white-colored wastes, are recycled, so the final color quality of regenerated fibers is uncontrollable. Color removal is required for the large-scale circulation of non-thermoplastic fibers Technologies for color removal from textile wastes are dye destruction or extraction, which take place during the pre-recycling process. Dye destruction methods, such as oxidation and photodegradation, may damage polymers and change the dyeability of the regenerated textiles [236]. To lower the chemical potential of the dyes in solution, harsh extraction conditions such as ultra-high temperature (>150 °C), lengthy time, and the use of excessive corrosive solutions can lead to structural damages to the polymers. Dye extraction focuses on finding solvents with high dye solubility [237], but dye extraction fails to remove all of the dye from textiles. Mu and Yang demonstrated the minimization of fiber density by solvents and temperatures, completely removing dispersed dyes, acid dyes, and direct dyes from PET, nylon, and cotton fibers, respectively [238].
When color removal takes place after textile degradation, the final colors are not predictable due to the uncontrollable mixture of colors in the collected textiles, for example, when cotton with multiple shades is dissolved in ionic liquids to spin regenerated cellulose fibers [239]. Li et al. used nitric acid-modified activated carbon (AC-HNO3) as an adsorbent for the removal of C.I. Disperse Red 60 (DR60) from the glycolysis products of waste PET fabrics and for the rapid decolorization of textile dyes using ceria and tin-doped ZnO nanoparticles [240,241]. Huang et al. presented ion-exchange resin (D201) to efficiently remove colorants after PET glycolysis [226]. The removal rate of the colorant and the retention efficiency of BHET was over 99% and 95%, respectively. Huang et al. presented a modified recrystallization process using ethyl acetate as the solvent needed to remove colored impurities from BHET. The process showed a decoloring rate of over 97.5% for the model-colored impurities, performing better than the reported physical adsorption processes [242].

7. Conclusions

The fashion industry faces some challenges in implementing a CE; it must consider environmental laws, policies, and the CE’s long-term viability. Digital innovation and industry 4.0 have significantly impacted CE transformation with data-driven product lifecycle analysis. By implementing IoT and RFID technology, physical production processes, such as material transportation, and the associated information flows, such as the visibility and traceability of numerous manufacturing operations, can be easily connected. Furthermore, the recognition rate of textile waste can be achieved with up to 100% accuracy using NIR technology. Mechanical and chemical recycling are common methods that take place during the textile recycling process. Textile materials such as cotton and wool can be converted into useable yarn (re-woven or knitted) by spinning carded yarn and mixed shoddy. In addition, garnering, carding, or air laying webs, which are subsequently mechanically, thermally, or chemically bonded, can be widely used to make non-woven fabrics from these materials. On the other hand, chemical treatment is seen as a promising method for textile recycling. Chemical textile recycling can be applied to degrade textile waste into feedstock monomers. Cotton waste from the textile industry has been identified as a suitable feedstock to produce cellulosic ethanol with a yield of more than 90% by means of enzymatic hydrolysis. It may be worthwhile for future experimental studies to investigate chemical recycling methods that use enzymatic hydrolysis with an ionic liquid to obtain high-value fibers. Nevertheless, the toxicity of the dissolving solvents is still a major concern; the amount of energy and water that are consumed are affected by environmental issues. Furthermore, real-time sensors to identify various types of textile materials are needed.

Author Contributions

Conceptualization, H.-S.W. and D.D.; validation, H.-S.W., D.D. and A.R.; investigation, D.D., L.A.W., A.B. and A.R.; writing—original draft preparation, D.D., L.A.W. and A.B.; writing—review and editing, D.D.; visualization, D.D.; supervision, H.-S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No applicable.

Acknowledgments

We thank the information obtained from the Far Eastern New Century Corporation Research & Development Center, Taiwan.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AIArtificial intelligence
BHETBis(2-hydroxyethyl) terephthalate
CBPConsolidated bioprocessing
CECircular economy
CIEC-Illuminant (light source C)
CNNConvolutional neural network
CTWCotton textile waste
EaActivation energy
EGEthylene Glycol
FCMFuzzy C-means
HOGHistogram of oriented gradients
ICTInformation and communication technology
IoTInternet of Things
k-NNk-nearest neighbors algorithm
LCALife cycle assessment
MLPMulti-layer perceptron
NIRNear-infrared spectroscopy
NMRNuclear magnetic resonance
PCAPrincipal component analysis
PETPolyethylene terephthalate
r-PETRecycled polyethylene terephthalate
RMBsRecycling business models
SAaaSSoftware as a service
SHFSeparated hydrolysis fermentation
SIMCASoft independent modeling of class analogy
SMFSubmerged fermentation
SMOsSmart manufacturing objects
SOMSelf-organizing map
SSFSimultaneous saccharification and fermentation
SSnFSemi-simultaneous saccharification and fermentation
SVMSupport vector machine
TILTTransform invariant low-rank textures
WEEEWaste electrical and electronic equipment

References

  1. Stone, C.; Windsor, F.M.; Munday, M.; Durance, I. Natural or synthetic–how global trends in textile usage threaten freshwater environments. Sci. Total Environ. 2020, 718, 134689. [Google Scholar] [CrossRef] [PubMed]
  2. Gardetti, M.Á. Sustainability in the textile and fashion industries: Animal ethics and welfare. In Textiles and Clothing Sustainability; Muthu, S., Ed.; Springer: Singapore, 2017; pp. 47–73. [Google Scholar]
  3. Chapagain, A.K.; Hoekstra, A.Y.; Savenije, H.H.; Gautam, R. The water footprint of cotton consumption: An assessment of the impact of worldwide consumption of cotton products on the water resources in the cotton producing countries. Ecol. Econ. 2006, 60, 186–203. [Google Scholar] [CrossRef]
  4. Pfister, S.; Bayer, P.; Koehler, A.; Hellweg, S. Environmental impacts of water use in global crop production: Hotspots and trade-offs with land use. Environ. Sci. Technol. 2011, 45, 5761–5768. [Google Scholar] [CrossRef] [PubMed]
  5. Navone, L.; Moffitt, K.; Hansen, K.-A.; Blinco, J.; Payne, A.; Speight, R. Closing the textile loop: Enzymatic fibre separation and recycling of wool/polyester fabric blends. J. Waste Manag. 2020, 102, 149–160. [Google Scholar] [CrossRef]
  6. Mäkelä, M.; Rissanen, M.; Sixta, H. Machine vision estimates the polyester content in recyclable waste textiles. Resour. Conserv. Recycl. 2020, 161, 105007. [Google Scholar] [CrossRef]
  7. MacArthur, E. A New Textiles Economy: Redesigning Fashion’s Future; Ellen MacArthur Foundation: Cowes, UK, 2017. [Google Scholar]
  8. Shirvanimoghaddam, K.; Motamed, B.; Ramakrishna, S.; Naebe, M. Death by waste: Fashion and textile circular economy case. Sci. Total Environ. 2020, 718, 137317. [Google Scholar] [CrossRef]
  9. Payne, A. Open-and closed-loop recycling of textile and apparel products. In Handbook of Life Cycle Assessment (LCA) of Textiles and Clothing; Elsevier: Amsterdam, The Netherlands, 2015; pp. 103–123. [Google Scholar]
  10. Heikkilä, P.; Cura, K.; Heikkilä, J.; Hinkka, V.; Ikonen, T.; Kamppuri, T.; Knuutila, H.; Kokko, M.; Lankiniemi, S.; Lehtinen, L. Telaketju: Towards Circularity of Textiles; VTT Research Report; No. VTT-R-00062-19; VTT Technical Research Centre of Finland: Espoo, Finland, 2019; 117p. [Google Scholar]
  11. Koszewska, M. Circular Economy—Challenges for the Textile and Clothing Industry. Autex Res. J. 2018, 18, 337–347. [Google Scholar] [CrossRef] [Green Version]
  12. Park, S.H.; Kim, S.H. Poly (ethylene terephthalate) recycling for high value added textiles. Fash. Text. 2014, 1, 1–17. [Google Scholar] [CrossRef] [Green Version]
  13. Sandin, G.; Peters, G.M. Environmental impact of textile reuse and recycling–A review. J. Clean. Prod. 2018, 184, 353–365. [Google Scholar] [CrossRef]
  14. Dahlbo, H.; Aalto, K.; Eskelinen, H.; Salmenperä, H. Increasing textile circulation—consequences and requirements. Sustain. Prod. Consum. 2017, 9, 44–57. [Google Scholar] [CrossRef]
  15. Muthu, S.S.; Li, Y.; Hu, J.Y.; Ze, L. Carbon footprint reduction in the textile process chain: Recycling of textile materials. Fibers Polym. 2012, 13, 1065–1070. [Google Scholar] [CrossRef]
  16. Franco, M.A. Circular economy at the micro level: A dynamic view of incumbents’ struggles and challenges in the textile industry. J. Clean. Prod. 2017, 168, 833–845. [Google Scholar] [CrossRef]
  17. Galvão, G.D.A.; de Nadae, J.; Clemente, D.H.; Chinen, G.; de Carvalho, M.M. Circular economy: Overview of barriers. Procedia CIRP 2018, 73, 79–85. [Google Scholar] [CrossRef]
  18. Hussain, A.; Kamboj, N.; Podgurski, V.; Antonov, M.; Goliandin, D. Circular economy approach to recycling technologies of postconsumer textile waste in Estonia: A review. Proc. Est. Acad. Sci. 2021, 70, 82–92. [Google Scholar]
  19. Manglani, H.; Hodge, G.L.; Oxenham, W. Application of the internet of things in the textile industry. Text. Prog. 2019, 51, 225–297. [Google Scholar] [CrossRef]
  20. Das, S.K.; Chinnappan, A.; Jayathilaka, W.A.D.M.; Gosh, R.; Baskar, C.; Ramakrishna, S. Challenges and potential solutions for 100% recycling of medical textiles. Mater. Circ. Econ. 2021, 3, 1–12. [Google Scholar] [CrossRef]
  21. Pagoropoulos, A.; Pigosso, D.C.; McAloone, T.C. The emergent role of digital technologies in the circular economy: A review. Procedia CIRP 2017, 64, 19–24. [Google Scholar] [CrossRef] [Green Version]
  22. Zhou, J.; Yao, X.; Zhang, J. Big data in wisdom manufacturing for industry 4.0. In Proceedings of the 2017 5th International Conference on Enterprise Systems (ES), Beijing, China, 22–24 September 2017; pp. 107–112. [Google Scholar]
  23. Fatimah, Y.A.; Govindan, K.; Murniningsih, R.; Setiawan, A. Industry 4.0 based sustainable circular economy approach for smart waste management system to achieve sustainable development goals: A case study of Indonesia. J. Clean. Prod. 2020, 269, 122263. [Google Scholar] [CrossRef]
  24. Haslinger, S.; Hietala, S.; Hummel, M.; Maunu, S.L.; Sixta, H. Solid-state NMR method for the quantification of cellulose and polyester in textile blends. Carbohydr. Polym. 2019, 207, 11–16. [Google Scholar] [CrossRef]
  25. Ruckebusch, C.; Orhan, F.; Durand, A.; Boubellouta, T.; Huvenne, J. Quantitative analysis of cotton–polyester textile blends from near-infrared spectra. Appl. Spectrosc. 2006, 60, 539–544. [Google Scholar] [CrossRef]
  26. Park, S.; Johnson, D.K.; Ishizawa, C.I.; Parilla, P.A.; Davis, M.F. Measuring the crystallinity index of cellulose by solid state 13 C nuclear magnetic resonance. Cellulose 2009, 16, 641–647. [Google Scholar] [CrossRef]
  27. Rodgers, J.; Beck, K. NIR characterization and measurement of the cotton content of dyed blend fabrics. Text. Res. J. 2009, 79, 675–686. [Google Scholar] [CrossRef]
  28. Lv, F.; Wang, C.; Zhu, P.; Zhang, C. Isolation and recovery of cellulose from waste nylon/cotton blended fabrics by 1-allyl-3-methylimidazolium chloride. Carbohydr. Polym. 2015, 123, 424–431. [Google Scholar] [CrossRef]
  29. Spathas, T. The Environmental Performance of High Value Recycling for the Fashion Industry. Master’s Thesis, Chalmers University of Technology, Gothenburg, Sweden, 2017. [Google Scholar]
  30. Rosa, P.; Sassanelli, C.; Terzi, S. Towards Circular Business Models: A systematic literature review on classification frameworks and archetypes. J. Clean. Prod. 2019, 236, 117696. [Google Scholar] [CrossRef]
  31. Winans, K.; Kendall, A.; Deng, H. The history and current applications of the circular economy concept. Renew. Sust. Energ. Rev. 2017, 68, 825–833. [Google Scholar] [CrossRef]
  32. Geissdoerfer, M.; Savaget, P.; Bocken, N.M.; Hultink, E.J. The Circular Economy–A new sustainability paradigm? J. Clean. Prod. 2017, 143, 757–768. [Google Scholar] [CrossRef] [Green Version]
  33. Kazancoglu, I.; Kazancoglu, Y.; Yarimoglu, E.; Kahraman, A. A conceptual framework for barriers of circular supply chains for sustainability in the textile industry. J. Sustain. Dev. 2020, 28, 1477–1492. [Google Scholar] [CrossRef]
  34. Allwood, J.M.; Laursen, S.E.; de Rodriguez, C.M.; Bocken, N.M. Well dressed?: The present and future sustainability of clothing and textiles in the United Kingdom. J. Home Econ. Inst. Aust. 2015, 22, 42. [Google Scholar]
  35. Manickam, P.; Duraisamy, G. 3Rs and circular economy. In Circular Economy in Textiles and Apparel; Elsevier: Amsterdam, The Netherlands, 2019; pp. 77–93. [Google Scholar]
  36. Chen, X.; Memon, H.A.; Wang, Y.; Marriam, I.; Tebyetekerwa, M. Circular Economy and sustainability of the clothing and textile Industry. Mater. Circ. Econ. 2021, 3, 1–9. [Google Scholar] [CrossRef]
  37. Balanay, R.; Halog, A. Tools for circular economy: Review and some potential applications for the Philippine textile industry. Circ. Econ. Text. Appar. 2019, 49–75. [Google Scholar] [CrossRef]
  38. Fortuna, L.M.; Diyamandoglu, V. Optimization of greenhouse gas emissions in second-hand consumer product recovery through reuse platforms. J. Waste Manag. 2017, 66, 178–189. [Google Scholar] [CrossRef] [PubMed]
  39. Lüdeke-Freund, F.; Gold, S.; Bocken, N.M. A review and typology of circular economy business model patterns. J. Ind. Ecol. 2019, 23, 36–61. [Google Scholar] [CrossRef] [Green Version]
  40. Pal, R.; Samie, Y.; Chizaryfard, A. Demystifying process-level scalability challenges in fashion remanufacturing: An interdependence perspective. J. Clean. Prod. 2021, 286, 125498. [Google Scholar] [CrossRef]
  41. Dissanayake, G.; Sinha, P. An examination of the product development process for fashion remanufacturing. Resour. Conserv. Recycl. 2015, 104, 94–102. [Google Scholar] [CrossRef]
  42. Niinimäki, K.; Hassi, L. Emerging design strategies in sustainable production and consumption of textiles and clothing. J. Clean. Prod. 2011, 19, 1876–1883. [Google Scholar] [CrossRef]
  43. Gwilt, A.; Rissanen, T. Shaping Sustainable Fashion: Changing the Way We Make and Use Clothes; Routledge: England, UK, 2012. [Google Scholar]
  44. Wen-hui, X.; Dian-yan, J.; Yu-ying, H. The remanufacturing reverse logistics management based on closed-loop supply chain management processes. Procedia Environ. Sci. 2011, 11, 351–354. [Google Scholar] [CrossRef]
  45. Vogtlander, J.G.; Scheepens, A.E.; Bocken, N.M.; Peck, D. Combined analyses of costs, market value and eco-costs in circular business models: Eco-efficient value creation in remanufacturing. J. Remanufacturing 2017, 7, 1–17. [Google Scholar] [CrossRef] [Green Version]
  46. De Silva, R.; Wang, X.; Byrne, N. Recycling textiles: The use of ionic liquids in the separation of cotton polyester blends. RSC Adv. 2014, 4, 29094–29098. [Google Scholar] [CrossRef]
  47. Li, X.; Hu, Y.; Du, C.; Lin, C.S.K. Recovery of glucose and polyester from textile waste by enzymatic hydrolysis. Waste Biomass Valorization 2019, 10, 3763–3772. [Google Scholar] [CrossRef] [Green Version]
  48. Kumar, P.S.; Yaashikaa, P. Recycled Fibres. In Sustainable Innovations in Recycled Textiles; Springer: Amsterdam, The Netherlands, 2018; pp. 1–7. [Google Scholar]
  49. Yin, Y.; Yao, D.; Wang, C.; Wang, Y. Removal of spandex from nylon/spandex blended fabrics by selective polymer degradation. Text. Res. J. 2014, 84, 16–27. [Google Scholar] [CrossRef]
  50. Huang, Y.-F.; Azevedo, S.G.; Lin, T.-J.; Cheng, C.-S.; Lin, C.-T. Exploring the decisive barriers to achieve circular economy: Strategies for the textile innovation in Taiwan. Sustain. Prod. Consum. 2021, 27, 1406–1423. [Google Scholar] [CrossRef]
  51. Martina, R.A.; Oskam, I.F. Practical guidelines for designing recycling, collaborative, and scalable business models: A case study of reusing textile fibers into biocomposite products. J. Clean. Prod. 2021, 128542. [Google Scholar] [CrossRef]
  52. Christensen, T.B. Towards a circular economy in cities: Exploring local modes of governance in the transition towards a circular economy in construction and textile recycling. J. Clean. Prod. 2021, 305, 127058. [Google Scholar] [CrossRef]
  53. Seadon, J.K. Sustainable waste management systems. J. Clean. Prod. 2010, 18, 1639–1651. [Google Scholar] [CrossRef]
  54. Payne, A.; Nay, Z.; Maguire, R. Regulating a circular economy for textiles in Australia. In Proceedings of the 4th PLATE 2021 Virtual Conference, Limerick, Ireland, 26–28 May 2021. [Google Scholar]
  55. Ghoreishi, M.; Happonen, A. The case of fabric and textile industry: The emerging role of digitalization, internet-of-Things and industry 4.0 for circularity. In Proceedings of the Sixth International Congress on Information and Communication Technology, London, UK, 25–26 February 2022; pp. 189–200. [Google Scholar]
  56. Abdullah, N.; Alwesabi, O.A.; Abdullah, R. Iot-based smart waste management system in a smart city. In Proceedings of the International Conference of Reliable Information and Communication Technology, Kuala Lumpur, Malaysia, 23–24 June 2018; pp. 364–371. [Google Scholar]
  57. Zhong, R.Y.; Dai, Q.; Qu, T.; Hu, G.; Huang, G.Q. RFID-enabled real-time manufacturing execution system for mass-customization production. Robot. Comput. Integr. Manuf. 2013, 29, 283–292. [Google Scholar] [CrossRef]
  58. Lu, B.; Bateman, R.; Cheng, K. RFID enabled manufacturing: Fundamentals, methodology and applications. Int. J. Agil. Syst. Manag. 2006, 1, 73–92. [Google Scholar] [CrossRef]
  59. Zhong, R.Y.; Li, Z.; Pang, L.; Pan, Y.; Qu, T.; Huang, G.Q. RFID-enabled real-time advanced planning and scheduling shell for production decision making. Int. J. Comput. Integr. Manuf. 2013, 26, 649–662. [Google Scholar] [CrossRef]
  60. Mavropoulos, A.; Nilsen, A.W. Industry 4.0 and Circular Economy: Towards a Wasteless Future or a Wasteful Planet? John Wiley & Sons: New Jersey, NJ, USA, 2020. [Google Scholar]
  61. Zhong, R.Y.; Xu, X.; Klotz, E.; Newman, S.T. Intelligent manufacturing in the context of industry 4.0: A review. Engineering 2017, 3, 616–630. [Google Scholar] [CrossRef]
  62. Kumar, K.; Zindani, D.; Davim, J.P. Intelligent Manufacturing. In Industry 4.0; Springer: Amsterdam, The Netherlands, 2019; pp. 1–17. [Google Scholar]
  63. Wu, D.; Greer, M.J.; Rosen, D.W.; Schaefer, D. Cloud manufacturing: Strategic vision and state-of-the-art. J. Manuf. Syst. 2013, 32, 564–579. [Google Scholar] [CrossRef] [Green Version]
  64. Lopes, S.; Machado, S. IoT based automatic waste segregator. In Proceedings of the 2019 International Conference on Advances in Computing, Communication and Control (ICAC3), Mumbai, India, 20–21 December 2019; pp. 1–5. [Google Scholar]
  65. Yu, K.H.; Zhang, Y.; Li, D.; Montenegro-Marin, C.E.; Kumar, P.M. Environmental planning based on reduce, reuse, recycle and recover using artificial intelligence. Environ. Impact Assess. Rev. 2021, 86, 106492. [Google Scholar] [CrossRef]
  66. Giacobbe, M.; Puliafito, C.; Scarpa, M. The big bucket: An iot cloud solution for smart waste management in smart cities. In Proceedings of the European Conference on Service-Oriented and Cloud Computing, Vienna, Austria, 5–7 September 2016; pp. 43–58. [Google Scholar]
  67. Merlino, G.; Bruneo, D.; Distefano, S.; Longo, F.; Puliafito, A. Stack4things: Integrating iot with openstack in a smart city context. In Proceedings of the 2014 International Conference on Smart Computing Workshops, Hongkong, China, 5 November 2014; pp. 21–28. [Google Scholar]
  68. Marin Perez, M. Analysis of European Post-Consumer Textile Waste for Automated Sorting. Master’s Thesis, Uppsala University, Stockholm, Sweden, 2021. [Google Scholar]
  69. Bergfjord, C.; Holst, B. A procedure for identifying textile bast fibres using microscopy: Flax, nettle/ramie, hemp and jute. Ultramicroscopy 2010, 110, 1192–1197. [Google Scholar] [CrossRef]
  70. Nørup, N.; Pihl, K.; Damgaard, A.; Scheutz, C. Development and testing of a sorting and quality assessment method for textile waste. J. Waste Manag. 2018, 79, 8–21. [Google Scholar] [CrossRef] [PubMed]
  71. Choudhuri, P.K.; Majumdar, P.K.; Sarkar, B. Thermal behaviour of textiles: A review. Man-Made Text. India. 2013, 41, 3. [Google Scholar]
  72. Sombatsompop, N.; Wood, A. Measurement of thermal conductivity of polymers using an improved Lee’s disc apparatus. Polym. Test. 1997, 16, 203–223. [Google Scholar] [CrossRef]
  73. Yousef, S.; Tatariants, M.; Tichonovas, M.; Sarwar, Z.; Jonuškienė, I.; Kliucininkas, L. A new strategy for using textile waste as a sustainable source of recovered cotton. Resour. Conserv. Recycl. 2019, 145, 359–369. [Google Scholar] [CrossRef]
  74. Liu, Z.; Li, W.; Wei, Z. Qualitative classification of waste textiles based on near infrared spectroscopy and the convolutional network. Text. Res. J. 2020, 90, 1057–1066. [Google Scholar] [CrossRef]
  75. Zitting, J. Optical Sorting Technology for Textile Waste: Development of an Identification Method with NIR Spectroscopy. Bechelor’s Thesis, Lahti University, Lahti, Finland, 2017. [Google Scholar]
  76. Wilts, H.; Garcia, B.R.; Garlito, R.G.; Gómez, L.S.; Prieto, E.G. Artificial intelligence in the sorting of municipal waste as an enabler of the circular economy. Resources 2021, 10, 28. [Google Scholar] [CrossRef]
  77. Xiong, E. Autonomous sorting of plastic resin using near-infrared and machine learning. CSFJ. 2021, 3, 1–5. [Google Scholar]
  78. Berghmans, A.C.; Huys, M.J.G. Progress for Identifying Objects Using an Optical Spectrometer and a Transport System. U.S. Patent 7,071,469, 4 July 2006. [Google Scholar]
  79. Zhou, C.; Han, G.; Via, B.K.; Song, Y.; Gao, S.; Jiang, W. Rapid identification of fibers from different waste fabrics using the near-infrared spectroscopy technique. Text. Res. J. 2019, 89, 3610–3616. [Google Scholar] [CrossRef]
  80. Zhou, J.; Yu, L.; Ding, Q.; Wang, R. Textile fiber identification using near-infrared spectroscopy and pattern recognition. Autex Res. J. 2019, 19, 201–209. [Google Scholar] [CrossRef] [Green Version]
  81. Chen, H.; Tan, C.; Lin, Z. Quantitative determination of wool in textile by near-infrared spectroscopy and multivariate models. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2018, 201, 229–235. [Google Scholar] [CrossRef] [PubMed]
  82. Iqbal Hussain, M.A.; Khan, B.; Wang, Z.; Ding, S. Woven fabric pattern recognition and classification based on deep convolutional neural networks. Electronics 2020, 9, 1048. [Google Scholar] [CrossRef]
  83. Xiao, Z.; Guo, Y.; Geng, L.; Wu, J.; Zhang, F.; Wang, W.; Liu, Y. Automatic recognition of woven fabric pattern based on TILT. Math. Probl. Eng. 2018, 2018, 9707104. [Google Scholar] [CrossRef]
  84. Kuo, C.-F.J.; Kao, C.-Y. Self-organizing map network for automatically recognizing color texture fabric nature. Fibers Polym. 2007, 8, 174–180. [Google Scholar] [CrossRef]
  85. Cura, K.; Rintala, N.; Kamppuri, T.; Saarimäki, E.; Heikkilä, P. Textile recognition and sorting for recycling at an automated line using near infrared spectroscopy. Recycling 2021, 6, 11. [Google Scholar] [CrossRef]
  86. Riba, J.-R.; Cantero, R.; Canals, T.; Puig, R. Circular economy of post-consumer textile waste: Classification through infrared spectroscopy. J. Clean. Prod. 2020, 272, 123011. [Google Scholar] [CrossRef]
  87. Peets, P.; Kaupmees, K.; Vahur, S.; Leito, I. Reflectance FT-IR spectroscopy as a viable option for textile fiber identification. Herit. Sci. 2019, 7, 1–10. [Google Scholar] [CrossRef]
  88. Morgado, V.; Gomes, L.; Bettencourt da Silva, R.J.N.; Palma, C. Validated spreadsheet for the identification of PE, PET, PP and PS microplastics by micro-ATR-FTIR spectra with known uncertainty. Talanta 2021, 234, 122624. [Google Scholar] [CrossRef]
  89. Gulich, B. Development of products made of reclaimed fibres. Recycl. Text. Camb. 2006, 1, 117–136. [Google Scholar]
  90. Valerio, O.; Muthuraj, R.; Codou, A. Strategies for polymer to polymer recycling from waste: Current trends and opportunities for improving the circular economy of polymers in South America. Curr. Opin. Green Sustain. Chem. 2020, 25, 100381. [Google Scholar] [CrossRef]
  91. Michaud, J.-C.; Farrant, L.; Jan, O.; Kjær, B.; Bakas, I. Environmental Benefits of Recycling—2010 Update. Material Change for a Better Environment. 2010. Available online: http://localhost:8080/xmlui/handle/123456789/174 (accessed on 22 August 2021).
  92. Euratex. Circular Textiles, Prospering in the Circular Economy. Available online: https://euratex.eu/wp-content/uploads/EURATEX-Prospering-in-the-Circular-Economy-2020.pdf (accessed on 22 August 2021).
  93. Pensupa, N.; Leu, S.-Y.; Hu, Y.; Du, C.; Liu, H.; Jing, H.; Wang, H.; Lin, C.S.K. Recent trends in sustainable textile waste recycling methods: Current situation and future prospects. Chem. Chem. Technol. Waste Valoriz. 2017, 189–228. [Google Scholar] [CrossRef]
  94. Bartlett, C.; McGill, I.; Willis, P. Textiles Flow and Market Development Opportunities in the UK; Waste & Resources Action Programme: Banbury, UK, 2013. [Google Scholar]
  95. Korhonen, M.-R.; Dahlbo, H. Reducing Greenhouse Gas Emissions by Recycling Plastics and Textiles into Products; Finnish Environment Institute: Helsinki, Finland, 2007. [Google Scholar]
  96. Johansson, L. On the Mechanical Recycling of Woven Fabrics: Improving the Reusable Fibre Yield of Mechanical Methods. Master’s Thesis, Uppala University, Stockholm, Sweden, 2020. [Google Scholar]
  97. Hutten, I.M. Handbook of Nonwoven Filter Media; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  98. Jin, K.; Banerji, A.; Kitto, D.; Bates, F.S.; Ellison, C.J. Mechanically robust and recyclable cross-linked fibers from melt blown anthracene-functionalized commodity polymers. ACS Appl. Mater. Interfaces 2019, 11, 12863–12870. [Google Scholar] [CrossRef] [PubMed]
  99. Tsai, P.P.-Y. Methods of Saturating Nonwoven Fabrics with Liquid and the Making of Electret Therof. U.S. Patent 17/056,276, 26 August 2021. [Google Scholar]
  100. Mondragon, G.; Kortaberria, G.; Mendiburu, E.; González, N.; Arbelaiz, A.; Peña-Rodriguez, C. Thermomechanical recycling of polyamide 6 from fishing nets waste. J. Appl. Polym. Sci. 2020, 137, 48442. [Google Scholar] [CrossRef]
  101. Russell, S.; Swan, P.; Trebowicz, M.; Ireland, A. Review of wool recycling and reuse. In Natural Fibres: Advances in Science and Technology towards Industrial Applications; Springer: Dordrecht, The Netherlands, 2016; pp. 415–428. [Google Scholar] [CrossRef]
  102. Alexander, C.; Reno, J. Economies of Recycling: The Global Transformation of Materials, Values and Social Relations; Zed Books Ltd.: London, UK, 2012. [Google Scholar]
  103. Klepp, I.G.; Tobiasson, T.S.; Laitala, K. Wool as an Heirloom: How Natural Fibres Can Reinvent Value in Terms of Money, Life-Span and Love. In Natural Fibres: Advances in Science and Technology towards Industrial Applications; Springer: Dordrecht, The Netherlands, 2016. [Google Scholar] [CrossRef]
  104. Harmsen, P.; Scheffer, M.; Bos, H. Textiles for circular fashion: The logic behind recycling options. Sustainability 2021, 13, 9714. [Google Scholar] [CrossRef]
  105. Klaus-nietrost, C.; Richard, H.; Weilach, C. Method of Reusing a Mixed Textile Comprising Cellulose and Synthetic Plastic. U.S. Patent 16/962,498, 18 March 2021. [Google Scholar]
  106. Piribauer, B.; Bartl, A. Textile recycling processes, state of the art and current developments: A mini review. Waste Manag. Res. 2019, 37, 112–119. [Google Scholar] [CrossRef]
  107. Liu, R.-G.; Shen, Y.-Y.; Shao, H.-L.; Wu, C.-X.; Hu, X.-C. An analysis of Lyocell fiber formation as a melt–spinning process. Cellulose 2001, 8, 13–21. [Google Scholar] [CrossRef]
  108. Kaminsky, W.; Scheirs, J. Feedstock Recycling and Pyrolysis of Waste Plastics: Converting Waste Plastics into Diesel and Other Fuels; J. Wiley & Sons: New Jersey, NJ, USA, 2006. [Google Scholar]
  109. Damayanti, D.; Wu, H.-S. Strategic possibility routes of recycled PET. Polymers 2021, 13, 1475. [Google Scholar] [CrossRef]
  110. Ambercycle. Creating a Circular System. Available online: www.ambercycle.com (accessed on 29 July 2021).
  111. BlockTexx. Textile Recovery Technology. Available online: https://www.blocktexx.com (accessed on 29 July 2021).
  112. FENC. New Process Recycles any PET Waste. Available online: http://news.fenc.com/news_detail.aspx?lang=en&id=4557 (accessed on 4 March 2021).
  113. Fiber, I. Infinited Fiber Company Brochure. Available online: https://infinitedfiber.com/wp-content/uploads/2019/06/IFC_Brochure_5-2019.pdf (accessed on 29 July 2021).
  114. Asikainen, S.; Määttänen, M.; Harlin, A.; Valta, K.; Sivonen, E. Method of Producing Dissolving Pulp, Dissolving Pulp and Use of Method. U.S. Patent 14/428,377, 13 August 2015. [Google Scholar]
  115. Ioncell. Ioncell® Process. Available online: https://ioncell.fi/research/ (accessed on 30 July 2021).
  116. Lenzing. Closing the Loops. Available online: https://www.lenzing.com/sustainability/production/resources/chemicals (accessed on 4 September 2021).
  117. BioSciences, T. Tyton BioSciences. Available online: www.tytonbio.com (accessed on 4 September 2021).
  118. Technologies, W.A. Available online: https://wornagain.co.uk/ (accessed on 22 May 2021).
  119. Sherwood, J. Closed-loop recycling of polymers using solvents. Johnson Matthey Technol. Rev. 2020, 4–15. [Google Scholar] [CrossRef]
  120. Ribul, M.; Lanot, A.; Tommencioni-Pisapia, C.; Purnell, P.; McQueen-Mason, S.J.; Baurley, S. Mechanical, chemical, biological: Moving towards closed-loop bio-based recycling in a circular economy of sustainable textiles. J. Clean. Prod. 2021, 129325. [Google Scholar] [CrossRef]
  121. Saha, S. Textile Recycling: The Chemical Recycling Process of Textiles. Available online: https://www.onlineclothingstudy.com/2020/08/textile-recycling-chemical-recycling.html (accessed on 24 October 2021).
  122. Coates, G.W.; Getzler, Y.D. Chemical recycling to monomer for an ideal, circular polymer economy. Nat. Rev. Mater. 2020, 5, 501–516. [Google Scholar] [CrossRef]
  123. Bruce, M.; Miller, W.D. Nylon Hydrolisis. U.S. Patent 2,840,606, 24 June 1958. [Google Scholar]
  124. Bodrero, S.; Canivenc, E.; Cansell, F. Chemical Recycling of Polyamide 6.6 and Polyamide 6 through a Two Step Ami-/Ammonolysis Process; Georgia Institute of Technology: Atlanta, GA, USA, 1999. [Google Scholar]
  125. Polk, M.B.; Leboeuf, L.L.; Shah, M.; Won, C.-Y.; Hu, X.; Ding, W. Nylon 66, nylon 46, and PET phase-transfer-catalyzed alkaline depolymerization at atmospheric pressure. Polym. Plast. Technol. Eng. 1999, 38, 459–470. [Google Scholar] [CrossRef]
  126. Rickson, R. Controlling sediment at source: An evaluation of erosion control geotextiles. Earth Surf. Process.Land. BGRG 2006, 31, 550–560. [Google Scholar] [CrossRef]
  127. Broda, J.; Przybyło, S.; Gawłowski, A.; Grzybowska-Pietras, J.; Sarna, E.; Rom, M.; Laszczak, R. Utilisation of textile wastes for the production of geotextiles designed for erosion protection. J. Text. Inst. 2019, 110, 435–444. [Google Scholar] [CrossRef]
  128. Yang, Z.; Li, X.; Si, J.; Cui, Z.; Peng, K. Morphological, mechanical and thermal properties of poly (lactic acid)(PLA)/cellulose nanofibrils (CNF) composites nanofiber for tissue engineering. J. Wuhan Univ. Technol. Mater. Sci. Ed. 2019, 34, 207–215. [Google Scholar] [CrossRef]
  129. Rizal, S.; Olaiya, F.G.; Saharudin, N.; Abdullah, C.; Ng, O.; Mohamad Haafiz, M.; Yahya, E.B.; Sabaruddin, F.; Khalil, H. Isolation of textile waste cellulose nanofibrillated fibre reinforced in polylactic acid-chitin biodegradable composite for green packaging application. Polymers 2021, 13, 325. [Google Scholar] [CrossRef]
  130. Barisci, S.; Oncel, M.S. The disposal of combed cotton wastes by pyrolysis. Int. J. Green Energy 2014, 11, 255–266. [Google Scholar] [CrossRef]
  131. Nunes, L.J.R.; Godina, R.; Matias, J.C.O.; Catalão, J.P.S. Economic and environmental benefits of using textile waste for the production of thermal energy. J. Clean. Prod. 2018, 171, 1353–1360. [Google Scholar] [CrossRef]
  132. Czajczyńska, D.; Anguilano, L.; Ghazal, H.; Krzyżyńska, R.; Reynolds, A.; Spencer, N.; Jouhara, H. Potential of pyrolysis processes in the waste management sector. Therm. Sci. Eng. Prog. 2017, 3, 171–197. [Google Scholar] [CrossRef]
  133. Czajczyńska, D.; Nannou, T.; Anguilano, L.; Krzyżyńska, R.; Ghazal, H.; Spencer, N.; Jouhara, H. Potentials of pyrolysis processes in the waste management sector. Energy Procedia 2017, 123, 387–394. [Google Scholar] [CrossRef]
  134. Qi, R.; Xu, Z.; Zhou, Y.; Zhang, D.; Sun, Z.; Chen, W.; Xiong, M. Clean solid fuel produced from cotton textiles waste through hydrothermal carbonization with FeCl3: Upgrading the fuel quality and combustion characteristics. Energy 2021, 214, 118926. [Google Scholar] [CrossRef]
  135. Xu, Z.; Qi, R.; Xiong, M.; Zhang, D.; Gu, H.; Chen, W. Conversion of cotton textile waste to clean solid fuel via surfactant-assisted hydrothermal carbonization: Mechanisms and combustion behaviors. Bioresour. Technol. 2021, 321, 124450. [Google Scholar] [CrossRef]
  136. Duman, G. Preparation of novel porous carbon from hydrothermal pretreated textile wastes: Effects of textile type and activation agent on structural and adsorptive properties. J. Water Process. Eng. 2021, 43, 102286. [Google Scholar] [CrossRef]
  137. Yasin, S.; Massimo, C.; Rovero, G.; Behary, N.; Perwuelz, A.; Giraud, S.; Migliavacca, G.; Chen, G.; Guan, J. An alternative for the end-of-life phase of flame retardant textile products: Degradation of flame retardant and preliminary settings of energy valorization by gasification. BioResources 2017, 12, 5196–5211. Available online: https://ojs.cnr.ncsu.edu/index.php/BioRes/article/viewFile/BioRes_12_3_5196_Yasin_Alternative_Flame_Retardant_Textile/5343 (accessed on 29 July 2021). [CrossRef]
  138. Sert, E.; Yılmaz, E.; Atalay, F.S. Chemical recycling of polyethlylene terephthalate by glycolysis using deep eutectic solvents. J. Polym. Environ. 2019, 27, 2956–2962. [Google Scholar] [CrossRef]
  139. Kalfas, G.A. Mathematical modeling of the depolymerization of polyamide mixtures-Part I: Kinetic mechanism and parametric studies in batch reactors. Polym. React. Eng. 1998, 6, 41–67. [Google Scholar] [CrossRef]
  140. Kwon, E.E.; Lee, T.; Ok, Y.S.; Tsang, D.C.; Park, C.; Lee, J. Effects of calcium carbonate on pyrolysis of sewage sludge. Energy 2018, 153, 726–731. [Google Scholar] [CrossRef]
  141. Lee, T.; Nam, I.-H.; Jung, S.; Park, Y.-K.; Kwon, E.E. Synthesis of nickel/biochar composite from pyrolysis of Microcystis aeruginosa and its practical use for syngas production. Bioresour. Technol. 2020, 300, 122712. [Google Scholar] [CrossRef]
  142. Damayanti, D.; Wu, H.S. Pyrolysis kinetic of alkaline and dealkaline lignin using catalyst. J. Polym. Res. 2018, 25, 7. [Google Scholar] [CrossRef]
  143. Miranda, R.; Sosa_Blanco, C.; Bustos-Martinez, D.; Vasile, C. Pyrolysis of textile wastes: I. Kinetics and yields. J. Anal. Appl. Pyrolysis 2007, 80, 489–495. [Google Scholar] [CrossRef]
  144. Lee, J.; Kim, K.-H.; Kwon, E.E. Biochar as a catalyst. Renew. Sust. Energ. Rev. 2017, 77, 70–79. [Google Scholar] [CrossRef]
  145. You, S.; Ok, Y.S.; Chen, S.S.; Tsang, D.C.; Kwon, E.E.; Lee, J.; Wang, C.-H. A critical review on sustainable biochar system through gasification: Energy and environmental applications. Bioresour. Technol. 2017, 246, 242–253. [Google Scholar] [CrossRef] [Green Version]
  146. Kwon, D.; Yi, S.; Jung, S.; Kwon, E.E. Valorization of synthetic textile waste using CO2 as a raw material in the catalytic pyrolysis process. Environ. Pollut. 2021, 268, 115916. [Google Scholar] [CrossRef]
  147. Yousef, S.; Eimontas, J.; Striūgas, N.; Trofimov, E.; Hamdy, M.; Abdelnaby, M.A. Conversion of end-of-life cotton banknotes into liquid fuel using mini-pyrolysis plant. J. Clean. Prod. 2020, 267, 121612. [Google Scholar] [CrossRef]
  148. Wang, S.; Dai, G.; Yang, H.; Luo, Z. Lignocellulosic biomass pyrolysis mechanism: A state-of-the-art review. Prog. Energy Combust. Sci. 2017, 62, 33–86. [Google Scholar] [CrossRef]
  149. Damayanti, D.; Wulandari, Y.R.; Wu, H.-S. Product Distribution of Chemical Product Using Catalytic Depolymerization of Lignin. Bull. Chem. React. Eng. Catal. 2020, 15, 432–453. [Google Scholar] [CrossRef]
  150. Molino, A.; Chianese, S.; Musmarra, D. Biomass gasification technology: The state of the art overview. J. Energy Chem. 2016, 25, 10–25. [Google Scholar] [CrossRef]
  151. Devi, L.; Ptasinski, K.J.; Janssen, F.J. A review of the primary measures for tar elimination in biomass gasification processes. Biomass Bioenerg. 2003, 24, 125–140. [Google Scholar] [CrossRef]
  152. Rittfors, J. Thermochemical Textile Recycling: Investigation of Pyrolysis and Gasification of Cotton and Polyester; Chalmers tekniska högskola: Göteborg, Sweden, 2020. [Google Scholar]
  153. Yousef, S.; Kalpokaitė-Dičkuvienė, R.; Baltušnikas, A.; Pitak, I.; Lukošiūtė, S.-I. A new strategy for functionalization of char derived from pyrolysis of textile waste and its application as hybrid fillers (CNTs/char and graphene/char) in cement industry. J. Clean. Prod. 2021, 128058. [Google Scholar] [CrossRef]
  154. Jagdale, P.; Nair, J.R.; Khan, A.; Armandi, M.; Meligrana, G.; Hernandez, F.R.; Rusakova, I.; Piatti, E.; Rovere, M.; Tagliaferro, A. Waste to life: Low-cost, self-standing, 2D carbon fiber green Li-ion battery anode made from end-of-life cotton textile. Electrochim. Acta 2021, 368, 137644. [Google Scholar] [CrossRef]
  155. Williams, P.T.; Reed, A.R. Pre-formed activated carbon matting derived from the pyrolysis of biomass natural fibre textile waste. J. Anal. Appl. Pyrolysis 2003, 70, 563–577. [Google Scholar] [CrossRef]
  156. Williams, P.T.; Reed, A.R. High grade activated carbon matting derived from the chemical activation and pyrolysis of natural fibre textile waste. J. Anal. Appl. Pyrolysis 2004, 71, 971–986. [Google Scholar] [CrossRef]
  157. Reed, A.R.; Williams, P.T. Thermal processing of biomass natural fibre wastes by pyrolysis. Int. J. Energy Res. 2004, 28, 131–145. [Google Scholar] [CrossRef]
  158. Xu, Z.; Gu, S.; Sun, Z.; Zhang, D.; Zhou, Y.; Gao, Y.; Qi, R.; Chen, W. Synthesis of char-based adsorbents from cotton textile waste assisted by iron salts at low pyrolysis temperature for Cr (VI) removal. Environ. Sci. Pollut. Res. 2020, 1–14. [Google Scholar] [CrossRef] [PubMed]
  159. Eimontas, J.; Yousef, S.; Striūgas, N.; Abdelnaby, M.A. Catalytic pyrolysis kinetic behaviour and TG-FTIR-GC–MS analysis of waste fishing nets over ZSM-5 zeolite catalyst for caprolactam recovery. Renew. Energy 2021, 179, 1385–1403. [Google Scholar] [CrossRef]
  160. da Silva, J.E.; Calixto, G.Q.; de Araújo Medeiros, R.L.B.; de Freitas Melo, M.A.; de Araújo Melo, D.M.; de Carvalho, L.P.; Braga, R.M. Colored cotton wastes valuation through thermal and catalytic reforming of pyrolysis vapors (Py-GC/MS). Sci. Rep. 2021, 11, 1–10. [Google Scholar] [CrossRef] [PubMed]
  161. Wang, J.; Jiang, J.; Ding, J.; Wang, X.; Sun, Y.; Ruan, R.; Ragauskas, A.J.; Ok, Y.S.; Tsang, D.C. Promoting Diels-Alder reactions to produce bio-BTX: Co-aromatization of textile waste and plastic waste over USY zeolite. J. Clean. Prod. 2021, 127966. [Google Scholar] [CrossRef]
  162. Yousef, S.; Eimontas, J.; Striūgas, N.; Mohamed, A.; Abdelnaby, M.A. Morphology, compositions, thermal behavior and kinetics of pyrolysis of lint-microfibers generated from clothes dryer. J. Anal. Appl. Pyrolysis 2021, 155, 105037. [Google Scholar] [CrossRef]
  163. Gholamzad, E.; Karimi, K.; Masoomi, M. Effective conversion of waste polyester–cotton textile to ethanol and recovery of polyester by alkaline pretreatment. Chem. Eng. Sci. 2014, 253, 40–45. [Google Scholar] [CrossRef]
  164. Subramanian, K.; Sarkar, M.K.; Wang, H.; Qin, Z.-H.; Chopra, S.S.; Jin, M.; Kumar, V.; Chen, C.; Tsang, C.-W.; Lin, C.S.K. An overview of cotton and polyester, and their blended waste textile valorisation to value-added products: A circular economy approach–research trends, opportunities and challenges. Crit. Rev. Environ. Sci. Technol. 2021, 1–22. [Google Scholar] [CrossRef]
  165. Nikolić, S.; Lazić, V.; Veljović, Đ.; Mojović, L. Production of bioethanol from pre-treated cotton fabrics and waste cotton materials. Carbohydr. Polym. 2017, 164, 136–144. [Google Scholar] [CrossRef]
  166. Yang, T.C.; Kumaran, J.; Amartey, S.; Maki, M.; Li, X.; Lu, F.; Qin, W. Biofuels and bioproducts produced through microbial conversion of biomass. Bioenergy Res. Adv. Appl. 2014, 71–93. [Google Scholar] [CrossRef]
  167. Elsayed, S.; Hellsten, S.; Guizani, C.; Witos, J.; Rissanen, M.; Rantamäki, A.H.; Varis, P.; Wiedmer, S.K.; Sixta, H. Recycling of superbase-based ionic liquid solvents for the production of textile-grade regenerated cellulose fibers in the lyocell process. ACS Sustain. Chem. Eng. 2020, 8, 14217–14227. [Google Scholar] [CrossRef]
  168. Hong, F.; Guo, X.; Zhang, S.; Han, S.-f.; Yang, G.; Jönsson, L.J. Bacterial cellulose production from cotton-based waste textiles: Enzymatic saccharification enhanced by ionic liquid pretreatment. Bioresour. Technol. 2012, 104, 503–508. [Google Scholar] [CrossRef] [PubMed]
  169. Sasaki, C.; Nakagawa, T.; Asada, C.; Nakamura, Y. Microwave-assisted hydrolysis of cotton waste to glucose in combination with the concentrated sulfuric acid impregnation method. Waste Biomass Valoriz. 2020, 11, 4279–4287. [Google Scholar] [CrossRef]
  170. Hasanzadeh, E.; Mirmohamadsadeghi, S.; Karimi, K. Enhancing energy production from waste textile by hydrolysis of synthetic parts. Fuel 2018, 218, 41–48. [Google Scholar] [CrossRef]
  171. Keskin, T.; Abubackar, H.N.; Arslan, K.; Azbar, N. Biohydrogen production from solid wastes. In Biohydrogen; Elsevier: Amsterdam, The Netherlands, 2019; pp. 321–346. [Google Scholar]
  172. Dimos, K.; Paschos, T.; Louloudi, A.; Kalogiannis, K.G.; Lappas, A.A.; Papayannakos, N.; Kekos, D.; Mamma, D. Effect of various pretreatment methods on bioethanol production from cotton stalks. Fermentation 2019, 5, 5. [Google Scholar] [CrossRef] [Green Version]
  173. Chandrashekhar, B.; Mishra, M.S.; Sharma, K.; Dubey, S. Bio-ethanol production from textile cotton waste via dilute acid hydrolysis and fermentation by Saccharomyces cerevisiae. J. Ecobiotechnol. 2011, 3, 6–9. [Google Scholar]
  174. Binczarski, M.J.; Malinowska, J.; Stanishevsky, A.; Severino, C.J.; Yager, R.; Cieslak, M.; Witonska, I.A. A Model Procedure for Catalytic Conversion of Waste Cotton into Useful Chemicals. Materials 2021, 14, 1981. [Google Scholar] [CrossRef] [PubMed]
  175. Quartinello, F.; Vecchiato, S.; Weinberger, S.; Kremenser, K.; Skopek, L.; Pellis, A.; Guebitz, G.M. Highly selective enzymatic recovery of building blocks from wool-cotton-polyester textile waste blends. Polymers 2018, 10, 1107. [Google Scholar] [CrossRef] [Green Version]
  176. Shen, F.; Xiao, W.; Lin, L.; Yang, G.; Zhang, Y.; Deng, S. Enzymatic saccharification coupling with polyester recovery from cotton-based waste textiles by phosphoric acid pretreatment. Bioresour. Technol. 2013, 130, 248–255. [Google Scholar] [CrossRef]
  177. Kuo, C.H.; Lin, P.J.; Lee, C.K. Enzymatic saccharification of dissolution pretreated waste cellulosic fabrics for bacterial cellulose production by Gluconacetobacter xylinus. J. Chem. Technol. Biotechnol. 2010, 85, 1346–1352. [Google Scholar] [CrossRef]
  178. Yousef, S.; Kuliešienė, N.; Sakalauskaitė, S.; Nenartavičius, T.; Daugelavičius, R. Sustainable green strategy for recovery of glucose from end-of-life euro banknotes. Waste Manag. 2021, 123, 23–32. [Google Scholar] [CrossRef]
  179. Piribauer, B.; Bartl, A.; Ipsmiller, W. Enzymatic textile recycling–best practices and outlook. Waste Manag. Res. 2021, 39, 1277–1290. [Google Scholar] [CrossRef] [PubMed]
  180. Zhu, L.; Pelaz, B.; Chakraborty, I.; Parak, W.J. Investigating possible enzymatic degradation on polymer shells around inorganic nanoparticles. Int. J. Mol. Sci. 2019, 20, 935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Feliu, N.; Docter, D.; Heine, M.; Del Pino, P.; Ashraf, S.; Kolosnjaj-Tabi, J.; Macchiarini, P.; Nielsen, P.; Alloyeau, D.; Gazeau, F. In vivo degeneration and the fate of inorganic nanoparticles. Chem. Soc. Rev. 2016, 45, 2440–2457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Soenen, S.J.; Himmelreich, U.; Nuytten, N.; Pisanic, T.R.; Ferrari, A.; De Cuyper, M. Intracellular nanoparticle coating stability determines nanoparticle diagnostics efficacy and cell functionality. Small 2010, 6, 2136–2145. [Google Scholar] [CrossRef] [PubMed]
  183. Spector, L.B. Covalent enzyme-substrate intermediates in transferase reactions. Bioorg. Chem. 1973, 2, 311–321. [Google Scholar] [CrossRef]
  184. Chanana, M.; Rivera_Gil, P.; Correa-Duarte, M.A.; Liz-Marzán, L.M.; Parak, W.J. Physicochemical properties of protein-coated gold nanoparticles in biological fluids and cells before and after proteolytic digestion. Angew. Chem. Int. Ed. 2013, 52, 4179–4183. [Google Scholar] [CrossRef]
  185. Duff, S.J.; Murray, W.D. Bioconversion of forest products industry waste cellulosics to fuel ethanol: A review. Bioresour. Technol. 1996, 55, 1–33. [Google Scholar] [CrossRef]
  186. Olguin-Maciel, E.; Singh, A.; Chable-Villacis, R.; Tapia-Tussell, R.; Ruiz, H.A. Consolidated bioprocessing, an innovative strategy towards sustainability for biofuels production from crop residues: An overview. Agronomy 2020, 10, 1834. [Google Scholar] [CrossRef]
  187. Damayanti, D.; Supriyadi, D.; Amelia, D.; Saputri, D.R.; Devi, Y.L.L.; Auriyani, W.A.; Wu, H.S. Conversion of Lignocellulose for Bioethanol Production, Applied in Bio-Polyethylene Terephthalate. Polymers 2021, 13, 2886. [Google Scholar] [CrossRef]
  188. Dahnum, D.; Tasum, S.O.; Triwahyuni, E.; Nurdin, M.; Abimanyu, H. Comparison of SHF and SSF processes using enzyme and dry yeast for optimization of bioethanol production from empty fruit bunch. Energy Procedia 2015, 68, 107–116. [Google Scholar] [CrossRef] [Green Version]
  189. Leong, Y.K.; Chew, K.W.; Chen, W.-H.; Chang, J.-S.; Show, P.L. Reuniting the Biogeochemistry of Algae for a Low-Carbon Circular Bioeconomy. Trends Plant Sci. 2021. [Google Scholar] [CrossRef]
  190. Parisutham, V.; Kim, T.H.; Lee, S.K. Feasibilities of consolidated bioprocessing microbes: From pretreatment to biofuel production. Bioresour. Technol. 2014, 161, 431–440. [Google Scholar] [CrossRef]
  191. Hasunuma, T.; Kondo, A. Consolidated bioprocessing and simultaneous saccharification and fermentation of lignocellulose to ethanol with thermotolerant yeast strains. Process. Biochem. 2012, 47, 1287–1294. [Google Scholar] [CrossRef]
  192. Okamoto, K.; Uchii, A.; Kanawaku, R.; Yanase, H. Bioconversion of xylose, hexoses and biomass to ethanol by a new isolate of the white rot basidiomycete Trametes versicolor. Springerplus 2014, 3, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Lynd, L.R.; Van Zyl, W.H.; McBride, J.E.; Laser, M. Consolidated bioprocessing of cellulosic biomass: An update. Curr. Opin. Biotechnol. 2005, 16, 577–583. [Google Scholar] [CrossRef] [PubMed]
  194. Jeihanipour, A.; Taherzadeh, M.J. Ethanol production from cotton-based waste textiles. Bioresour. Technol. 2009, 100, 1007–1010. [Google Scholar] [CrossRef]
  195. Sanchis-Sebastiá, M.; Novy, V.; Stigsson, L.; Galbe, M.; Wallberg, O. Towards circular fashion–transforming pulp mills into hubs for textile recycling. RSC Advances 2021, 11, 12321–12329. [Google Scholar] [CrossRef]
  196. Sanchis-Sebastiá, M.; Ruuth, E.; Stigsson, L.; Galbe, M.; Wallberg, O. Novel sustainable alternatives for the fashion industry: A method of chemically recycling waste textiles via acid hydrolysis. J. Waste Manag. 2021, 121, 248–254. [Google Scholar] [CrossRef] [PubMed]
  197. Jeihanipour, A.; Karimi, K.; Niklasson, C.; Taherzadeh, M.J. A novel process for ethanol or biogas production from cellulose in blended-fibers waste textiles. J. Waste Manag. 2010, 30, 2504–2509. [Google Scholar] [CrossRef] [PubMed]
  198. Kuo, C.-H.; Lin, P.-J.; Wu, Y.-Q.; Ye, L.-Y.; Yang, D.-J.; Shieh, C.-J.; Lee, C.-K. Simultaneous saccharification and fermentation of waste textiles for ethanol production. BioResources 2014, 9, 2866–2875. [Google Scholar] [CrossRef] [Green Version]
  199. McIntosh, S.; Vancov, T.; Palmer, J.; Morris, S. Ethanol production from cotton gin trash using optimised dilute acid pretreatment and whole slurry fermentation processes. Bioresour. Technol. 2014, 173, 42–51. [Google Scholar] [CrossRef]
  200. Yildirir, E. Chemical Recycling of Waste Plastics via Hydrothermal Processing. Ph.D. Thesis, University of Leeds, Leeds, UK, 2015. [Google Scholar]
  201. Wikberg, H.; Grönberg, V.; Jermakka, J.; Kemppainen, K.; Kleen, M.; Laine, C.; Paasikallio, V.; Oasmaa, A. Hydrothermal refining of biomass: An overview and future perspectives. Tappi J. 2015, 14, 195–207. [Google Scholar] [CrossRef]
  202. Tang, K.; Zhang, Y.; Lin, D.; Han, Y.; Chen, C.-T.A.; Wang, D.; Lin, Y.-S.; Sun, J.; Zheng, Q.; Jiao, N. Cultivation-independent and cultivation-dependent analysis of microbes in the shallow-sea hydrothermal system off Kueishantao Island, Taiwan: Unmasking heterotrophic bacterial diversity and functional capacity. Front. Microbiol. 2018, 9, 279. [Google Scholar] [CrossRef]
  203. Hongthong, S.; Leese, H.S.; Allen, M.J.; Chuck, C.J. Assessing the Conversion of Various Nylon Polymers in the Hydrothermal Liquefaction of Macroalgae. Environments 2021, 8, 34. [Google Scholar] [CrossRef]
  204. Zheng, C.; Ma, X.; Yao, Z.; Chen, X. The properties and combustion behaviors of hydrochars derived from co-hydrothermal carbonization of sewage sludge and food waste. Bioresour. Technol. 2019, 285, 121347. [Google Scholar] [CrossRef] [PubMed]
  205. Hou, W.; Ling, C.; Shi, S.; Yan, Z.; Zhang, M.; Zhang, B.; Dai, J. Separation and characterization of waste cotton/polyester blend fabric with hydrothermal method. Fibers Polym. 2018, 19, 742–750. [Google Scholar] [CrossRef]
  206. Shi, S.; Zhang, M.; Ling, C.; Hou, W.; Yan, Z. Extraction and characterization of microcrystalline cellulose from waste cotton fabrics via hydrothermal method. J. Waste Manag. 2018, 82, 139–146. [Google Scholar] [CrossRef]
  207. Lin, Y.; Ma, X.; Peng, X.; Yu, Z. A mechanism study on hydrothermal carbonization of waste textile. Energy Fuels 2016, 30, 7746–7754. [Google Scholar] [CrossRef]
  208. Li, Y.; Shao, M.; Huang, M.; Sang, W.; Zheng, S.; Jiang, N.; Gao, Y. Enhanced remediation of heavy metals contaminated soils with EK-PRB using β-CD/hydrothermal biochar by waste cotton as reactive barrier. Chemosphere 2022, 286, 131470. [Google Scholar] [CrossRef]
  209. Xu, Z.; Zhang, T.; Yuan, Z.; Zhang, D.; Sun, Z.; Huang, Y.; Chen, W.; Tian, D.; Deng, H.; Zhou, Y. Fabrication of cotton textile waste-based magnetic activated carbon using FeCl 3 activation by the Box–Behnken design: Optimization and characteristics. RSC Adv. 2018, 8, 38081–38090. [Google Scholar] [CrossRef] [Green Version]
  210. Xu, Z.; Zhou, Y.; Sun, Z.; Zhang, D.; Huang, Y.; Gu, S.; Chen, W. Understanding reactions and pore-forming mechanisms between waste cotton woven and FeCl3 during the synthesis of magnetic activated carbon. Chemosphere 2020, 241, 125120. [Google Scholar] [CrossRef]
  211. Kawamura, K.; Sako, K.; Ogata, T.; Tanabe, K. Environmentally friendly, hydrothermal treatment of mixed fabric wastes containing polyester, cotton, and wool fibers: Application for HMF production. Bioresour. Technol. Rep. 2020, 11, 100478. [Google Scholar] [CrossRef]
  212. Asaoka, Y.; Funazukuri, T. Hydrothermal saccharification of cotton cellulose in dilute aqueous formic acid solution. Res. Chem. Intermed. 2011, 37, 233–242. [Google Scholar] [CrossRef]
  213. Keh, E.Y.M.; Yao, L.; Liao, X.; Liu, Y.; Cheuk, K.; Chan, A. Method for Separating and Recycling a Waste Polyester-Cotton Textile by Means of a Hydrothermal Reaction Catalyzed by an Organic Acid. U.S. Patent 2020/0262108 A1, 20 August 2020. [Google Scholar]
  214. McKinney, R. Ammonolysis of Nylon. US 5302756A, 12 April 1994. [Google Scholar]
  215. Molino, A.; Larocca, V.; Chianese, S.; Musmarra, D. Biofuels production by biomass gasification: A review. Energies 2018, 11, 811. [Google Scholar] [CrossRef] [Green Version]
  216. Labaki, M.; Jeguirim, M. Thermochemical conversion of waste tyres—A review. Environ. Sci. Pollut. Res. 2017, 24, 9962–9992. [Google Scholar] [CrossRef]
  217. Samolada, M.; Zabaniotou, A. Potential application of pyrolysis for the effective valorisation of the end of life tires in Greece. Environ. Dev. 2012, 4, 73–87. [Google Scholar] [CrossRef]
  218. Portofino, S.; Donatelli, A.; Iovane, P.; Innella, C.; Civita, R.; Martino, M.; Matera, D.A.; Russo, A.; Cornacchia, G.; Galvagno, S. Steam gasification of waste tyre: Influence of process temperature on yield and product composition. J. Waste Manag. 2013, 33, 672–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Yasin, S. Eco-Design for End-of-Life Phase of Flame Retardant Textiles. Ph.D. Thesis, Université Lille 1-Sciences et Technologies, Lille, France, 2017. [Google Scholar]
  220. Cañete Vela, I.; Maric, J.; Seemann, M. Valorisation of textile waste via steam gasification in a fluidized bed reactor. In Proceedings of the 7th International Conference on Sustainable Solid Waste Management, Crete Island, Greece, 26–29 June 2019. [Google Scholar]
  221. Wen, C.; Wu, Y.; Chen, X.; Jiang, G.; Liu, D. The pyrolysis and gasification performances of waste textile under carbon dioxide atmosphere. J. Therm. Anal. Calorim. 2017, 128, 581–591. [Google Scholar] [CrossRef]
  222. Wu, Y.; Wen, C.; Chen, X.; Jiang, G.; Liu, G.; Liu, D. Catalytic pyrolysis and gasification of waste textile under carbon dioxide atmosphere with composite Zn-Fe catalyst. Fuel Process. Technol. 2017, 166, 115–123. [Google Scholar] [CrossRef]
  223. Ziyaei, M.D.; Barikani, M.; Honarkar, H. Recycling of Polyethylene Terephthalate (PET) via Glycolysis Method for Synthesis Waterborne Polyurethane. In Proceedings of the International Seminar on Polymer Science and Technology, Tehran, Iran, 10–22 November 2018; pp. 520–523. [Google Scholar]
  224. Guo, Z.; Adolfsson, E.; Tam, P.L. Nanostructured micro particles as a low-cost and sustainable catalyst in the recycling of PET fiber waste by the glycolysis method. J. Waste Manag. 2021, 126, 559–566. [Google Scholar] [CrossRef] [PubMed]
  225. Al-Sabagh, A.; Yehia, F.; Eshaq, G.; Rabie, A.; ElMetwally, A. Greener routes for recycling of polyethylene terephthalate. Egypt. J. Pet. 2016, 25, 53–64. [Google Scholar] [CrossRef] [Green Version]
  226. Heiran, R.; Ghaderian, A.; Reghunadhan, A.; Sedaghati, F.; Thomas, S.; hossein Haghighi, A. Glycolysis: An efficient route for recycling of end of life polyurethane foams. J. Polym. Res. 2021, 28, 1–19. [Google Scholar] [CrossRef]
  227. Al-Sabagh, A.M.; Yehia, F.Z.; Eissa, A.-M.M.; Moustafa, M.E.; Eshaq, G.; Rabie, A.-R.M.; ElMetwally, A.E. Glycolysis of poly (ethylene terephthalate) catalyzed by the Lewis base ionic liquid [Bmim][OAc]. Ind. Eng. Chem. Res. 2014, 53, 18443–18451. [Google Scholar] [CrossRef]
  228. Guo, Z.; Eriksson, M.; de la Motte, H.; Adolfsson, E. Circular recycling of polyester textile waste using a sustainable catalyst. J. Clean. Prod. 2021, 283, 124579. [Google Scholar] [CrossRef]
  229. Hommez, B.; Goethals, E.J. Degradation of nylon-6 by glycolysis. Part 1: Identification of degradation products. J. Macromol. Sci.—Pure Appl. Chem. 1998, 35, 1489–1505. [Google Scholar] [CrossRef]
  230. Huczkowski, P.; Kapko, J.; Olesiak, R. Degradation of nylon-6 in ethylene glycol. Polymer 1978, 19, 77–80. [Google Scholar] [CrossRef]
  231. Huczkowski, P.; Kapko, J. Degradation of nylon-6 in ethylene glycol: 2. Mathematical illustration of degradation. Polymer 1980, 21, 86–88. [Google Scholar] [CrossRef]
  232. Holland, B.J.; Hay, J.N. Thermal degradation of nylon polymers. Polym. Int. 2000, 49, 943–948. [Google Scholar] [CrossRef]
  233. Kim, K.J.; Dhevi, D.M.; Lee, J.S.; Dal Cho, Y.; Choe, E.K. Mechanism of glycolysis of nylon 6, 6 and its model compound by ethylene glycol. Polym. Degrad. Stab. 2006, 91, 1545–1555. [Google Scholar] [CrossRef]
  234. Hu, Y.; Wang, Y.; Zhang, X.; Qian, J.; Xing, X.; Wang, X. Synthesis of poly (ethylene terephthalate) based on glycolysis of waste PET fiber. J. Macromol. Sci. A 2020, 57, 430–438. [Google Scholar] [CrossRef]
  235. Shukla, S.; Harad, A.M. Glycolysis of polyethylene terephthalate waste fibers. J. Appl. Polym. Sci. 2005, 97, 513–517. [Google Scholar] [CrossRef]
  236. Maryan, A.S.; Montazer, M.; Damerchely, R. Discoloration of denim garment with color free effluent using montmorillonite based nano clay and enzymes: Nano bio-treatment on denim garment. J. Clean. Prod. 2015, 91, 208–215. [Google Scholar] [CrossRef]
  237. Groves, E.; Palenik, C.S.; Palenik, S. A survey of extraction solvents in the forensic analysis of textile dyes. Forensic Sci. Int. 2016, 268, 139–144. [Google Scholar] [CrossRef] [PubMed]
  238. Mu, B.; Yang, Y. Complete separation of colorants from polymeric materials for cost-effective recycling of waste textiles. Chem. Eng. Sci. 2022, 427, 131570. [Google Scholar] [CrossRef]
  239. Haslinger, S.; Wang, Y.; Rissanen, M.; Lossa, M.B.; Tanttu, M.; Ilen, E.; Määttänen, M.; Harlin, A.; Hummel, M.; Sixta, H. Recycling of vat and reactive dyed textile waste to new colored man-made cellulose fibers. Green Chem. 2019, 21, 5598–5610. [Google Scholar] [CrossRef] [Green Version]
  240. Padmanaban, A.; Murugadoss, G.; Venkatesh, N.; Hazra, S.; Rajesh Kumar, M.; Tamilselvi, R.; Sakthivel, P. Electrochemical determination of harmful catechol and rapid decolorization of textile dyes using ceria and tin doped ZnO nanoparticles. J. Environ. Chem. Eng. 2021, 9, 105976. [Google Scholar] [CrossRef]
  241. Li, Y.; Li, M.; Lu, J.; Li, X.; Ge, M. Decoloration of waste PET alcoholysis liquid by an electrochemical method. Water Sci. Technol. 2018, 77, 2463–2473. [Google Scholar] [CrossRef]
  242. Huang, J.; Yan, D.; Dong, H.; Li, F.; Lu, X.; Xin, J. Removal of trace amount impurities in glycolytic monomer of polyethylene terephthalate by recrystallization. J. Environ. Chem. Eng. 2021, 106277. [Google Scholar] [CrossRef]
Figure 1. The critical point of CE.
Figure 1. The critical point of CE.
Polymers 13 03834 g001
Figure 2. The sorting and identification of textiles using IoT.
Figure 2. The sorting and identification of textiles using IoT.
Polymers 13 03834 g002
Figure 3. Mechanisms of possible textile waste recycling routes.
Figure 3. Mechanisms of possible textile waste recycling routes.
Polymers 13 03834 g003
Figure 4. The recycling of garments by open and closed-loop systems.
Figure 4. The recycling of garments by open and closed-loop systems.
Polymers 13 03834 g004
Figure 5. The possible pathways to recycling textiles using chemical methods.
Figure 5. The possible pathways to recycling textiles using chemical methods.
Polymers 13 03834 g005
Table 1. The recognition rate with different types of analyzers.
Table 1. The recognition rate with different types of analyzers.
TextileAnalyzer Mathematic Algorithms Recognition Rate, %Ref.
Pure (polyester slash; wool; cotton; polyester normal; nylon).
Mix (polyester/nylon; polyester/wool; polyester/cotton slash; polyester/cotton).
NIRSVM, MLP + CNN92–98 [74]
Cotton, polyester, polyamide, acrylic, silk, wool. NIRSIMCA93–97 [79]
Polypropylene, polyethylene terephthalate, polylactic acid, cashmere, Tencel, cotton, wool NIRPCA + SIMCA100[80]
WoolNIRPLS, ECR [81]
Satin, twill, plainResNet-50CNN99.3[82]
Satin, twill, plain Epson Scan V330TILT, HOG, FCM clustering, histogram equalization data 94.5[83]
Non-woven fabric, plain weave, twill weave, double jersey, satin weaveCanon scanner 9950F, Self-organizing map (SOM) networkCIE + Co-occurrence matrix92.6[84]
Cotton, polyester, viscose NIRNA76–100[85]
CIE: C-Illuminant (light source C), FCM: fuzzy C-means, HOG: histogram of oriented gradients, PCA: principal component analysis, SIMCA: soft independent modeling of class analogy, SOM: self-organizing map, SVM: support vector machine, TILT: transform invariant low-rank textures. PLS: partial least squares. ECR: elastic component regression. NA: not available.
Table 2. Various chemical processes applied in textile recycling at the industrial scale.
Table 2. Various chemical processes applied in textile recycling at the industrial scale.
CompanyMaterialMethodCountryRef.
AmbercycleTextileEnzymatic treatmentLos Angeles[110]
BlockTexxCotton and polyester (PET)Fiber separation
technology
Australia[111]
FENCPET and textileHydrolysisTaiwan[112]
Infinited FiberCellulose and polyester (PET)Alkaline extractionFinland[113,114]
IoncellCotton and polyester (PET)Ionic liquid solvent
polymer dissolution
Finland[115]
LenzingPre-consumer cotton and post-consumer garmentsClosed loops and chemical recoveryAustria[116]
Tyton BioSciencesCotton pulp, polyester (PET), poly-cotton blendsSubcritical water treatmentDanville[117]
Worn Again TechnologiesCotton and polyester (PET)Dissolution polymer
Solvent separation
UK[118]
Table 3. The drawbacks, advantages, and disadvantages of chemically recycling textiles.
Table 3. The drawbacks, advantages, and disadvantages of chemically recycling textiles.
Type of Technology Advantages Disadvantages Ref.
Pyrolysis
Simple process
Many different kinds of raw materials can be used for the pyrolysis process.
High-temperature reaction process
High energy consumption
[132,133]
Enzymatic
Hydrolysis
The novel microorganism enhances the production of biopolymers and conversion rate. It can effectively be spun into novel waste textile fibers
Under mild conditions
▪ Low energy demand
Utilizes environmentally friendly solvents and chemicals
Only possible to recycle certain materials using enzymatic hydrolysis such as rayon, cotton, hemp, etc.
Needs large amounts of water
[120]
Hydrothermal
Low ash
High heating values
Relatively low temperatures than gasification and pyrolysis processes
Reduced oxygen content
Long reaction time
Lower purity and heterogeneity
[134,135,136]
Gasification
Potentially mixed textile waste
High energy consumption
High temperature is needed
[137]
Glycolysis
Low energy consumption
Low selectivity
Without a catalyst, slow processing
[138]
Ammonolysis
Various types of textile waste can be applied
Generates a mixture of primary, secondary amines
Applied toxic solvent (ammonia)
High pressure and temperature
[139]
Table 4. Yields of solid, liquid, and gas products produced via textile waste pyrolysis.
Table 4. Yields of solid, liquid, and gas products produced via textile waste pyrolysis.
Raw Materials ReactorsCatalyst Reaction Time, Min Temperature, °CYield (%)Major Products Ref.
Solid Liquid Gas
Egyptian banknote ELCBs
Cotton, 100%
BatchNA3070015.6529.2855.072-Propanone, toluene, 3-furaldehyde, 2-furalmethanol, benzaldehyde, phenol, acetophenone[147]
Hemp, 100%BatchNA1207502448.527.50Activated carbon[155]
Flax, 100%Fixed bedZnCl222045044.80NANAActivated carbon[156]
Egyptian banknote ELCBs
Cotton, 100%
BatchNA3060017.2630.2852.461,3-Dioxolane-2-propanal, 2-methyl, furfural, 2-furanmethanol, 4,4-ethylenedioxy-1-pentylamine[147]
Flax, 100%Fixed bedZnCl222045044.80NANAActivated carbon[156]
Cotton, 70%
Polyester, 30%
BatchDye-Originating Heavy Metals35500–70017.7937.5944.62Activated carbon [147]
Egyptian banknote ELCBs
Cotton, 100%
BatchNA3050020.7439.7539.51Toluene, furfural, 2-furanmethanol, 2-furancarboxaldehyde, 5-methyl[147]
Hemp, 100%Fixed bedZnCl222045041.60NANAActivated carbon[156]
Cotton, 100%Fixed bed Na2CO312060016.2529.4954.26Furans, ketones[130]
Coir, 100%Fixed bedNA24080037.4047.4018.20Activated carbon[157]
Abaca, 100%Fixed bedNA24080028.6048.1023.60Activated carbon[157]
Cotton, 100%BatchNA7070012.5074.0013.50Double bond carboxyl and carbonyl liquid[143]
Low Grade
biomass fiber (Flax, 100%)
BatchNA120900205525Activated carbon[155]
Flax, 100%Fixed bedH3PO422045039.20NANAActivated carbon[156]
Biomass fiber waste Jute, 100%Fixed bedNA24080024.6059.6015.90Activated carbon[157]
Table 5. Conditions for converting textile fiber to glucose via enzymatic hydrolysis.
Table 5. Conditions for converting textile fiber to glucose via enzymatic hydrolysis.
Textile
Materials
Type of
Pretreatment
Condition of Pretreatment EnzymeCondition of HydrolysisGlucose
Yield, %
Ref.
T (°C)T (h)T (°C)T (h)
Cotton, 90%
Wool, 5%
Polyester, 5%
Protease
(Enzymatic)
5048Cellic CTec3® and Savinase 12T®50 7095.0[175]
Used JeansPhosphoric acid
(Acid)
507Cellulase Trichoderma reesei and Aspergillus niger509679.2[176]
Cotton red T-shirt, 100%
cotton black T-shirt, 100%
N-methyl-morpholine-N-oxide& Phosphoric Acid
(Acid)
501Cellulase AP3, Aspergillus niger507287.0–95.0[177]
Cotton red T-shirt, 100%
cotton black T-shirt, 100%
NaOH/Urea
(Base)
501Cellulase AP3, Aspergillus niger507248.0–55.0[177]
Cotton T-shirts, 100%1-allyl-3-methylimidazolium chloride (Ionic liquid)1101.5G. xylinus (Acetobacter aceti subsp. xylinus or A. xylinus)502481.6[168]
Towels, cellulose content 87.8%Untreated PretreatmentNANACellulase2000.0374.2[169]
Waste blue jeans (polyester/cotton), 40%/60%Sodium
carbonate
(Base)
1502Celluclast 1.5 L and β-glucosidase457281.71[170]
Cotton, 100%Sodium
carbonate
(Base)
1502Celluclast 1.5 L and β-glucosidase457288.0[170]
Textile from End-of-life euro banknotes NaOH/Urea
(Base)
−206Cellulase5038296.0[178]
Table 6. Ethanol production after enzymatic hydrolysis and microorganism fermentation of textile wastes.
Table 6. Ethanol production after enzymatic hydrolysis and microorganism fermentation of textile wastes.
Textile lType of PretreatmentHydrolysis Enzyme for Hydrolysis Fermentation ProcessMicroorganism Type of Fermentation EthanolYield, %Ref.
T (°C)T (h)T (°C)T (h)
Polyester/cotton
(50%/50%)
N-methyl morpholine-N-oxide (Cellulose solvent)NA48Cellulase and β-glucosidase3024Saccharomyces cerevisiaeSHF91.91[197]
Polyester/cotton (40%/60%)NaOH/Urea
(Alkali)
457230 FPU C ellulase and 60 IU β-glucosidase3672Saccharomyces cerevisiaeSSF70.00[163]
Cotton, 100%Na2S2O4 and Na2CO3
(Alkali)
5048Cellulase AP33748Zymomonas mobilisSSF90.00[198]
Viscose/polyester
(60%/40%)
N-methylmorpholine-N-oxide (Cellulose solvent)NA48Cellulase and β-glucosidase3024Saccharomyces cerevisiaeSHF94.99[197]
Cotton, 100%Na2CO3
(Alkali)
4572Cellulase and β-glucosidase3224Saccharomyces cerevisiae CCUG 53310NA69.40[170]
Cotton ginning trash, 100%Sulfuric acid (Acid)5096Cellic CTec 2 cellulase30NASaccharomyces cerevisiaeSSF70.00[199]
SHF: separated hydrolysis dermentation, SSF: simultaneous saccharification and fermentation, NA: not available.
Table 7. Products and conditions of textile recycling using hydrothermal methods.
Table 7. Products and conditions of textile recycling using hydrothermal methods.
FeedstockReactor Catalyst Time, h Temperature, °CMajor Product Yield, %Ref.
A blue cotton/polyester, BatchHydrochloric acid3150Cellulose powder49.3[205]
Waste cotton fabricsBatchHydrochloric acid1.6150Microcrystalline cellulose85.5[206]
Cotton/synthetic fibers, Autoclave NA90280Volatile compound (CH4, C2H4)98.0[207]
Waste cottonsHydrothermal reactor β-cyclodextrin3250Activated carbonNA[208]
Cotton textile wasteTubularFerric chloride 1700Nanopowder 32.6[209]
Waste cotton wovenTubularFerric chloride 1.5700Activated carbonNA[210]
Light blue and white uniform with polyester/cotton, Hydrothermal reactorCitric acid 12255′-hydroxymethylfurfural12.5[211]
Cotton Semi-batchFormic acid1250Glucose83.8[212]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Damayanti, D.; Wulandari, L.A.; Bagaskoro, A.; Rianjanu, A.; Wu, H.-S. Possibility Routes for Textile Recycling Technology. Polymers 2021, 13, 3834. https://doi.org/10.3390/polym13213834

AMA Style

Damayanti D, Wulandari LA, Bagaskoro A, Rianjanu A, Wu H-S. Possibility Routes for Textile Recycling Technology. Polymers. 2021; 13(21):3834. https://doi.org/10.3390/polym13213834

Chicago/Turabian Style

Damayanti, Damayanti, Latasya Adelia Wulandari, Adhanto Bagaskoro, Aditya Rianjanu, and Ho-Shing Wu. 2021. "Possibility Routes for Textile Recycling Technology" Polymers 13, no. 21: 3834. https://doi.org/10.3390/polym13213834

APA Style

Damayanti, D., Wulandari, L. A., Bagaskoro, A., Rianjanu, A., & Wu, H. -S. (2021). Possibility Routes for Textile Recycling Technology. Polymers, 13(21), 3834. https://doi.org/10.3390/polym13213834

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop