Next Article in Journal
Enhanced Antibacterial Activity of Vancomycin Loaded on Functionalized Polyketones
Previous Article in Journal
Mattel’s ©Barbie: Preventing Plasticizers Leakage in PVC Artworks and Design Objects through Film-Forming Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Environment-Friendly Polyurethanes from Polyols Recovered from the Recycling and Renewable Resources: A Review

1
School of Mechanical and Precision Instrument Engineering, Xi’an University of Technology, Xi’an 710048, China
2
School of Printing, Packaging Engineering and Digital Media Technology, Xi’an University of Technology, Xi’an 710048, China
3
Key Laboratory of Auxiliary Chemistry and Technology for Chemical Industry, Ministry of Education, Shaanxi University of Science and Technology, Xi’an 710021, China
4
Shaanxi Collaborative Innovation Center of Industrial Auxiliary Chemistry and Technology, Shaanxi University of Science and Technology, Xi’an 710021, China
*
Authors to whom correspondence should be addressed.
Polymers 2024, 16(13), 1889; https://doi.org/10.3390/polym16131889
Submission received: 3 June 2024 / Revised: 25 June 2024 / Accepted: 27 June 2024 / Published: 2 July 2024

Abstract

:
Polyurethane (PU) is among the most universal polymers and has been extensively applied in many fields, such as construction, machinery, furniture, clothing, textile, packaging and biomedicine. Traditionally, as the main starting materials for PU, polyols deeply depend on petroleum stock. From the perspective of recycling and environmental friendliness, advanced PU synthesis, using diversified resources as feedstocks, aims to develop versatile products with excellent properties to achieve the transformation from a fossil fuel-driven energy economy to renewable and sustainable ones. This review focuses on the recent development in the synthesis and modification of PU by extracting value-added monomers for polyols from waste polymers and natural bio-based polymers, such as the recycled waste polymers: polyethylene terephthalate (PET), PU and polycarbonate (PC); the biomaterials: vegetable oil, lignin, cashew nut shell liquid and plant straw; and biomacromolecules: polysaccharides and protein. To design these advanced polyurethane formulations, it is essential to understand the structure–property relationships of PU from recycling polyols. In a word, this bottom-up path provides a material recycling approach to PU design for printing and packaging, as well as biomedical, building and wearable electronics applications.

Graphical Abstract

1. Introduction

PU, a kind of versatile polymer, is commonly produced by the synthesis of polyols and polyisocyanate through step-growth polymerization (Figure 1). PU, in the form of soft and hard foam plastics, elastomers, coatings, adhesives, films, hydrogels, waterproofing materials, paving materials and other forms, is widely used in many industries, such as leather, textile, printing and packaging, food processing, machinery, furniture, construction and even biomedical fields [1,2,3,4,5,6]. The soft segment of PU, composed of oligomer polyols, is mainly derived from petrochemical resources. However, with the crisis of energy shortage and environmental protection, the recycling of more sustainable resources as feedstocks is an excellent choice. Recently, there are two general methods for preparing sustainable polyols: recycling waste polymers to valuable monomers for waste-based polyols [7,8,9]; and exploring biomass resources to make monomers for bio-based polyols [10,11,12].
Plastic products are ubiquitous in our daily life and agricultural production. When these products reach the end of their life or use, they will be discarded and become waste plastic. There is a growing interest in the chemical recycling of waste polymers to recover valuable materials [13]. At present, some waste polymers could be degraded into value-added polyols as the raw material for PU, such as PET [14,15], PU [16] and PC wastes [17]. Meanwhile, the development and utilization of renewable resources to extract bio-based polyols to create sustainable “green” PU materials has also become a new bright spot [18]. Bio-based PU has undergone vigorous development in recent years, due to it saving resources, causing less environmental pollution, and having easy accessibility, a low cost and biodegradability [19]. Several bio-renewable materials have been introduced as feedstocks for PU, such as vegetable oil, nut shell, plant straw, lignin, starch, terpene and rosin, etc. [3,20,21]. These waste-based and bio-based polyols as feedstock are then reacted with various diisocyanate to facilitate PU production [22].
In recent years, several excellent reviews have discussed the recycling of waste PU materials and the advanced applications of waste PU derived materials in different fields [23,24,25]. However, recent advances in the preparation of PU materials from the extraction of valuable components from various recyclable resources have not been reviewed in detail. Additionally, some recent reviews have focused on the synthesis of bio-based PU, especially vegetable oil [26,27,28], and few of them have highlighted the waste-based polyol [29]. However, no review specifically covers the waste-based and bio-based PU as a whole. The challenge is how to produce PU with properties similar to traditional polyurethanes through effective resource conversion. Herein, we endeavor to focus on ever-expanding research of recyclable and environmentally friendly polyurethanes. The mechanism, types, synthesis, properties and applications of waste-based polyurethane and bio-based polyurethane were analyzed and compared with raw materials. Therefore, in this review we intended to aim at the most significant waste-based and bio-based PU and their products for sustainable printing and packaging, biomedical, buildings and wearable electronics applications. Finally, the cost and output of the waste-based and bio-based raw materials of PU are compared from a circular economy perspective. The true expansion of this circular economy market requires concerted action across the value chain.

2. Waste-Based Feedstocks for Polyurethanes

PU can be a kind of block polymer made from long and short chain raw materials. In general, long-chain polyols form soft segments, while hard segments are composed of short chain polyisocyanates [30]. Polyether, polyester, polycarbonate and other oligomers polyols can be used as soft segments [31]. The hard segments have a high glass transition temperature and melting temperature, providing the high hardness and strength of PU. Nevertheless, the soft segments have a lower glass transition temperature, giving PU flexibility, cross-linking ability and elasticity [32]. At the early stage, petroleum-based monomers were mainly used as precursors for polyols. Recently, there lies a huge potential in the development of new methods and technologies to produce polyols from waste polymers. Polymers are made up of a series of monomers reacted together in a polymerization process. The waste-based polymers can be classified based on their monomer sources [33,34], which have been key raw materials for fresh polymers. They can be chemical degraded to oligomer with carboxyl and hydroxyl end groups as building blocks to produce new polyester [35].

2.1. Waste Polyethylene Terephthalate for PU

Polyethylene terephthalate (PET), the main type of thermoplastic polyester, is one of the most multifunctional plastics prepared from terephthalic acid and ethylene glycol via polycondensation reaction nowadays [36,37]. PET is widely used to produce hollow containers in the packaging industry, such as soft drinking bottles in daily life. The number of global consumptions of plastic bottles currently is 1 million per minute, reaching 500 billion a year [38]. Most of these PET plastics are waste in landfills, thrown into incinerators or into rivers and oceans. So the recycling of plastic waste has become an urgent requirement of all sectors of society, and it is the urgent responsibility of the packaging industry [39]. There is an advanced technology that recycles plastic waste into raw materials. By chemical recycling, waste PET plastic bottles are extracted to polyester polyol, a key raw material for PU production, to enhance the commercial added value [40,41].
The main principle of the chemical recovery of PET is a depolymerization reaction, resulting in a polymer chain being broken up into oligomer polyols. Glycolysis [42,43,44,45], hydrolysis [46], alcoholysis [47,48] and biodegradation [49,50] are developing fast, and mainly applied on the scale of waste PET recycling to obtain high-value monomers and oligomers for the synthesis of PU (Table 1). In the following sections, we describe them separately.

2.1.1. Glycolysis

Glycolysis, which is a reaction that converts an inhomogeneous system into a single phase, has been one of the most extensively prospective chemical recycling routes for most waste PET [51,52,53,54]. At the same time, it is the most commercially used route, and has been widely adopted by a number of well-known companies such as DuPont, Dow Chemical, Goodyear, Shell Polyester, Zimmer, Eastman, Kodak, and so on [55]. The glycolysis of PET is the transformation of PET to bis (hydroxyethyl) terephthalate (BHET) with the excess ethylene glycol as a transesterification agent, and the catalyst is mainly metal acetate (Figure 2A) [51,56]. This process essentially involves an esterification reaction through the breaking of the ester bond in the PET chain to produce BHET, dimer and oligomers [57].
In past and present, various oligomer polyols synthesized from waste PET plastic bottles via glycolysis are demonstrated along with their potential applications in Table 1. Atta et al. recycled PET waste to suitable polyol oligomers using trimethylolpropane or pentaerythritol for preparing PU foams (Figure 2B) [43]. The degradation of PET with ethylene glycol (EG) lead to the formation of hydroxyl groups in the resulting oligomers. Luo et al. reported that waste PET plastic bottles were degraded by diethylene glycol (DEG), producing monomeric or oligomeric diols with low molecular weight; then, PET degradation products, as raw materials, reacted with crude glycerol for producing polyols and PU foams (Figure 2C) [42]. Scremin et al. introduced the glycolysis reaction of PET using diethylene glycol, and the final glycolysis product PET polyol was used in PU adhesives by the one-shot method [52]. As can be seen, oligomers produced by glycolysis are mostly terminated by hydroxyl groups. In addition, Ramin Shamsi et al. reported that an efficient and reasonable green synthesis route was adopted, to add carbon nanotubes into polyurethane prepared from PET waste to obtain nanocomposites with excellent mechanical properties [58]. Meanwhile, the swelling properties of crosslinked polyurethanes in various non-polar and polar solvents were studied. The diffusion distribution and adsorption kinetics in different solvents were evaluated [59]. Therefore, the degradation product, polyols, can be applied into a valuable raw material for the later preparation of other polymer compounds.
Figure 2. Glycolysis of PET. (A) The recovery of polyol from PU and BHET from PET plastic waste by transesterification using ethyleneglycol as a transesterification agent, metal salts, Lewis acids, phosphine/phosphite ligands with metal salts and low valent [Ni(COD)2] as catalytic precursors; reproduced with permission [51]. Copyright 2019, Elsevier. (B) PET depolymerization with Pentaerythritol (Per). Reproduced with permission [43]. Copyright 2013, Wiley. (C) Polyols were synthesized by glycolysis with polyester waste as raw material, diethylene glycol and crude glycerol as oligomer mixture (d), diols (c), triols (a,b). Reproduced with permission [42]. Copyright 2014, Springer.
Figure 2. Glycolysis of PET. (A) The recovery of polyol from PU and BHET from PET plastic waste by transesterification using ethyleneglycol as a transesterification agent, metal salts, Lewis acids, phosphine/phosphite ligands with metal salts and low valent [Ni(COD)2] as catalytic precursors; reproduced with permission [51]. Copyright 2019, Elsevier. (B) PET depolymerization with Pentaerythritol (Per). Reproduced with permission [43]. Copyright 2013, Wiley. (C) Polyols were synthesized by glycolysis with polyester waste as raw material, diethylene glycol and crude glycerol as oligomer mixture (d), diols (c), triols (a,b). Reproduced with permission [42]. Copyright 2014, Springer.
Polymers 16 01889 g002

2.1.2. Hydrolysis

The hydrolysis of PET is performed so that the ester bond is attacked by water molecules under a certain temperature and pressure, producing terephthalic acid (TPA) and EG [60]. It is the reverse of the polycondensation reaction used to make the polymers in the first place, in which the addition of water or water vapor to the process causes decomposition [61]. Here are some new hydrolysis methods. Vjekoslav et al. investigated a convenient and quantitative hydrolysis method of waste PET into TPA by mechanochemical milling at ambient temperature and pressure (Figure 3A) [62]. Kang et al. developed a method using the microwave-assisted hydrolysis of PET into TPA by ZSM-5 zeolite catalyzed in neutral media [63]. TPA is extensively used in chemical building blocks for the synthesis of other chemicals, such as biodegradable plastics and the reproduction of PET by polymerization with EG, a characteristic of closed-loop recycling. Not only that, a new way to convert PET hydrolysis products to higher-value monomers, in which waste PET was decomposed into TPA and EG monomers and then the monomers were polymerized to aromatics and aromatic-derived compounds by serious microbial catalysts, is to be applied for manufacturing drug, cosmetics, disinfectant, animal feeds, bio-based monomers, and so on (Figure 3B) [46].

2.1.3. Alcoholysis

Another effective method for the trans-esterification reaction of the waste PET with methanol is alcoholysis, to prepare the dimethyl terephthalate (DMT) and EG. DMT is a monomer of polyester, mainly applied for the synthesis of polyester. Liu et al. recycled waste PET to DMT and EG under a supercritical methanol, and researched the effect of alcoholysis temperature and time on yield (Figure 4A) [64]. Actually, DMT from alcoholysis products is recycled in the generation of polyester. Jadhav et al. researched that transesterification reaction of DMT and EG to prepare BHET, which is a precursor for PET [65]. Therefore, a closed-loop process for waste PET treatment and synthesis was formed. What’s more, it is a good solution for the maximum utilization of resources, and to release the pressure of waste pollution. Imek et al. found that waste PET depolymerized to dioctyl terephthalate by alcoholysis with isooctyl alcohol, which was applied in construction materials [47].

2.1.4. Biodegradation

However, some new technologies appearing in recycling PET waste are biodegraded with bacteria. In 2016, Shosuke Yoshida et al. isolated a novel bacterium, which hydrolyzes PET to produce MHET, TPA and EG [49]. Sadeghi et al. introduced Bis (2-hydroxy ethylene) terephthalamide from PET to the ring-opening polymerization of Ɛ-caprolactone, and then synthesized a new biodegradable PU with HDI [66]. Wei et al. revealed that postconsumer PET food packaging containers were degraded by hydrolysis as promising biocatalysts, and the degradation production was more than 50% at 70 °C [50]. The nuclear magnetic resonance spectroscopy showed that the oxydiethylene unit, oxyethylene unit and 2-hydroxyethyl could form during the PET enzymatic degradation. The schematic illustration of the amorphous fraction of PET was degraded to ethylene glycol and terephthalic acid, indicating the exo-type chain scission (Figure 4B).
There are various PET degradation methods to meet the recycling treatment of different situations. By comparing the data presented in Table 1, it becomes evident that diverse degradation methods necessitate distinct conditions and yield dissimilar degradation products. The products are classified as follows: glycolysis converts PET to BHET, with ethylene glycol or other glycol, hydrolysis degrades PET to TPA under acidic or basic conditions, and the alcoholysis reaction depolymerizes PET to DMT using methanol. These resultant products can be employed for polyurethane preparation through various treatments to generate a wide range of polyurethane products. In particular, the glycolysis of PET involves the transesterification of hydroxyl compounds, resulting in reaction products with terminal hydroxyl groups that can be used to synthesize PU. These polyol products are then used as raw materials for the preparation of fresh polymer, such as PET, a characteristic of closed-loop recycling, as well as PU, used for producing various foams, coatings and adhesives [42,52,53].

2.2. Waste Polyurethane for PU

It is well known that PU is a block polymer synthesized from polyols and isocyanates, which are extremely valuable petrochemical materials. So, PU waste can also be recycled or recovered as a real value material in order to be put to further use. Because the strong polar carboxylate group and high density of hydrogen bonds between the molecules of PU is insoluble in non-polar groups, it possesses good oil resistance, chemical resistance, mechanical performance and wear resistance [67,68]. Owing to its excellent performance, it is widely used in many fields. As a consequence, a lot of PU waste exists in many forms: foams, films, elastomer and adhesive applied to cars [69], furniture [70], electrical appliances [71], buildings [72] and so on. From the direction of ecology, it is extremely easy for PU waste to contaminate the soil and destroy its components, as PU is difficult to degrade and its products, such as aromatic amines, are toxic [73].
On the other hand, PU feedstock recovery processes, to obtain high-value polyols as raw materials, are one of the most productive research topics today. The PU waste recovery methods include a series of physical and chemical processes, through which waste PU can either be cut into fragments which will be dissolved into fundamental hydrocarbon units, or depolymerized into constituent monomers as raw material to prepare new chemical products [16,74]. A series of research about the degradation of PU waste to obtain polyol performed by some reactions have been carried out. The recovery of the original raw materials, essentially the monomers of the polyols, can be applied in the synthesis of new PU products [75]. In order to understand the chemical recovery of PU waste in more detail, it is introduced in the following parts: hydrolysis, aminolysis, phosphorolysis and the hydroglycolysis of urethane and urea groups [76,77,78,79,80,81,82].
The main products of the hydrolysis reaction of polyether-based PU result are a diamine, a polyol and carbon dioxide [83]. The resulting polyol, with a similar composition to that of original polyether polyol, could replace the original polyols used for PU [84]. Furthermore, the resulting diamines are applied in isocyanates synthesis as the precursors, or given rigid characteristics by the alkoxylation of polyols or alkylene oxide. Additionally, due to the existence of ester groups, the polyester-based PU is hydrolyzed into monomers, including diacids and glycols or polyols [85]. Actually, the hydrolysis of polyester and polyether PU initially takes place in the hard segment through the decomposition of weak intermolecular hydrogen bonds between urethane groups on PU chains. The next step is the hydrolysis of the C=O group in the soft segment [86].
The glycolysis of the PU polymer is actually a transesterification reaction between carbamic ester and glycol, in which hydroxyl group of ethylene glycol replaces the ester group of urethane band [87,88,89,90,91]. A polyol mixture was composed of waste PU glycolysis products and aromatic polyols extracted from diphenylmethane diisocyanate, which could build up the physical and mechanical properties, heat resistance and fire resistance of rigid PU foams [92,93]. Shin et al. found that PU foam was made from recycled polyols with superior mechanical strength and thermal insulation performance compared with conventional foam [94]. Interestingly, glycolysis with polyfunctional alcohols can yield a mixture of biphasic products. It is composed of apolar polyether polyols in the upper phase, and aromatic compounds and polar glycolysis agents in the lower phase. Vanbergen et al. investigated the recovery of polyether polyols with high purity (97%) by fractional glycolysis in diglycerol, which could be used to replace the original polyether polyols in the later preparation of novel flexible PU foam materials [95]. Using split-phase glycolysis technology to recover PU production waste and product waste after consumption, closed-loop recycling can be realized. In addition, the secondary reaction is also unavoidable, resulting in various by-products, such as amines, isocyanates and unsaturation [96]. The main advantages of glycolysis are the low reaction temperature, short glycolysis time and the glycolysis product for producing new PU. This means it is also available for when these mixed polyol and isocyanate products could be reused and recycled as feedstock for the regeneration of PU with superior performance, and has been put into industrial production [87,97].
The aminolysis of PU is actually a chemical reaction, in which the ester group of urethane is exchanged with the amine group [16]. It takes place with dibutylamine, ethanolamine, lactam or lactam admixture and other amines react to break the chemical bond of PU waste under the high-temperature and high-pressure conditions [98]. The resulting decomposed polyol, the main products of ammonolysis, could be used as an alternative virgin polyol in fabricating PU. However, some researchers found that amine contaminants seriously affect the properties of polyols. Consequently, through the reaction of propylene oxide, the amine contaminants are separated from the recycled polyols, which can be applied to synthesize new PU with excellent performance [99,100]. In particular, aminolysis with monoethanolamine forms a hydroxyalkyl urea and an alcohol. Then, an intramolecular reaction occurs in hydroxyalkyl urea to form a cyclic urethane. Lastly, the reaction between the oxazolidone and monoethanolamine produce a bis(hydroxyethyl) urea as a chain extender in PU.
The phosphorolysis of polyurethanes is the ester exchange reaction processing with phosphoric and phosphonic acids, occurring between urethane and ester alkoxy group [101,102]. The process allowed the final products to mainly contain a mixture of oligomer of phosphorus and chlorine [103]. It can be used as a raw material of new PU, especially in the improvement of flame-retardant properties, adhesion properties and ultraviolet resistance [104,105]. Heinen et al. studied that the more phosphorus content in the oxygen index of PU foam, the greater the influence on the reduction in flammability [106].
Hydroglycolysis is a multiple method which is performed by combing water and glycols under a certain condition. PU reacts with diethylene glycol in an alkaline environment to form polyols and various intermediate chemicals [107]. The recovered polyols can also be resynthesized into PU products [108]. Unfortunately, the obtained polyol product is normally more complicated than conventional glycolysis. However, it is still a good way to deal with the complex mixture of dirty and contaminated PU waste that would otherwise have to go to landfills [109].
As can be seen from Table 2, most of the waste PU degradation products by different methods contain polyols. Therefore, the degradation products can be used as new raw materials to synthesize PU again. However, the products of waste PU after ammonolysis also include aromatic amine alkanolamine derivatives. These can be used to synthesize a variety of materials, including epoxy resins, polyesters, and polycarbonates. The PU waste must be treated and recycled effectively, not only to prevent pollution and protect environment, but also to reduce production cost and improve the resource utilization rate. More briefly, the pure monomer materials are obtained from waste PU through effective chemical treatment, and then these monomers are used as raw materials to prepare new products. Today, many companies exhibit their recovered polyols from waste in the chemicals market, compared with PU that has not been resynthesized.

2.3. Waste Polycarbonate for PU

Polycarbonate (PC) is an engineering plastic made from carbon dioxide (CO2) and epoxides, which has been widely used to prepare biocompatible biomaterials with environmental and economic benefits [110,111]. PC waste might be just like other waste, re-processed to recover raw materials by means of chemical recycling, such as methanolysis, hydrolysis, glycolysis and hydroglycolysis [112,113,114,115]. The carbonate group of PC could be completely converted to new phenol-containing carbamate derivatives, especially 2, 2-bis(4-hydroxyphenyl)propane (BPA) with mono- and di-hydroxyalkyl ethers, which are subsequently used as PC polyols in the synthesis of PU materials. Simultaneously, low molecular weight PC has been commercialized as a feedstock for the preparation of PU materials, such as elastomers, foams, adhesives, coatings [116,117,118,119,120,121].
Polycarbonate polyols show excellent chemical and thermal properties, especially better hydrolysis resistance, as well as the higher ageing resistance and biodegradability of PU products than polyester and polyether polyols in PU [17]. Beneš et al. successfully applied these to transform PC waste from scrap vehicles to valuable liquid polyols with hydroxyl number 250 mg KOH·g−1 by reacting them with transesterified oils (Figure 5) [122]. However, the PC recycling could be degraded into hydroxyl N,N′-diphenylene-isopylidenyl biscarbamates with low crystalline, resulting in low tensile properties. It was found that 2,2′-(ethyldioxy) diethylamine was added to the recovered PC monomer without any chemical cross-linking or blending reaction, and excellent thermoplastic properties of PU elastomers were obtained [117]. Additionally, recent studies demonstrated that polycarbonate diols in the soft segment of waterborne polyurethane (WPU) dispersions assured the excellent hydrolysis resistance and UV-curable ability. Moreover, it has been determined that polycarbonate PU is a hopeful strategy in biomedical field, such as interventional devices. Li et al. has found that novel phospholipid-based polycarbonate PU films could be coated on biomedical metal plates to build a bionic antifouling surface with excellent tensile properties [119].
With increasing attention paid to safety issues, there is a sustainable growth interest in developing green synthesis pathways for related polymers, such as non-isocyanate PUs (NIPUs). The synthesis of NIPUs using dicyclic carbonates prepared from carbon dioxide and epoxides has been widely studied [123,124]. Yang et al. proposed that an ideal route towards high molecular weight NIPUs was those synthesized by ammonolysis reaction with dicyclic carbonates from renewable resources [125].
To sum up, PET, PU and PC waste could be degraded into value-added polyols as the raw material for PU. This allows for the preparation of a rich variety of polyurethane product types, including foam, coating, and adhesive. The technology of preparing polyols from recycling resources makes this the leader in sustainable PU commercialization. The method can be a promising alternative to the traditional mechanical recovery of waste. Generally, since the raw materials applied in PU stem from fossil-based energy, there is a growing trend towards recyclable alternatives to petroleum. In this manner, the utilization of waste resources as chemical feedstocks for the chemical industry instead of oil is a sustainable alternative. This technology not only solves the solid-waste recycling issue, but serves as a source of raw stock in place of petroleum-based products, so there should be room for development in today’s world.

3. Bio-Based Feedstocks for Polyurethanes

Biomass resources, a kind of renewable energy, and its rational development and utilization, will effectively solve energy and ecological environment issues. With the rapid development of the industry, fossil resources tend to be exhausted gradually, so using low-cost biomass resources as raw materials meets the requirements of green development in both economic and environmental aspects. Polyols, an important component of PU, can be obtained from the biomass materials available, such as vegetable oil, cashew nut shell liquid, rosin, terpene and lignin [126]. Bio-based polyol was used as the soft segment and reacted with isocyanate to obtain the corresponding PU. These bio-based PUs have large potential in the polymer field and have excellent thermal-physical and mechanical properties, competing with traditional petroleum-based PUs. In particular, WPU merged as a new chemical candidate for ecofriendly adhesives and coatings has been greatly promising for bio-based applications. The content of this section is to review the process of bio-based PUs and modified PU by biomass materials.

3.1. PU from Renewable Monomers

3.1.1. Vegetable Oil

Vegetable oils (VOs) are one of the most abundant renewable resources, which hold great promise as raw materials to support the chemical conversion in a wide range. This can arise from the characteristic molecular structure, which mainly contains triglyceride, three long chain fatty acids, C=C bonds, as well as the actively terminal ester group, as illustrated in Figure 6A [127]. Thus, the desired structure can be obtained by chemical reactions between VOs as the soft segment (polyols) and isocyanate as the hard segment in PU, as depicted in Figure 6B. Some VOs have natural hydroxyl groups and can be directly used as polyols, such as castor oil. However, some VOs must be modified to introduce hydroxyl groups into the chains according to the C=C bonds and terminal ester group, such as soybean, sunflower, linseed oil, etc. [2]. Owing to the merits of green, inherent biodegradability and low cost, VOs have been employed to enhance the elasticity and mechanical properties of PU in industry [128] However, the demerits caused by the hydrophilic carboxylate groups in VOs-based PU should not be ignored, including low resistance for chemical agents and heat, and physicomechanical stability [129].
Castor oil (CO) can be directly employed as a polyol to react with isocyanates, which possesses active hydroxyl bonds (hydroxyl index ~163 mg KOH/g), illustrated in Figure 6A [130]. The three distributed active hydroxyl groups can facilitate the crosslinking of PU soft segment to obtain significant mechanical properties and thermal resistance [26]. It is reported that the unmodified castor oil contributes to the mechanical properties, and the water and chemical agent resistance of the WPU because of the existence of hydrophobic triglyceride structure [131,132]. In addition, the castor oil can also react with some diols, such as ethylene glycol (EG), glycerol, pentaerythritol, and sorbitol, to produce derivatives with molecular weights and different hydroxyl indexes.
All these derivatives are able to react with isocyanates to prepare WPU coatings [133]. Interestingly, it is expected that the derivatives from castor oil can be used as isocyanates to potentially prepare bio-based WPU [127]. As for the VOs without active hydroxyl groups, including soybean, sunflower, linseed, jatropha, palm oil and so on, it is necessary to introduce active hydroxyl groups into these VOs’ chains by chemical modification. Currently, five routes, including thiol-ene reaction, epoxidation, hydroformylation, ozonolysis and transesterification, have been developed to convert these VOs into PU polyols by decorating the C=C bonds or esters on their chains, as depicted in Figure 6A.
Figure 6. (A) The chemical composition of vegetable oil (VO)-based PU and the most common routes of VO-based polyols: castor oil (CO), palm oil-based polyol (POP), jatropha oil-based polyol (JOP), sunflower oil-based polyol (SUP), linseed oil-based polyol (LOP), soybean oil-based polyol (SOP). (B) The chemical structure (soft and hard segments) of polyurethane.
Figure 6. (A) The chemical composition of vegetable oil (VO)-based PU and the most common routes of VO-based polyols: castor oil (CO), palm oil-based polyol (POP), jatropha oil-based polyol (JOP), sunflower oil-based polyol (SUP), linseed oil-based polyol (LOP), soybean oil-based polyol (SOP). (B) The chemical structure (soft and hard segments) of polyurethane.
Polymers 16 01889 g006
For example, soybean oil is commonly modified via thiol-ene reaction and epoxidation. The C=C bonds can be epoxidized to graft to the active hydroxyl groups, which can endow soybean oil multi-functional properties as bio-polyols, bio-emulsifier and even non-isocyanates agent for the preparation of PU [83]. Dodangeh et al. synthesized a bio-polyol by epoxidized soybean oil to prepare PU. The lap shear strength and tensile strength can be as high as 6.478 MPa and 7.173 MPa, respectively [134]. Moreover, the epoxidized soybean oil can also play the role of bio-based emulsifier with glutaric acid, which can be alternative for preparing WPU instead of typical petroleum-based emulsifiers [135]. As to the sunflower oil, it can be epoxidized to obtain active hydroxyl groups to prepare bio-polyols [136,137]. Owing to the carboxylic acid group on the chain, it even can be used as a hydrophilic dihydroxy acid to endow PU with hydrophilicity. This can solve the issue of the lack of hydrophilic dihydroxy acid for WPU dispersions [91]. The bio-polyols from sunflower oil can also be prepared with high primary hydroxyl functionality by thiol-ene reaction [138]. Owing to the high content of hydroxyl polyols from sunflower oil, various kinds of chemical products, such as PU and polyester, can be synthesized and applied in a wide range with the assistance of ultraviolet curing [139,140,141]. Linseed oil can be modified via an esterification process, and is mainly employed as a chain extender for preparing WPU [142,143]. The obtained WPU presents good adhesion, durability and light resistance in wood coatings. Similar to the sunflower oil, the modified linseed oil can also be used as hydrophilic chain extender instead of dimethylol propionic acid (DMPA) for the preparation of anionic WPU [144]. Jatropha oil is modified to prepare polyol by epoxidation and oxirane ring opening reactions, with the advantages of low cost and easy extraction [145,146,147]. It can contribute to the colloidal stability and rheological properties for WPU, with various hydroxyl numbers [148,149]. Palm oil, consisting of triglycerides and diglycerides, can be modified to form an active polyol through epoxidation and transesterification processes for significant use in PU [150]. The modified polyol can be totally used to synthesize aliphatic PU and even PU foam with the absence of petrochemical-based polyol, indicating that the palm oil-based polyol has excellent reacting activity with isocyanates [151,152].
Generally, different kinds of PU products can be prepared by using the different bio-polyols from VOs, due to the abundant functional sites on the VOs chains (such as ester groups and C=C bonds). Apart from polyols, diols and active –OH terminal polymer can also be obtained to react with isocyanate for preparing PU, or to react with other fatty acids for preparing non-isocyanate PUs. Thus, the VOs can not only used to replace petroleum-based polyols, but also can be modified and employed for eliminating isocyanates for PU, which can supply a new path for producing bio-based PU.

3.1.2. Cashew Nut Shell Liquid

Cashew nut shell liquid (CNSL) is natural aromatic oil from cashew nuts, and can be used for the extraction of cardanol, which is a natural phenol containing a 15-carbon hydrocarbon chain in the middle and one, two or three non-conjugated double bonds, as depicted in Figure 7A [153,154,155]. Cardanol can be functionalized to prepare bio-polyols via esterification, epoxidation and hydrogenation processes, etc., and then reacted with isocyanates to synthesize PU (Figure 7B). The PU commonly presents good mechanical (compression strength of 471 kPa), thermal (higher than 250 °C) and fire properties. Wang et al. epoxidized cardanol to obtain polyol via a thiol-based click reaction [156]. The obtained polyol presents a high hydroxyl number, which can react with hexamethylene diisocyanate (HDI) trimers to synthesize thermosetting PU. The results show that the increasing number of hydroxyl groups in polyols has a positive effect on the crosslinking density of PU. In addition, the CNSL is able to be directly employed as polyol to prepare PU foams.

3.1.3. Terpene

Terpene is a kind of natural hydrocarbons, which is widely found in limonene, pinene, myrcene, etc. [157]. It possesses a high reactivity due to the existence of relatively active C=C bonds in the various linear, cyclic and even polycyclic chains [158]. As one of the earliest used natural polymers, it derives two main functional polymers, including polyisoprene and polyolefins [159,160]. Among the derivatives of terpene, limonene can be the most important, which has been a commercial monoterpene from citrus fruits, as depicted in Figure 7C [161]. It can be facilely modified into bio-polyols through epoxidation or the thiol-ene click process under ultraviolet light (Figure 7C) or catalyst (Figure 7D), followed by the polymerization process to produce PU and NIPU, as well as other similar polymers, as illustrated in Figure 7E [157,162]. Bahr et al. prepared a terpene-based NIPU from cyclic limonene dicarbonate. Firstly, the limonene was epoxidized and followed by catalytic carbonation with CO2 to prepare cyclic limonene dicarbonate without ester groups. Then, the cyclic limonene dicarbonate was reacted with diamines in a non-isocyanate environment. The stiffness (Young’s modulus of 4100 MPa) and glass transition temperatures (62 °C) of NIPU might be enhanced significantly by increasing amine functionality, which reacts with cyclic limonene dicarbonate. It implies that the limonene dioxide can play a key role in synthesizing PU [163,164,165].
Figure 7. (A) Chemical structure of cardanol in cashew nut shell; (B) Polyols derived from cashew nut shells were used as the main building blocks of polyurethane. Reproduced with permission [156]. Copyright 2018, American Chemical Society. (C) Synthesis of limonene polyols by thiol-ene “click” chemistry. Reproduced with permission [162]. Copyright 2014, Springer. (D) Synthesis path from limonene to isocyanate polyurethane. Reproduced with permission [164]. Copyright 2012, Royal Society of Chemistry. (E) The application of limonene for different polymers. Reproduced with permission [157]. Copyright 2020, Science Press.
Figure 7. (A) Chemical structure of cardanol in cashew nut shell; (B) Polyols derived from cashew nut shells were used as the main building blocks of polyurethane. Reproduced with permission [156]. Copyright 2018, American Chemical Society. (C) Synthesis of limonene polyols by thiol-ene “click” chemistry. Reproduced with permission [162]. Copyright 2014, Springer. (D) Synthesis path from limonene to isocyanate polyurethane. Reproduced with permission [164]. Copyright 2012, Royal Society of Chemistry. (E) The application of limonene for different polymers. Reproduced with permission [157]. Copyright 2020, Science Press.
Polymers 16 01889 g007

3.1.4. Rosin

Rosin is a promising biomass for PU synthesis due to the merits of abundance, easy modification, and unique structure with cycloaliphatic, aromatic groups and hydrophenanthrene ring [166,167]. It is commonly used to prepare PU with excellent mechanical rigidity, heat resistance and adhesion [168]. Before the employment in PU, it is usually used directly or modified via isomerization and esterification processes [169,170]. Similar to the limonene, rosin can be also modified to prepare rosin-based cyclic carbonate, which is used to synthesize NIPU with amines [171]. This obviously broadens the raw materials for preparing NIPU in a green way. Notably, rosin is particularly added as a feedstock during the PU preparation process, which can endow PU with some interesting properties. It is reported that the heat resistance, viscoelasticity and antibacterial properties of PU can be increased by adding rosin. The PU presents an antibacterial rate above 99.2% against S.aureus [172]. Additionally, because the hydrophenanthrene ring of rosin is non-planar, it can supply a possibility of conformational transformation under the function of pressure. Thus, the rosin-based PU can possess excellent deformation and stretching strength [166,173]. The designed PU can present significant shape memory property at a high recovery rate. Zhang et al. reported this kind of PU by using a rosin-based diol as chain extender. It is found that the PU is good at shape memory, with an extensively strain higher than 1000% [174]. There is no doubt that these excellent mechanical and antibacterial properties can expand the application of PU in packaging areas.

3.2. PU from Lignin

There are a number of works of literature about the application of lignin and the derivatives in PU. Therefore, the lignin-based PU will be summarized briefly. As one of the most vital biomasses, lignin is easy to obtain from various resources, such as wood, grass, plant straw, etc. [175]. Unlike VOs, lignin and its derivatives do not compete with the food supply. Furthermore, there are abundant aromatic groups, phenolic and alcoholic hydroxyl, methoxy, and some other active groups [176]. As a comparison, lignin can be a better choice from an either resource- or from chemical structure-based perspective. As shown in Figure 8A, lignin mainly consists of the monomers of coniferyl, p-coumaryl and sinapyl alcohol. By changing the structures of monomers via oxygen radicals, lignin can present complex crosslinks with significant strength and hardness, as depicted in Figure 8B [177,178,179,180]. The modified lignin and the derivatives can be employed as polyol to prepare lignin-based PU [181,182]. Lignin can also indeed be added directly to PU during the polymerization process to supply multi-functional properties, such as antibacterial, oxidation and ultraviolet resistance, and even flame retardancy [183,184,185].
For the lignin from wood, the hydroxyl groups on lignin chains are able to react with isocyanates directly, under mild conditions, to obtain PU [186]. It is reported that lignin can be pre-treated via a solvent fractionation process to purify the complex heterogeneous structure and extract the homogeneous fractions of lignin [187,188]. Wang et al. investigated the function of softwood Kraft lignin fractions in lignin-based PU. The fractions were used as primary hydroxyl groups. It was shown that the rigidity and deformation resistance of PU could be enhanced by increasing the molecular weight of lignin [189]. Unfortunately, the obtained PU products are usually brittle and hard, by using lignin as the single polyol. To improve the viscoelasticity of PU, the other polyols are employed as secondary polyols, such as polyether, poly(ethylene glycol), and polyester, etc. [190,191]. This can provide a new path to design the mixture of lignin and synthesized polyols.
To improve the reactivity of lignin, it can be modified through various processes, including oxidation, esterification, isocyanate modification, oxypropylation, etc., as illustrated in Figure 8B. Among them, oxidation can be the most employed method to increase the alcoholic hydroxyl groups of lignin, as well as the functionality [192,193]. Zhang et al. mixed the ozone oxidized lignin with polyethylene glycol as the polyol mixture, to react with isocyanate for preparing PU (Figure 8C). It implies that the oxidized lignin presents a high hydroxyl content with good reactivity. The obtained oxidized lignin/polyethylene glycol PU shows a high degree of crosslinking with good performance [194]. Wang et al. purified lignin via co-solvent enhanced lignocellulosic fractionation pretreatment to obtain lignin-based products with narrow distribution and relatively low molecular weight. Then, the obtained lignin-based products were mixed with secondary hydroxyl groups (aliphatic diols and polyethers) to synthesize PU. It was found that secondary hydroxyl groups interacted with lignin molecules in different degrees, resulting in significant mechanical and thermal properties for PU [195]. Thus, lignin is surly a key candidate feedstock for PU, with unique characteristics [196,197,198]. Besides the roles of polyols, lignin can be used as fillers to modify PU elastomer.
Figure 8. (A) The main alcohol structure in lignin; blue is octanol, green is coniferol, red is coumarol. (B) Common methods for lignin hydroxyl modification. Reproduced with permission [177]. Copyright 2020, Wiley. (C) Schematic diagram of lignin-based polyurethane synthesis reaction, internal structure diagram of LPU and OLPU. Reproduced with permission [194]. Copyright 2019, Elsevier.
Figure 8. (A) The main alcohol structure in lignin; blue is octanol, green is coniferol, red is coumarol. (B) Common methods for lignin hydroxyl modification. Reproduced with permission [177]. Copyright 2020, Wiley. (C) Schematic diagram of lignin-based polyurethane synthesis reaction, internal structure diagram of LPU and OLPU. Reproduced with permission [194]. Copyright 2019, Elsevier.
Polymers 16 01889 g008
Notably, lignin from plant straw shows different properties due to the lignocellulosic biomass from agricultural crops, including soybean [199], rape [200], wheat [201] and corn straws [202]. The lignin-based bio-polyols can be obtained by extracting products from these straws via a liquefaction process, which mainly consists of decomposition, esterification, and condensation polymerization processes [203]. Most of the liquefaction products are polyesters or polyether polyols with rich hydroxyl groups, indicating a good reactivity with isocyanates [204]. Hu et al. extracted bio-polyols from soybean straw by glycerol. The bio-polyols (hydroxyl numbers of 440–540 mg KOH g−1) are easily reacted with isocyanates to synthesize PU with good performance. The obtained PU can possess significant biodegradability [205,206]. It is suggested that the bio-polyols from plant straw can accomplish both the recycling of waste straws and the green synthesis of PU, which properly meets the needs of an environmentally friendly process.
With regard to the use of renewable resources and sustainable development, it is notable that there is a growing interest in the preparation of polymer materials using biomass as a raw material. The bio-based resource source is generally rich and the chemical structure is complex, with a large number of phenol hydroxyl and alcohol hydroxyl active groups. This allows for polycondensation or graft copolymerization chemical reactions. In addition to the extraction of a variety of valuable products, polyols and active -OH end polymers, which can be achieved through different means, can also be utilized in the synthesis of polyurethane. Some bio-based resource can also be directly involved in the synthesis of polyols, such as castor oil.

4. Biomacromolecules for Polyurethane

The synthesis of novel bioactive molecule-based polyurethanes has attracted more and more attention. Up till now, the most frequently used biomacromolecules are mainly polysaccharides and protein for PU synthesis and modification.

4.1. PU from Polysaccharides

Polysaccharides are carbohydrate polymers with a large molecular weight, and usually exist in the form of a main chain followed by side chain. Various carbohydrate polymers, including cellulose, starch, and cyclodextrin, etc., from plants, and chitosan from animals, have been used to prepare PU for wide applications in packaging, biomedicine, printing, coatings and adhesives, etc. [207,208,209]. Because of the branched structure and functional group of carbohydrates, it can be used as the composite material/filler in PU polymer matrices by covalent linkage with isocyanate to form a crosslinking network. Meanwhile, owing to the merits of renewable, non-toxicity, abundance and biodegradability, polysaccharides have been recommended as a promising candidate for preparing bio-PU with biodegradability [210].

4.1.1. Cellulose for PU

Cellulose has been the most important, intriguing and abundant natural polymer, which consists of linear and long-chain glucose. It is abundant in hydroxyl groups, which supplies multiple points for possible H-bonding functions with other polymers [211,212]. According to the micro-morphology, the main cellulose products are cellulose fiber (micro- and nano-scale), cellulose nanocrystal (CNC), and bacterial cellulose (BC) [213,214], all of which have been devoted to WPU modification with excellent biodegradability, mechanical properties, and renewability [215,216,217]. Owing to the challenge of cellulose dissolving at a molecular level, it is hard to obtain a homogenous cellulose solution for polymerization during PU synthesis. Therefore, the different cellulose products are mainly used as fillers to modify PU, which are attested to be an effective way to improve the performance of PU [218,219,220].
It is reported that the cellulose nanofiber (CNF) can be introduced into WPU via the chemical grafting and physical blending processes, respectively. It is found that hydrogen bonding plays a key role to improve the performance of PU for physical blending, while a chemical bonding forms with PU chains for chemical grafting, as depicted in Figure 9A. Commonly, the chemical structure changes little, due to the insolubility of cellulose, and then the thermal stability is similar for PU and CNF-modified PU. Notably, CNF contributes to the tensile strength (~41.8 MPa, higher than the pure PU of ~26.3 MPa) of PU significantly [221].
As for CNC, it can be the most attractive biomass in nanoscales. It presents a rod-like nano-morphology with high crystallinity, which benefits the thermal and mechanical properties due to the hydrogen bonds’ interaction in nano-levels with PU chains. Our previous works have demonstrated this by using CNC from waste paper to reinforce WPU, as shown in Figure 9B [222,223]. CNC can also be modified further to endow WPU with novel properties, including electrical [224] and water resistance [225], antimicrobial properties, etc. [226]. In addition, CNC can also be mixed with VOs to modify WPU to present multi-functional properties, such as high hardness and compatibility, elastic recovery, and even good adhesion [227]. Zhang et al. combined CNC with carbon nanotubes as mixing fillers to modify WPU with high compressibility. This suggests that this composite can be used to prepare piezoresistive sensor with excellent sensitivity. It can hold great potential application in artificial electronic skin, as depicted in Figure 9C [224].
BC is a unique cellulose product from bacteria strains. It plays an important role in medicine, food and tissue engineering areas. Owing to the strong inter-chain hydrogen bonds, BC presents 3D nanofiber networks with significant mechanical properties in vivo. The unique structure and properties provide the possibility for applying BC as a hydrogel to cure wounds [228]. Urbina et al. prepared a gelatinous and translucent BC/WPU composite membrane with shape memory property. The shape memory property is adjusted by temperature, and the shape recovery ratio can be as high as 92.8 ± 6.3%. It is an excellent candidate for biomedical products, as illustrated in Figure 9D [229].
Figure 9. (A) WPU modified by cellulose nanofibers for application of wood coating; photograph of WPU, which contains 0.1 wt% of nanocelllose by chemical modification; photograph and schematic diagram of bending state of nanocellulose-modified polyurethane film; schematic diagram of hydrogen bond formation between PU macromolecule chain and nanocellulose (0.1 wt%) prepared by physical blending method, and the chemical bond by chemical modification method; the FTIR and TG curves of the two modified PU films; the tensile strength of physical modification of polyurethane (PM PU) and chemical modification of polyurethane (CM PU) with different nanocellulose content. Reproduced with permission [221]. Copyright 2019, MDPI. (B) AFM phase images and TG curves of WPU modified by cellulose nanocrystal. Reproduced with permission [223]. Copyright 2019, Elsevier. (C) Schematic illustration of WPU modified by cellulose nanocrystal/carbon nanotube application of piezoresistive sensors; a three-dimensional map of resistance and the inset is photograph of artificial electronic skin. Reproduced with permission [224]. Copyright 2020, Elsevier. (D) Fold-unfold shape memory test process and shape recovery curve in water, illustrated with wet film photograph and SEM image of cross-section 10,000. Reproduced with permission [229]. Copyright 2019, Elsevier.
Figure 9. (A) WPU modified by cellulose nanofibers for application of wood coating; photograph of WPU, which contains 0.1 wt% of nanocelllose by chemical modification; photograph and schematic diagram of bending state of nanocellulose-modified polyurethane film; schematic diagram of hydrogen bond formation between PU macromolecule chain and nanocellulose (0.1 wt%) prepared by physical blending method, and the chemical bond by chemical modification method; the FTIR and TG curves of the two modified PU films; the tensile strength of physical modification of polyurethane (PM PU) and chemical modification of polyurethane (CM PU) with different nanocellulose content. Reproduced with permission [221]. Copyright 2019, MDPI. (B) AFM phase images and TG curves of WPU modified by cellulose nanocrystal. Reproduced with permission [223]. Copyright 2019, Elsevier. (C) Schematic illustration of WPU modified by cellulose nanocrystal/carbon nanotube application of piezoresistive sensors; a three-dimensional map of resistance and the inset is photograph of artificial electronic skin. Reproduced with permission [224]. Copyright 2020, Elsevier. (D) Fold-unfold shape memory test process and shape recovery curve in water, illustrated with wet film photograph and SEM image of cross-section 10,000. Reproduced with permission [229]. Copyright 2019, Elsevier.
Polymers 16 01889 g009

4.1.2. Starch for PU

Starch is not only a daily necessity in food for humans, but also the key natural polymer for chemical and material industries, due to the impressive physicochemical properties [230]. The introduction of starch into PU possesses significant superiorities, such as outstanding mechanical properties, elongation to break and modulus, biodegradability, which can extend the application ranges of PU in food, medicine and papermaking areas, etc. [231,232,233]. The hydrophilic nature of starch promotes the dispersion of it in WPU, as well as the preparation of NIPU composites [234,235]. However, one of the key problems is the micro-phase separation of physical incompatibility for starch/PU hybrids. It is suggested that the employment of a mixture of polyester (hydrophobic) and polyether polyol (hydrophilic) with starch can solve this problem. The prepared starch–polyurethane hybrid presents a good performance of elongation at break (187%), Young’s modulus (383 MPa) and contact angle (112°), and good transparency, as depicted in Figure 10A,B [236]. Similar to cellulose, starch also presents nanocrystals with a typical diameter (average equivalent diameter in the range 25–40 nm), which can strengthen the thermal stability and mechanical properties of WPU [210,237,238].
In addition, cyclodextrin is an enzyme produced from starch, and contains 6, 7 and 8 glucose units of cyclodextrin, which are called α-, β- and γ-cyclodextrin, respectively. Because it is composed of glucose, it has no toxic side effects, and can be absorbed by the body. Therefore, it is highly valued in the pharmaceutical field, used widely as a drug filler and binder, with the general properties of starch. Polysaccharide, as a multihydroxyl compound, can be conveniently used as a crosslinking agent in PU. Lee et al. introduced a β-cyclodextrin molecule, effectively migrated into the main chain of WPU via a sol–gel reaction to reinforce the thermal and mechanical properties, and biodegradability [239]. The β-cyclodextrin molecule consisted of d-glucopyranose groups linked by α-1,4-glycosidic bonds. Due to the presence of the external hydroxyl group of β-cyclodextrin, it has a hydrophobic inner cavity and a hydrophilic outer cavity. At the same time, it can be used as a multifunctional crosslinker to form chemical hybrids with organic polymers. Konieczny et al. synthesized stable PU films using the diisocyanate and an acetylated original, partially hydrolyzed amylopectin/white dextrin as a crosslinking agent [240].

4.1.3. Chitosan for PU

Owing to the characteristics of non-toxicity, antibacterial, biological activity, cytocompatibility and degradability, chitosan have attracted huge attention, with great promise as biomaterials in both academic and industrial contexts [241,242]. There are active organic groups, including hydroxyl and amino groups, to support the chemical modification with other polymers, especially for PU [243,244,245,246]. For example, chitosan has been employed as a chain extender for preparing WPU with self-healing properties. The research found that the self-healing efficiency is about 47% for the tensile strength, which is much higher than that of WPU without chitosan (only 4%) [244]. This can be due to the exchange reaction between the hydroxyl group on the chitosan chain and the carbamate on the PU chain at high temperatures. Lin et al. used chitosan as a cross-linker to prepare biodegradable PU through a dynamic Schiff reaction at room temperature. As a result, the PU gels are sensitive to low pH values, with excellent self-healing properties of nearly 100% recovery after damage, as depicted in Figure 10C,D [247]. Furthermore, chitosan can be in situ incorporated in the form of nanoflakes into PU. It supports the PU with good shape recovery (~100% shape recovery ratio), an increase in tensile strength (~130%) and elongation at break (~500%). Meanwhile, the chitosan/PU composites also present a relatively high thermal stability of 360 °C. Therefore, chitosan can be introduced to PU synthesis either in chemical method or physically blending, resulting in different PU products. The chitosan supports PU to have great potential as biomedical materials [248].

4.1.4. Sodium Alginate for PU

Sodium alginate is a kind of hydrophilic anionic polysaccharide, which is commonly used as seasoning in foodstuffs [249]. Owing to its water solubility, biological compatibility, and multiple physicochemical properties, it has been applied in biomedicine, packaging and printing, and even tissue engineering [250,251]. Wang et al. prepared sodium alginate/WPU composites by physically blending a sodium alginate aqueous solution with VO-based WPU. It is revealed that the hydrophobicity, mechanical properties, and thermal stability improve, and the elongation at break decreases by increasing the sodium alginate content in the composite. The changes of the composite properties can be caused by the improvement of crosslink density from the sodium alginate molecules with many active points, which provide the possibility for PU chains to form interpenetrating networks, as illustrated in Figure 10E. When the sodium alginate content in WPU increases from 5% to 20%, the tensile strength and Young’s modulus of the sodium alginate/WPU composite film gradually increases to the maximum of 26.5 MPa and 511 MPa (Figure 10F) [250].
Figure 10. (A,B) Schematic diagram showing the physical crosslinking of starch and PU dispersion; stress–strain curves of HAGS (high amylose glycerol plasticized starch film), LDPE (low density polyethylene) and HAGS-AEEPU (Anionic polyurethane) films. Reproduced with permission [236]. Copyright 2018, Elsevier. (C,D) WPU modified by chitosan. Schematic representation of waterborne polyurethane with chitosan crosslinkers to form self-healing hydrogel or cryogel; the strain of self-healing hydrogel was up to 500%. Reproduced with permission [247]. Copyright 2020, Wiley. (E,F) Schematic diagram of process of sodium alginate (SA) modified WPU; stress–strain curves of PU films with different SA content. Reproduced with permission [250]. Copyright 2019, Elsevier.
Figure 10. (A,B) Schematic diagram showing the physical crosslinking of starch and PU dispersion; stress–strain curves of HAGS (high amylose glycerol plasticized starch film), LDPE (low density polyethylene) and HAGS-AEEPU (Anionic polyurethane) films. Reproduced with permission [236]. Copyright 2018, Elsevier. (C,D) WPU modified by chitosan. Schematic representation of waterborne polyurethane with chitosan crosslinkers to form self-healing hydrogel or cryogel; the strain of self-healing hydrogel was up to 500%. Reproduced with permission [247]. Copyright 2020, Wiley. (E,F) Schematic diagram of process of sodium alginate (SA) modified WPU; stress–strain curves of PU films with different SA content. Reproduced with permission [250]. Copyright 2019, Elsevier.
Polymers 16 01889 g010

4.1.5. Glucomannan

Glucomannan, extracted from tubers, bulbs, cork, and roots, belongs to a kind of mannans family of polysaccharides, with structural and storage functions in plants [252]. It is mainly derived from konjac glucomannan, which is composed of β-1, 4 linked D-mannan and d-glucose monomers arranged in a block of mixed residues forming a linear chain [253]. The glucomannan can be successfully incorporated in PU to synthesize nanocomposites, while the synergistic effects of filler (glucomannan) and substrate (PU) provide outstanding mechanical properties due to the intermolecular hydrogen bond between PU and glucomannan [254]. It was found that PU and various types of glucomannan formed an interpenetrating polymer network-based graft material. The nitro derivatives of konjac glucomannan and castor oil-derived PU formed a semi-interpenetrating polymer network sheet by hydrogen bonding, which had good compatibility [253].

4.2. PU from Protein

Soy protein (SP) is a by-product of soybean oil industry. It is composed of amino acids, carboxylic acids, hydroxyl groups and mercaptan chemical functional groups. It is a kind of polymer, containing amino acids, which is considered to be a new raw material to replace polymers as petroleum resources to synthesize plastics [255]. Meanwhile, SP can be used as a filler in PU plastic, with many desirable properties, such as abundant resources, non-toxicity, low cost, biodegradability and renewability. Madbouly et al. prepared hydrolyzable PU/SP shape-memory polymer blends with improved crystallization behavior in environmentally friendly aqueous dispersions, which can be applied in multifunctional potential biomedical contexts [256]. Zhang et al. developed castor oil-based PU foam by a self-rising method, using water as foaming agent and SP as a reinforcing filler. The result showed that the crosslinking degree of the composite foams increased with the reaction of amino and hydroxyl groups on SP with polyisocyanates [255].
Not only plant proteins but also animal proteins can be used as raw materials for PU synthesis. A total of 90% of the feather structure is made up of a protein called feather keratin. Saucedo-Rivalcoba et al. showed that keratin is grafted into the PU structure by the amino and carboxyl groups of the amino acid structure to form polyurethane-keratin membranes. And beyond that, amino acids are the basic component of proteins, containing amino and carboxyl functional groups. It can also be involved in the synthesis of PU. Chan-Chan et al. reported the synthesis of more biocompatible and biodegradable PU, using unmodified amino acids as chain extenders via urea and amide formation, such as L-arginine, glycine or L-aspartic acid [257]. Among them, arginine and glycine have good effects on the properties of polyurethane materials, such as high molecular weight, deformation ability, tensile strength and high storage modulus. According to their good mechanical properties and biodegradability, PU prepared from amino acid has been extensively used in biomedical applications.
In general, the structure of biomacromolecule polysaccharides is complex and difficult to dissolve, which has led to their primary use in modifying polyurethane. This is achieved through chemical grafting and physical blending, which introduces the polysaccharides into the polyurethane chain and improves their mechanical and thermal properties.The protein is integrated into the polyurethane structure through the formation of covalent bonds between the amino and carboxyl groups of the amino acid structure, resulting in the synthesis of bio-based polyurethane products with biodegradability. As illustrated in Table 3, a diverse range of bio-based raw materials can be employed in the synthesis of polyurethane, resulting in a multitude of product types with distinct properties.

5. Sustainability for Polyurethane

Polyurethane is one of the most versatile materials in the world. It can be produced in a variety of forms and used in a variety of markets. According to the data, the polyurethane market was ranked sixth in global polymer production in 2018, with a total production of approximately 65.5 billion dollars. The market is expected to reach $105.2 billion by 2025 [258]. However, according to the data, recycled industrial waste and consumption waste polyurethane only accounts for 30% of the total content. Because the recycling of resources is largely driven by economic factors, cheap new materials recovered at the end of the life cycle are proving attractive. These products and processes must be further developed to make renewable technologies more competitive. Ultimately, the polyurethane feedstock minimizes dependence on non-renewable petrochemical products.
From an economic point of view, the total annual output of waste-based feedstocks and bio-based feedstocks and the cost of recycling these materials are shown in Table 4. We collected data and sorted out the total annual output of waste-based raw materials and bio-based raw materials and the cost of recycling these materials for a comparative analysis from an economic perspective. As can be seen from Table 4, waste-based raw materials generally have low recycling prices and relatively high annual yields. Due to the type and planting area of bio-based raw materials, the yield of some raw materials, such as CNSL, terpene, and rosin is low, so the price is higher. However, chitosan, cellulose and lignin are lower, due to the wide source and high yield. Generally speaking, the recycling of a single material does not require the removal of different plastics, greatly reducing the complexity of the process, and the homogeneous parts that can be easily removed from the end product all have a high recycling potential. There are many kinds of PU products produced from renewable recycled raw materials, including adhesives, elastomers, foams and coatings. At the same time, all kinds of PU products have excellent properties and can be widely used in various fields, such as cushion packaging, oil sorption, fireproof materials, biomedical implants.
In general, despite the recycling of resources to achieve the high value of the economic cycle, the industry is still facing many challenges. From the upstream design, collection and classification, to the last part of the regeneration process, the regeneration process is developing rapidly at present, but the real expansion of this market requires concerted action across the value chain. Upstream classification and sorting systems tend to be more efficient, and downstream access to high-quality raw materials will be less efficient.
Table 4. Recovery of different feedstocks of PU and their novel applications.
Table 4. Recovery of different feedstocks of PU and their novel applications.
FeedstockYield
(t/y)
Price ($/t)ProductsApplicationsRef.
PET19.9 × 106418–448PU FoamsCushion packaging
Oil sorption
Fireproof materials
[42,43,259]
PU12 × 106150–300PU FoamsFoaming agent[89,260]
PC9.8 × 106700–900PU films
WPU
Biomimetic anti-fouling
Coatings
[119,121]
VOs90 × 1061800–2000PU FoamsMattress
Ink
[3,261,262]
CNSL0.03 × 106450–700PU Foams
coatings
Building materials
Anticorrosive coatings
[153,263]
Terpene0.45 × 1063000–3500PU FoamsFlame retardancy[264]
Rosin0.65 × 1061800–2000ElastomerShape memory material[174]
Lignin50 × 10650–200Elastomer
WPU
Anti-bacterial coatings
Shoe Sole
[157,189]
Cellulose100 × 10620–45WPU
Coatings
Wood antibacterial[215]
Starch30 × 106450–600PU filmsBiomedical materials[239]
Chitosan100 × 1092500–3000PU gels
Elastomer
Shape memory material
Biomedical materials
[247,248]
Sodium alginate0.23 × 106600–800WPUBiomedical materials[252]
Glucomannan0.25 × 106500–1000WPUFood packaging[252]
Protein0.2 × 106200–350ElastomerBiomedical materials[256]

6. Conclusions and Prospects

Polyurethane is a kind of multipurpose resin widely used in many fields, in which one of main chemical composition is polyols. Apparently, the main resources of polyurethane are still petroleum feedstock. The scarcity of petroleum oil resources pulls the feed-stock of polyurethane toward recycled and renewable raw materials. These urgent requirements are driving research into more advanced polyurethane designs and have elevated effective approaches regarding the diversity of raw materials. To improve the environmental pollution problems, the two main ways summarized in this work are the solution, namely reusing waste petroleum-based polymers, and developing novel feedstock from natural resources. Therefore, the recent developments in PU preparation from waste, biomass and biomacromolecule resources have been reviewed concisely.
The PET, PU, and PC wastes are recovered and degraded into value-added polyols as the raw material to synthesize waste-based PU products, i.e., foams, films, coatings, and waterborne dispersion. Furthermore, the bio-based polyols are obtained from vegetable oil, cashew nut shell liquid, plant straw, terpene, lignin and some new biodegradable materials: polylactic acid and polycaprolactone. The bio-based resource sources, with a large number of phenol hydroxyl and alcohol hydroxyl active groups, are used to synthesize PU by polycondensation or graft copolymerization chemical reactions, and the bio-based PU products similarly have outstanding thermal stability and mechanical properties, applied to eco-friendly adhesive coatings and degradable packaging. Biomacromolecule polysaccharides are achieved through chemical grafting and physical blending, which introduces the polysaccharides into the PU chain and improves their mechanical and thermal properties. Meanwhile, the modification of PU is conducted by various biomacromolecules, such as polysaccharides and protein, to develop multi-purpose practical applications such as food packaging, biodegradable plastics and biomedical engineering. The utilization of recyclable and biomass materials has the effect of reducing the consumption of disposable mineral resources, such as petroleum, and the application of harmful chemicals. Furthermore, it provides a sustainable method of producing polyurethane. The incorporation of bio-based polyols renders polyurethane materials biodegradable, thereby mitigating the environmental contamination associated with PU products.
In a word, the use of waste-based recycled products or bio-based renewable raw materials in the production of PU not only satisfies the maximum utilization of resources, but also reduces the production costs for the circular economy principles. At the same time, it achieves the requirement of promoting green environmental protection and reducing pollution. Sustainable PU products, from recycling and renewable resources, have already gained huge attention for their promising application in printing and packaging, biomedical materials, buildings, and wearable electronics. While the future of PU, from the perspective of recycling and renewable resources, still holds great opportunity, it will also consist of tremendous challenges.
  • An important challenge for recycling resources is improved extraction, degradation and transformation, which are valuable building blocks that will be optimized at the cost of both the performance and monomers. At the same time, the purity and stability of the monomers prepared from recycling resources will also be an important consideration in the future research of PU.
  • From the upstream design, collection and classification, to the last part of the regeneration process, the regeneration process is developing rapidly at present, but the real expansion of this market requires concerted action across the value chain. Upstream classification and sorting systems tend to be more efficient, and downstream access to high-quality raw materials will be less efficient.
  • A crucial, and sometimes underestimated, synthesis route is the need to ensure the raw materials are compatible with current equipment condition. Although it is still early days to directly quantify and compare traditional petrochemical PU, there is enough research to show that recyclable materials have gradually become part of the view of the PU production, particularly regarding environmental pollution and fossil resource depletion.
  • Strengthen the connection between the solid waste classification and recovery system and the renewable resource recovery system, connect the renewable resource industry chain with the waste industry chain, improve the resource utilization rate, and strive to explore the industrialization development of the circular economy of polyurethane.

Author Contributions

Conceptualization, data curation and writing—original draft preparation M.P.; validation and funding acquisition, C.F.; writing—review and editing and resources X.Z.; data curation, Y.L.; visualization, D.W.; formal analysis, W.L.; supervision, L.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support provided by the National Key Research and Development Program of China [No. 2021YFD1600402], the National Natural Science Foundation of China [Grant No. 52372093]; Key R & D Program of Shaanxi Province [No. 2022SF-168]; Central Guidance on Local Science and Technology Development Fund of Shaanxi Province [No. 2020-ZYYD-NCC-9]; Shaanxi Provincial Department of Education Collaborative Innovation Center Project [20JY052]; National Natural Science Foundation of China (No. 52072075); the Open Foundation of Key Laboratory of Auxiliary Chemistry and Technology for Chemical Industry, Ministry of Education, Shaanxi University of Science and Technology (No. KFKT2022-04) and Shaanxi Collaborative Innovation Center of Industrial Auxiliary Chemistry and Technology, Shaanxi University of Science and Technology (No. KFKT2022-04).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, H.R.; Li, T.; Li, J.; Zhao, R.H.; Ding, A.; Xu, F.J. Structural Engineering of Polyurethanes for Biomedical Applications. Prog. Polym. Sci. 2024, 151, 101803. [Google Scholar] [CrossRef]
  2. Poopalam, K.D.; Ismail, T.N.M.T.; Hanzah, N.A.; Alias, A.H.; Wahab, N.A.; Ibrahim, Z.; Subramaniam, V.; Armylisas, A.N.; Idris, Z. Utilization of oil palm biomass and Polyurethanes as sustainable construction materials: A review. Dev. Built. Environ. 2024, 17, 100380. [Google Scholar] [CrossRef]
  3. Gunatillake, P.A.; Dandeniyage, L.S.; Adhikari, R.; Bown, M.; Shanks, R.; Adhikari, B. Advancements in the Development of Biostable Polyurethanes. Polym. Rev. 2019, 59, 391–417. [Google Scholar] [CrossRef]
  4. Cevher, D.; Surdem, S. Polyurethane Adhesive Based on Polyol Monomers BHET and BHETA Depolymerised from PET Waste. Int. J. Adhes. 2021, 105, 102799. [Google Scholar] [CrossRef]
  5. Li, L.; Yu, T.L. Curing comparison and performance investigation of polyurethane concrete with retarders. Constr. Build. Mater. 2022, 326, 126883. [Google Scholar] [CrossRef]
  6. Jiang, R.J.; Zheng, X.Y.; Zhu, S.S.; Li, W.Y.; Zhang, H.W.; Liu, Z.H.; Zhou, X. Recent Advances in Functional Polyurethane Chemistry: From Structural Design to Applications. Chemistryselect 2023, 8, e202204132. [Google Scholar] [CrossRef]
  7. Stadler, B.M.; Hinze, S.; Tin, S.; de Vries, J.G. Hydrogenation of Polyesters to Polyether Polyols. Chemsuschem 2019, 12, 4082–4087. [Google Scholar] [CrossRef]
  8. Liu, H.; Zhang, Z.P.; Yu, X.H.; Kan, S.Y.; Luo, Y.; Han, K.B.; Liang, Y.Z.; Gao, J.P. Preparation of Polyol From Waste Polyethylene Terephthalate (PET) and Its Application to Polyurethane (PU) Modified Asphalt. Constr. Build Mater. 2024, 427, 136286. [Google Scholar] [CrossRef]
  9. Grdadolnik, M.; Drincic, A.; Oreski, A.; Onder, O.C.; Utrosa, P.; Pahovnik, D.; Zagar, E. Insight into Chemical Recycling of Flexible Polyurethane Foams by Acidolysis. ACS Sustain. Chem. Eng. 2022, 10, 1323–1332. [Google Scholar] [CrossRef]
  10. Alshabebi, A.S.; Alrashed, M.M.; El Blidi, L.; Haider, S. Preparation of Bio-Based Polyurethane Coating from Citrullus colocynthis Seed Oil: Characterization and Corrosion Performance. Polymers 2024, 16, 214. [Google Scholar] [CrossRef]
  11. Stuhr, R.; Bayer, P.; Stark, C.B.W.; Jacobi von Wangelin, A. Light-Driven Waste-To-Value Upcycling: Bio-Based Polyols and Polyurethanes from the Photo-Oxygenation of Cardanols. Chemsuschem 2021, 14, 3325–3332. [Google Scholar] [CrossRef] [PubMed]
  12. Cui, Z.; Han, J.W.; Sun, J.; Chen, Y.; Wang, X.G.; Gu, X.Y.; Zhang, S. Preparation of Aliphatic Bio-Based Polyester Polyols for Water-Blown and Halogen-Free Flame Retardant Rigid Polyurethane Foam. Polym. Degrad. Stab. 2024, 226, 110832. [Google Scholar] [CrossRef]
  13. Kazemi, M.; Kabir, S.F.; Fini, E.H. State of the Art in Recycling Waste Thermoplastics and Thermosets and Their Applications in Construction. Resour. Conserv. Recycl. 2021, 174, 105776. [Google Scholar] [CrossRef]
  14. Caki, S.M.; Risti, I.S.; Cincovi, M.M.; Nikolic, N.C.; Nikolic, L.B.; Cvetinov, M.J. Synthesis And Properties Biobased Waterborne Polyurethanes from Glycolysis Product of PET Waste and Poly(Caprolactone) Diol. Prog. Org. Coat. 2017, 105, 111–122. [Google Scholar] [CrossRef]
  15. Zhou, X.; Wang, C.; Fang, C.; Yu, R.; Li, Y.; Lei, W. Structure and thermal properties of various alcoholysis products from waste poly(ethylene terephthalate). Waste Manag. 2019, 85, 164–174. [Google Scholar] [CrossRef] [PubMed]
  16. Simón, D.; Borreguero, A.M.; Lucas, A.D.; Gutiérrez, C.; Rodríguez, J.F. Sustainable Polyurethanes: Chemical Recycling to Get It; Springer: Cham, Switzerland, 2014; Volume 32, pp. 229–260. [Google Scholar] [CrossRef]
  17. Lee, D.K.; Tsai, H.B.; Tsai, R.S.; Chen, P.H. Preparation and Properties of Transparent Thermoplastic Segmented Polyurethanes Derived from Different Polyols. Polym. Eng. Sci. 2007, 47, 695–701. [Google Scholar] [CrossRef]
  18. Savani, N.G.; Naveen, T.; Dholakiya, B.Z. A review on the synthesis of maleic anhydride based polyurethanes from renewable feedstock for different industrial applications. J. Polym. Res. 2023, 30, 175. [Google Scholar] [CrossRef]
  19. Noreen, A.; Zia, K.M.; Zuber, M.; Tabasum, S.; Zahoor, A.F. Bio-Based Polyurethane: An Efficient and Environment Friendly Coating Systems: A Review. Prog. Org. Coat. 2016, 91, 25–32. [Google Scholar] [CrossRef]
  20. Su, Y.; Ma, S.Q.; Wang, B.B.; Xu, X.W.; Feng, H.Z.; Hu, K.Z.; Zhang, W.Q.; Zhou, S.C.; Weng, G.S.; Zhu, J. High-Performance Castor Oil-Based Polyurethane Thermosets: Facile Synthesis and Properties. React. Funct. Polym. 2023, 183, 105496. [Google Scholar] [CrossRef]
  21. Petrovic, Z.S. Polyurethanes from Vegetable Oils. Polym. Rev. 2008, 48, 109–155. [Google Scholar] [CrossRef]
  22. Sarim, M.; Alavi Nikje, M.M.; Dargahi, M. Synthesis and characterization of polyurethane rigid foam by using feedstocks received from renewable and recyclable resources. J. Porous Mater. 2023, 30, 1337–1356. [Google Scholar] [CrossRef]
  23. Zia, K.M.; Bhatti, H.N.; Bhatti, I.A. Methods for Polyurethane and Polyurethane Composites, Recycling and Recovery: A Review. React. Funct. Polym. 2007, 67, 675–692. [Google Scholar] [CrossRef]
  24. Yuan, L.Y.; Zhou, W.; Shen, Y.; Li, Z.B. Chemically Recyclable Polyurethanes Based on Bio-Renewable Gamma-Butyrolactone: From Thermoplastics to Elastomers. Polym. Degrad. Stabil. 2022, 204, 110116. [Google Scholar] [CrossRef]
  25. Peng, Y.; Yuan, X.; Jiang, L.; Yang, J.; Liu, Z.; Zhao, Y.; Chen, H. The Fabricating Methods, Properties and Engineering Applications of Foamed Concrete with Polyurethane: A Review. Int. J. Environ. Sci. Technol. 2022, 20, 2293–2312. [Google Scholar] [CrossRef]
  26. Pfister, D.P.; Xia, Y.; Larock, R.C. Recent Advances in Vegetable Oil-Based Polyurethanes. Chemsuschem 2011, 4, 703–717. [Google Scholar] [CrossRef] [PubMed]
  27. Xu, C.P.; Paone, E.; Rodriguez-Padron, D.; Luque, R.; Mauriello, F. Reductive Catalytic Routes towards Sustainable Production of Hydrogen, Fuels and Chemicals from Biomass Derived Polyols. Renew. Sustain. Energy Rev. 2020, 127, 109852. [Google Scholar] [CrossRef]
  28. Wang, M.; Liu, M.J.; Lu, J.M.; Wang, F. Photo Splitting of Bio-Polyols and Sugars to Methanol and Syngas. Nat. Commun. 2020, 11, 1083. [Google Scholar] [CrossRef] [PubMed]
  29. Paciorek-Sadowska, J.; Borowicz, M.; Isbrandt, M. New Poly(lactide-urethane-isocyanurate) Foams Based on Bio-Polylactide Waste. Polymers 2019, 11, 481. [Google Scholar] [CrossRef] [PubMed]
  30. Khorshidi, F.H.; Najafi, S.K.; Najafi, F.; Pizzi, A.; Behrooz, R.; Sandberg, D. The Extraction of Polyol for The Synthesis of Lignin-Based Polyurethane Coatings—A Review. Wood Mater. Sci. Eng. 2024, 19, 794–802. [Google Scholar] [CrossRef]
  31. Klempner, D.; Frisch, K.C. Advances in Urethane: Science and Technology; CRC Press: Boca Raton, FL, USA, 1998. [Google Scholar]
  32. Sonnenschein, M.F. Polyurethanes; Wiley: New York, NY, USA, 2014. [Google Scholar]
  33. Kazemi, M.; Fini, E.H. State of the art in the application of functionalized waste polymers in the built environment. Resour. Conserv. Recycl. 2021, 177, 105967. [Google Scholar] [CrossRef]
  34. Zhou, X.; Deng, J.; Fang, C.; Yu, R.; Lei, W.; He, X.; Zhang, C. Preparation and characterization of lysozyme@carbon nanotubes/waterborne polyurethane composite and the potential application in printing inks. Prog. Org. Coat. 2020, 142, 105600. [Google Scholar] [CrossRef]
  35. Karpati, L.; Szarka, G.; Hartman, M.; Vargha, V. Oligoester and Polyester Production via Acido-Alcoholysis of PET Waste. Period. Polytech. Chem. Eng. 2018, 62, 336–344. [Google Scholar] [CrossRef]
  36. Bicerano, J. Crystallization of Polypropylene and Poly(Ethylene Terephthalate). Polym. Rev. 2008, 38, 391–479. [Google Scholar] [CrossRef]
  37. Smith, R.L.; Takkellapati, S.; Riegerix, R. Recycling of Plastics in the United States: Plastic Material Flows and Polyethylene Terephthalate (PET) Recycling Processes. ACS Sustain. Chem. Eng. 2022, 10, 2084–2096. [Google Scholar] [CrossRef] [PubMed]
  38. Welle, F. Twenty Years of PET Bottle to Bottle Recycling-an Overview. Resour. Conserv. Recycl. 2011, 55, 865–875. [Google Scholar] [CrossRef]
  39. Caputto, M.D.D.; Navarro, R.; Valentin, J.L.; Marcos-Fernandez, A. Chemical Upcycling of Poly(Ethylene Terephthalate) Waste: Moving to A Circular Model. J. Polym. Sci. 2022, 60, 3269–3283. [Google Scholar] [CrossRef]
  40. Duldner, M.; Dolana, S.; Bartha, E.; Teodorescu, F.; Slabu, A.; Tincu, R.; Sarbu, A.; Ghiurea, M.; Gavrila, A.M.; Ghebaur, A.; et al. Nano-composite semi-flexible polyurethane foams containing montmorillonite intercalated/exfoliated in situ during the synthesis of the intermediate polyester-polyols obtained from PET wastes. Eur. Polym. J. 2023, 196, 112280. [Google Scholar] [CrossRef]
  41. Senra, E.M.; Silva, A.L.N.; Pacheco, E.B.A.V. A Review of Waterborne Polyurethane Coatings and Adhesives with Polyester Polyol from Poly(Ethylene Terephthalate) Waste. J. Polym. Environ. 2023, 31, 3719–3739. [Google Scholar] [CrossRef]
  42. Luo, X.L.; Li, Y.B. Synthesis and Characterization of Polyols and Polyurethane Foams from PET Waste and Crude Glycerol. J. Polym. Environ. 2014, 22, 318–328. [Google Scholar] [CrossRef]
  43. Atta, A.M.; Brostow, W.; Datashvili, T.; El-Ghazawy, R.A.; Lobland, H.E.H.; Hasan, A.R.M.; Perez, J.M. Porous Polyurethane Foams Based on Recycled Poly(Ethylene Terephthalate) for Oil Sorption. Polym. Int. 2013, 62, 116–126. [Google Scholar] [CrossRef]
  44. Sangalang, A.; Bartolome, L.; Kim, D.H. Generalized Kinetic Analysis of Heterogeneous PET Glycolysis: Nucleation-Controlled Depolymerization. Polym. Degrad. Stab. 2015, 115, 45–53. [Google Scholar] [CrossRef]
  45. Fang, P.T.; Liu, B.; Xu, J.L.; Zhou, Q.; Zhang, S.J.; Ma, J.Y.; Lu, X.M. High-Efficiency Glycolysis of Poly(Ethylene Terephthalate) by Sandwich-Structure Polyoxometalate Catalyst with Two Active Sites. Polym. Degrad. Stab. 2018, 156, 22–31. [Google Scholar] [CrossRef]
  46. Kim, H.T.; Kim, J.K.; Cha, H.G.; Kang, M.J.; Lee, H.S.; Khang, T.U.; Yun, E.J.; Lee, D.H.; Song, B.K.; Park, S.J.; et al. Biological Valorization of Poly(ethylene terephthalate) Monomers for Upcycling Waste PET. ACS Sustain. Chem. Eng. 2019, 7, 19396–19406. [Google Scholar] [CrossRef]
  47. Imek, B.; Uygunolu, T.; Korucu, H.; Kocakerim, M.M. Analysis of the Effects of Dioctyl Terephthalate Obtained from Polyethylene Terephthalate Wastes on Concrete Mortar: A Response Surface Methodology Based Desirability Function Approach Application. J. Clean. Prod. 2017, 170, 437–445. [Google Scholar] [CrossRef]
  48. Liu, L.; Hao, P.P.; Zhang, R.Q.; Zhang, Y.J.; Zhou, Q.; Lu, X.M.; Yan, D.X.; Li, Y.; Shi, C.Y. Preparation of High-Performance Polyurethane Elastomers Via Direct Use of Alcoholyzed Waste PET by High Molecular Weight Polyester Diols. J. Appl. Polym. Sci. 2023, 140, e54238. [Google Scholar] [CrossRef]
  49. Yoshida, S.; Hiraga, K.; Takehana, T.; Taniguchi, I.; Yamaji, H.; Maeda, Y.; Toyohara, K.; Miyamoto, K.; Kimura, Y.; Oda, K.A. Bacterium that Degrades and Assimilates Poly(Ethylene Terephthalate). Science 2016, 351, 1196–1199. [Google Scholar] [CrossRef] [PubMed]
  50. Wei, R.; Breite, D.; Song, C.; Grasing, D.; Ploss, T.; Hille, P.; Schwerdtfeger, R.; Matysik, J.; Schulze, A.; Zimmermann, W. Biocatalytic Degradation Efficiency of Postconsumer Polyethylene Terephthalate Packaging Determined by Their Polymer Microstructures. Adv. Sci. 2019, 6, 1900491. [Google Scholar] [CrossRef] [PubMed]
  51. Esquer, R.; Garcia, J.J. Metal-Catalysed Poly(Ethylene) Terephthalate and Polyurethane Degradations by Glycolysis. J. Organomet. Chem. 2019, 902, 120972. [Google Scholar] [CrossRef]
  52. Scremin, D.M.; Miyazaki, D.Y.; Lunelli, C.E.; Silva, S.A.; Zawadzki, S.F. Pet Recycling by Alcoholysis Using a New Heterogeneous Catalyst: Study and Its Use in Polyurethane Adhesives Preparation. Macromol. Symp. 2019, 383, 1800027. [Google Scholar] [CrossRef]
  53. Kathalewar, M.; Dhopatkar, N.; Pacharane, B.; Sabnis, A.; Raut, P.; Bhave, V. Chemical Recycling of PET Using Neopentyl Glycol: Reaction Kinetics and Preparation of Polyurethane Coatings. Prog. Org. Coat. 2013, 76, 147–156. [Google Scholar] [CrossRef]
  54. Nunes, S.; da Silva, M.J.V.; da Silva, D.C.; Freitas, A.R.; Rosa, F.A.; Rubira, A.F.; Muniz, E.C. PET Depolymerisation In Supercritical Ethanol Catalysed by [Bmim][BF4]. Rsc Adv. 2014, 4, 20308–20316. [Google Scholar] [CrossRef]
  55. Raheem, A.B.; Noor, Z.Z.; Hassan, A.; Abd Hamid, M.K.; Samsudin, S.A.; Sabeen, A.H. Current Developments in Chemical Recycling of Post-Consumer Polyethylene Terephthalate Wastes for New Materials Production: A Review. J. Clean. Prod. 2019, 225, 1052–1064. [Google Scholar] [CrossRef]
  56. Pham, C.T.; Nguyen, B.T.; Nguyen, M.T.; Nguyen, T.H.; Hoang, C.N.; Nguyen, N.N.; Lee, P.C.; Kim, J.; Hoang, D. The Advancement of Bis(2-Hydroxyethyl)Terephthalate Recovered from Post-Consumer Poly(Ethylene Terephthalate) Bottles Compared to Commercial Polyol for Preparation of High Performance Polyurethane. J. Ind. Eng. Chem. 2021, 93, 196–209. [Google Scholar] [CrossRef]
  57. Mendiburu-Valor, E.; Larraza, I.; Echeverria-Altuna, O.; Harismendy, I.; Pena-Rodriguez, C.; Eceiza, A. Thermoset polyurethanes from biobased and recycled components. J. Polym. Environ. 2023, 31, 4946–4959. [Google Scholar] [CrossRef]
  58. Shamsi, R.; Sadeghi, G.M.M.; Vahabi, H.; Seyfi, J.; Sheibani, R.; Zarrintaj, P.; Laoutid, F.; Saeb, M.R. Hopes Beyond PET Recycling: Environmentally Clean and Engineeringly Applicable. J. Polym. Environ. 2019, 27, 2490–2508. [Google Scholar] [CrossRef]
  59. Shamsi, R.; Sadeghi, G.M.M. Novel Polyester Diol Obtained From PET Waste and Its Application in The Synthesis of Polyurethane and Carbon Nanotube-Based Composites: Swelling Behavior and Characteristic Properties. RSC Adv. 2016, 6, 38399–38415. [Google Scholar] [CrossRef]
  60. Scheirs, J.; Kaminsky, W. Feedstock Recycling and Pyrolysis of Waste Plastics: Converting Waste Plastics into Diesel and Other Fuels; Focus on Catalysts; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2006; pp. 161–192. [Google Scholar] [CrossRef]
  61. Azeem, M.; Attallah, O.A.; Tas, C.E.; Fournet, M.B. All Green Microwave Assisted 99% Depolymerisation of Polyethylene Terephthalate into Value Added Products via Glycerol Pre-treatment and Hydrolysis Reaction. J. Polym. Environ. 2023, 32, 303–315. [Google Scholar] [CrossRef]
  62. Strukil, V. Highly Efficient Solid-State Hydrolysis of Waste Polyethylene Terephthalate by Mechanochemical Milling and Vapor-Assisted Aging. Chemsuschem 2020, 14, 330–338. [Google Scholar] [CrossRef]
  63. Kang, M.J.; Yu, H.J.; Jegal, J.; Kim, H.S.; Cha, H.G. Depolymerization of PET into Terephthalic Acid in Neutral Media Catalyzed by the ZSM-5 Acidic Catalyst. Chem. Eng. J. 2020, 398, 125655. [Google Scholar] [CrossRef]
  64. Liu, Q.L.; Li, R.S.; Fang, T. Investigating and Modeling PET Methanolysis under Supercritical Conditions by Response Surface Methodology Approach. Chem. Eng. J. 2015, 270, 535–541. [Google Scholar] [CrossRef]
  65. Jadhav, A.L.; Malkar, R.S.; Yadav, G.D. Zn-and Ti-Modified Hydrotalcites for Transesterification of Dimethyl Terephthalate with Ethylene Glycol: Effect of the Metal Oxide and Catalyst Synthesis Method. Acs Omega 2020, 5, 2088–2096. [Google Scholar] [CrossRef] [PubMed]
  66. Sadeghi, G.M.M.; Shamsi, R.; Sayaf, M. From Aminolysis Product of PET Waste to Novel Biodegradable Polyurethanes. J. Polym. Environ. 2011, 19, 522–534. [Google Scholar] [CrossRef]
  67. Chattopadhyay, D.K.; Webster, D.C. Thermal stability and flame retardancy of polyurethanes. Prog. Polym. Sci. 2009, 34, 1068–1133. [Google Scholar] [CrossRef]
  68. Delavardea, A.; Savin, G.; Derkenne, P.; Boursier, M.; Morales-Cerrada, R.; Nottelet, B.; Pinaud, J.; Caillol, S. Sustainable polyurethanes: Toward new cutting-edge opportunities. Prog. Polym. Sci. 2024, 151, 101805. [Google Scholar] [CrossRef]
  69. Shaaban, A.; Elsabbagh, A.M. Crashworthiness Optimization of Impact Attenuators Constructed of Polyurethane Foam. Int. J. Automot. Technol. 2022, 23, 389–401. [Google Scholar] [CrossRef]
  70. Liu, Q.; Gao, D.; Xu, W. Effect of Polyurethane Non-Transparent Coating Process on Paint Film Performance Applied on Modified Poplar. Coatings 2022, 12, 39. [Google Scholar] [CrossRef]
  71. Choi, S.M.; Shin, E.J.; Zo, S.M.; Rao, K.M.; Seok, Y.J.; Won, S.Y.; Han, S.S. Revised Manuscript with Corrections: Polyurethane-Based Conductive Composites: From Synthesis to Applications. Int. J. Mol. Sci. 2022, 23, 1938. [Google Scholar] [CrossRef] [PubMed]
  72. Nering, K.; Kowalska-Koczwara, A. Determination of Vibroacoustic Parameters of Polyurethane Mats for Residential Building Purposes. Polymers 2022, 14, 314. [Google Scholar] [CrossRef] [PubMed]
  73. Gu, X.H.; Zhu, S.W.; Liu, S.W.; Liu, Y. Analysis of the Influencing Factors of the Efficient Degradation of Waste Polyurethane and Its Scheme Optimization. Polymers 2023, 15, 2337. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, B.Y.; Westman, Z.; Richardson, K.; Lim, D.; Stottlemyer, A.L.; Farmer, T.; Gillis, P.; Hooshyar, N.; Vlcek, V.; Christopher, P.; et al. Polyurethane Foam Chemical Recycling: Fast Acidolysis with Maleic Acid and Full Recovery of Polyol. ACS Sustain. Chem. Eng. 2024, 12, 4435–4443. [Google Scholar] [CrossRef]
  75. Gu, X.; Wang, X.; Guo, X.; Liu, S.; Lou, C.; Liu, Y. Study on Efficient Degradation of Waste PU Foam. Polymers 2023, 15, 2359. [Google Scholar] [CrossRef] [PubMed]
  76. Chung, Y.C.; Choi, J.W.; Choi, N.E.; Chun, B.C. Endurance of linear and cross-linked shape memory polyurethane under rigorous hydrolysis conditions. Fiber. Polym. 2009, 10, 576–582. [Google Scholar] [CrossRef]
  77. Gu, X.H.; Wang, X.Y.; Guo, X.Y.; Liu, S.W.; Li, Q.; Liu, Y. Study and Characterization of Regenerated Hard Foam Prepared by Polyol Hydrolysis of Waste Polyurethane. Polymers 2023, 15, 1445. [Google Scholar] [CrossRef] [PubMed]
  78. Dai, Z.Y.; Hatano, B.; Kadokawa, J.; Tagaya, H. Effect of Diaminotoluene on the Decomposition of Polyurethane Foam Waste in Superheated Water. Polym. Degrada. Stab. 2002, 76, 179–184. [Google Scholar] [CrossRef]
  79. Trzebiatowska, P.J.; Benešb, H.; Datta, J. Evaluation of the Glycerolysis Process and Valorisation of Recovered Polyol in Polyurethane Synthesis. React. Funct. Polym. 2019, 139, 25–33. [Google Scholar] [CrossRef]
  80. Najafi-Shoa, S.; Barikani, M.; Ehsani, M.; Ghaffari, M.; Vandalvand, M. Efficient and Eco-Friendly UV-Cured Polyurethane Coating: Harnessing Thiol-Yne Systems for Corrosion Protection. Mater. Today Commun. 2024, 39, 109037. [Google Scholar] [CrossRef]
  81. Paciorek-Sadowska, J.; Czupryński, B.; Liszkowska, J. Glycolysis of Rigid Polyurethane-Polyisocyanurate Foams. J. Elastomers Plast. 2016, 48, 340–353. [Google Scholar] [CrossRef]
  82. Beneš, H.; Roesner, J.; Holler, P.; Synkova, H.; Kotek, J.; Horak, Z. Glycolysis of Flexible Polyurethane Foam in Recycling of Car Seats. Polym. Adv. Technol. 2007, 18, 149–156. [Google Scholar] [CrossRef]
  83. Motokucho, S.; Nakayama, Y.; Morikawa, H.; Nakatani, H. Environment-Friendly Chemical Recycling of Aliphatic Polyurethanes by Hydrolysis in a CO2-Water System. J. Appl. Polym. Sci. 2018, 135, 45897. [Google Scholar] [CrossRef]
  84. Mahoney, L.R.; Weiner, S.A.; Ferris, F.C. Hydrolysis of Polyurethane Foam Waste. Environ. Sci. Technol. 1974, 8, 135–139. [Google Scholar] [CrossRef]
  85. Hung, C.S.; Barlow, D.E.; Varaljay, V.A.; Drake, C.A.; Crouch, A.L.; Russell, J.N.; Nadeau, L.J.; Crookes-Goodson, W.J.; Biffinger, J.C. The Biodegradation of Polyester and Polyester Polyurethane Coatings Using Papiliotrema Laurentii. Int. Biodeterior. Biodegrad. 2019, 139, 34–43. [Google Scholar] [CrossRef]
  86. Yamane, S.; Watanabe, R.; Ata, S.; Mizukado, J.; Shinzawa, H. Solvent-Induced Degradation of Polyurethane Studied by Two-Dimensional (2D) Infrared (IR) Correlation Spectroscopy. Vib. Spectrosc. 2020, 108, 103062. [Google Scholar] [CrossRef]
  87. Simón, D.; Borreguero, A.M.; De Lucas, A.; Rodríguez, J.F. Glycolysis of Viscoelastic Flexible Polyurethane Foam Wastes. Polym. Degrad. Stab. 2015, 116, 23–35. [Google Scholar] [CrossRef]
  88. Kemona, A.; Piotrowska, M. Polyurethane Recycling and Disposal: Methods and Prospects. Ploymers 2020, 12, 1752. [Google Scholar] [CrossRef]
  89. Simón, D.; de Lucas, A.; Rodríguez, J.F.; Borreguero, A.M. Flexible Polyurethane Foams Synthesized Employing Recovered Polyols from Glycolysis: Physical and Structural Properties. J. Appl. Polym. Sci. 2017, 134, 45087. [Google Scholar] [CrossRef]
  90. Datta, J.; Haponiuk, J.T. Advanced Coating of Interior of Tanks for Rising Environmental Safety-Novel Applications of Polyurethanes. Pol. Marit. Res. 2008, 15, 8–13. [Google Scholar] [CrossRef]
  91. Ko, J.Y.; Zarei, M.; Lee, S.G.; Cho, K.L. Single-Phase Recycling of Flexible Polyurethane Foam by Glycolysis and Oxyalkylation: Large-Scale Industrial Evaluation. ACS Sustain. Chem. Eng. 2023, 11, 10074–10082. [Google Scholar] [CrossRef]
  92. Shin, S.R.; Maim, V.D.; Lee, D.S. Chemical Recycling of Used Printed Circuit Board Scraps: Recovery and Utilization of Organic Products. Processes 2019, 7, 22. [Google Scholar] [CrossRef]
  93. Ionescu, M. Chemistry and Technology of Polyols for Polyurethanes; Rapra Technology Limited: Shawbury, UK, 2005. [Google Scholar]
  94. Shin, S.R.; Kim, H.N.; Liang, J.Y.; Lee, S.H.; Lee, D.S. Sustainable Rigid Polyurethane Foams Based on Recycled Polyols from Chemical Recycling of Waste Polyurethane Foams. J. Appl. Polym. Sci. 2019, 136, 47916. [Google Scholar] [CrossRef]
  95. Vanbergen, T.; Verlent, I.; De Geeter, J.; Haelterman, B.; Claes, L.; De Vos, D. Recycling of Flexible Polyurethane Foam by Split-Phase Alcoholysis: Identification of Additives and Alcoholyzing Agents to Reach Higher Efficiencies. Chemsuschem 2020, 13, 3835–3843. [Google Scholar] [CrossRef]
  96. Zhu, P.; Cao, Z.B.; Chen, Y.; Zhang, X.J.; Qian, G.R.; Chu, Y.L.; Zhou, M. Glycolysis Recycling of Rigid Waste Polyurethane Foam from Refrigerators. Environ. Technol. 2014, 35, 2676–2684. [Google Scholar] [CrossRef]
  97. Simon, D.; Rodriguez, J.F.; Carmona, M.; Serrano, A.; Borreguero, A.M. Glycolysis of Advanced Polyurethanes Composites Containing Thermoregulating Microcapsules. Chem. Eng. J. 2018, 350, 300–311. [Google Scholar] [CrossRef]
  98. Kanaya, K.; Takahashi, S. Decomposition of Polyurethane Foams by Alkanolamines. J. Appl. Polym. Sci. 2010, 51, 675–682. [Google Scholar] [CrossRef]
  99. Modesti, M.; Simioni, F. Chemical Recycling of Reinforced Polyurethane from the Automotive Industry. Polym. Eng. Sci. 1996, 36, 2173–2178. [Google Scholar] [CrossRef]
  100. Nikje, M.M.A.; Garmarudi, A.B.; Idris, A.B. Polyurethane Waste Reduction and Recycling: From Bench to Pilot Scales. Des. Monomers Polym. 2011, 14, 395–421. [Google Scholar] [CrossRef]
  101. Troev, K.; Atanasov, V.I.; Tsevi, R. Chemical Degradation of Polyurethanes I: Degradation of Microporous Polyurethane Elastomer by Dimethyl Phosphonate. Polym. Degrad. Stab. 2000, 67, 159–165. [Google Scholar] [CrossRef]
  102. Troev, K.; Grancharov, G.; Tsevi, R.; Tsekova, A. A Novel Approach to Recycling of Polyurethanes: Chemical Degradation of Flexible Polyurethane Foams by Triethyl Phosphate. Polymer 2000, 41, 7017–7022. [Google Scholar] [CrossRef]
  103. Troev, K.; Tsekova, A.; Tsevi, R. Chemical Degradation of Polyurethanes II: Degradation of Microporous Polyurethane Elastomer by Phosphoric Acid Esters. J. Appl. Polym. Sci. 2000, 76, 886–893. [Google Scholar] [CrossRef]
  104. Mitova, V.; Grancharov, G.; Molero, C.; Borreguero, A.M.; Troev, K.; Rodriguez, J.F. Chemical Degradation of Polymers (Polyurethanes, Polycarbonate and Polyamide) by Esters of H-phosphonic and Phosphoric Acids. J. Macromol. Sci. Part A 2013, 50, 774–795. [Google Scholar] [CrossRef]
  105. Abolins, A.; Yakushin, V.; Vilsone, D. Properties of Polyurethane Coatings Based on Linseed Oil Phosphate Ester Polyol. J. Renew. Mater. 2018, 6, 737–745. [Google Scholar] [CrossRef]
  106. Heinen, M.; Gerbase, A.E.; Petzhold, C.L. Vegetable Oil-Based Rigid Polyurethanes and Phosphorylated Flame-Retardants Derived from Epoxydized Soybean Oil. Polym. Degrad. Stab. 2014, 108, 76–86. [Google Scholar] [CrossRef]
  107. Nikje, M.M.A.; Nikrah, M.; Mohammadi, F.H.A. Microwave-Assisted Polyurethane Bond Cleavage via Hydroglycolysis Process at Atmospheric Pressure. J. Cell. Plast. 2008, 44, 367–380. [Google Scholar] [CrossRef]
  108. Simón, D.; Borreguero, A.M.; de Lucas, A.; Rodríguez, J.F. Recycling of Polyurethanes from Laboratory to Industry, a Journey towards the Sustainability. Waste Manag. 2018, 76, 147–171. [Google Scholar] [CrossRef]
  109. Gerlock, J.; Braslaw, J.; Zinbo, M. Polyurethane Waste Recycling; Glycolysis and Hydroglycolysis of Water-Blown Foams. Ind. Eng. Chem. Process. Des. Dev. 1984, 23, 545–552. [Google Scholar] [CrossRef]
  110. Ansari, I.; Singh, P.; Mittal, A.; Mahato, R.I.; Chitkara, D. 2,2-Bis(Hydroxymethyl) Propionic Acid Based Cyclic Carbonate Monomers and Their (Co)Polymers as Advanced Materials for Biomedical Applications. Biomaterials 2021, 275, 120953. [Google Scholar] [CrossRef] [PubMed]
  111. Wei, P.R.; Bhat, G.A.; Darensbourg, D.J. Enabling New Approaches: Recent Advances in Processing Aliphatic Polycarbonate-Based Materials. Angew. Chem. Int. Ed. 2023, 62, e202307507. [Google Scholar] [CrossRef] [PubMed]
  112. Lee, H.W.; Yoo, K.; Borchardt, L.; Kim, J.G. Chemical Recycling of Polycarbonate and Polyester without Solvent and Catalyst: Mechanochemical Methanolysis. Green Chem. 2024, 26, 2087–2093. [Google Scholar] [CrossRef]
  113. Fehér, Z.; Németh, R.; Kiss, J.; Balterer, B.; Verebélyi, K.; Iván, B.; Kupai, J. A Silica-Supported Organocatalyst for Polycarbonate Methanolysis Under Mild and Economic Conditions. Chem. Eng. J. 2024, 485, 149832. [Google Scholar] [CrossRef]
  114. Kim, D.; Kim, B.K.; Cho, Y.; Han, M.; Kim, B.S. Kinetics of Polycarbonate Glycolysis in Ethylene Glycol. Ind. Eng. Chem. Res. 2009, 48, 685–691. [Google Scholar] [CrossRef]
  115. Nikje, M.M.A.; Askarzadeh, M. Chemical Recycling of Polycarbonate Wastes into Bisphenol a by Using Green Solvent Composition. Polimery 2013, 58, 292–294. [Google Scholar] [CrossRef]
  116. Darensbourg, D.J.; Yeung, A.D. A Concise Review of Computational Studies of The Carbon Dioxide–Epoxide Copolymerization Reactions. Polym. Chem. 2014, 5, 3949–3962. [Google Scholar] [CrossRef]
  117. Huang, Y.C.; Huang, Y.H.; Chen, L.Y.; Dai, C.A.; Dai, S.H.A.; Chen, Y.H.; Wu, C.H.; Jeng, R.J. Robust Thermoplastic Polyurethane Elastomers Prepared from Recycling Polycarbonate. Polymer 2021, 212, 123296. [Google Scholar] [CrossRef]
  118. Jang, T.; Kim, H.J.; Jang, J.B.; Kim, T.H.; Lee, W.; Seo, B.; Ko, W.B.; Lim, C.S. Synthesis of Waterborne Polyurethane Using Phosphorus-Modified Rigid Polyol and its Physical Properties. Polymer 2021, 13, 432. [Google Scholar] [CrossRef] [PubMed]
  119. Li, P.C.; Cai, W.H.; Li, X.; Wang, K.B.; Zhou, L.; You, T.X.; Wang, R.; Chen, H.; Zhao, Y.C.; Wang, J. Preparation of Phospholipid-Based Polycarbonate Urethanes for Potential Applications of Blood-Contacting Implants. Regen. Biomater. 2020, 7, 491–504. [Google Scholar] [CrossRef] [PubMed]
  120. Orgiles-Calpena, E.; Aran-Ais, F.; Torro-Palau, A.M.; Orgiles-Barcelo, C. Novel Polyurethane Reactive Hot Melt Adhesives Based on Polycarbonate Polyols Derived from CO2 For the Footwear Industry. Int. J. Adhes. Adhes. 2016, 70, 218–224. [Google Scholar] [CrossRef]
  121. Garcia-Pacios, V.; Costa, V.; Colera, M.; Martin-Martinez, J.M. Waterborne Polyurethane Dispersions Obtained with Polycarbonate of Hexanediol Intended for Use as Coatings. Prog. Org. Coat. 2011, 71, 136–146. [Google Scholar] [CrossRef]
  122. Benes, H.; Paruzel, A.; Hodan, J.; Trhlikova, O. Impact of Natural Oil-Based Recycled Polyols on Properties of Cast Polyurethanes. J. Renew. Mater. 2018, 6, 697–706. [Google Scholar] [CrossRef]
  123. Lambeth, R.H.; Henderson, T.J. Organocatalytic Synthesis of (Poly) Hydroxyurethanes from Cyclic Carbonates and Amines. Polymer 2013, 54, 5568–5573. [Google Scholar] [CrossRef]
  124. Delebecq, E.; Pascault, J.P.; Boutevin, B.; Ganachaud, F. On the Versatility of Urethane/Urea Bonds: Reversibility, Blocked Isocyanate, and Non-isocyanate Polyurethane. Chem. Rev. 2013, 113, 80–118. [Google Scholar] [CrossRef]
  125. Yang, X.X.; Wang, S.B.; Liu, X.X.; Huang, Z.; Huang, X.J.; Xu, X.; Liu, H.; Wang, D.; Shang, S.B. Preparation of Non-isocyanate Polyurethanes from Epoxy Soybean Oil: Dual Dynamic Networks to Realize Self-healing and Reprocessing under Mild Conditions. Green Chem. 2021, 23, 6349–6355. [Google Scholar] [CrossRef]
  126. Tran, M.H.; Lee, E.Y. Production of polyols and polyurethane from biomass: A review. Environ. Chem. Lett. 2023, 21, 2199–2223. [Google Scholar] [CrossRef]
  127. Wang, L.; Gao, Q.; Li, J.; Jiang, L.; Wang, S.H.; Fan, H.J.; Chen, Y.; Yan, J.; Xiang, J. Tuning the Hydrophobicity of Bio-Based Waterborne Polyurethane by Leveraging a Diol Derived from Oleic Acid. Ind. Crops Prod. 2022, 187, 115400. [Google Scholar] [CrossRef]
  128. Kaikade, D.S.; Sabnis, A.S. Polyurethane foams from vegetable oil-based polyols: A review. Polym. Bull. 2023, 80, 2239–2261. [Google Scholar] [CrossRef] [PubMed]
  129. Vahabi, H.; Rastin, H.; Movahedifar, E.; Antoun, K.; Brosse, N.; Saeb, M.R. Flame Retardancyof Bio-Based Polyurethanes: Opportunities and Challenges. Polymers 2020, 12, 1234. [Google Scholar] [CrossRef] [PubMed]
  130. Hormaiztegui, M.E.V.; Aranguren, M.I.; Mucci, V.L. Synthesis and Characterization of a Waterborne Polyurethane Made from Castor Oil and Tartaric Acid. Eur. Polym. J. 2018, 102, 151–160. [Google Scholar] [CrossRef]
  131. Gurunathan, T.; Chung, J.S. Physicochemical Properties of Amino-Silane-Terminated Vegetable Oil-Based Waterborne Polyurethane Nanocomposites. ACS Sustain. Chem. Eng. 2016, 4, 4645–4653. [Google Scholar] [CrossRef]
  132. Liang, H.Y.; Wang, S.W.; He, H.; Wang, M.Q.; Liu, L.X.; Lu, J.Y.; Zhang, Y.; Zhang, C.Q. Aqueous Anionic Polyurethane Dispersions from Castor Oil. Ind. Crops Prod. 2018, 122, 182–189. [Google Scholar] [CrossRef]
  133. Luong, N.D.; Sinh, L.H.; Minna, M.; Jurgen, W.; Torsten, W.; Matthias, S.; Jukka, S. Synthesis and Characterization of Castor Oil-Segmented Thermoplastic Polyurethane with Controlled Mechanical Properties. Eur. Polym. J. 2016, 81, 129–137. [Google Scholar] [CrossRef]
  134. Dodangeh, F.; Dorraji, M.S.S.; Rasoulifard, M.H.; Ashjari, H.R. Synthesis and Characterization of Alkoxy Silane Modified Polyurethane Wood Adhesive Based on Epoxidized Soybean Oil Polyester Polyol. Compos. Part B 2020, 187, 107857. [Google Scholar] [CrossRef]
  135. Liu, L.X.; Lu, J.Y.; Zhang, Y.; Liang, H.Y.; Liang, D.S.; Jiang, J.Z.; Lu, Q.M.; Quirino, R.L.; Zhang, C.Q. Thermosetting Polyurethanes Prepared with The Aid of a Fully Bio-Based Emulsifier with High Bio-Content, High Solid Content, and Superior Mechanical Properties. Green Chem. 2019, 21, 526–537. [Google Scholar] [CrossRef]
  136. Shendi, H.K.; Omrani, I.; Ahmadi, A.; Farhadian, A.; Babnejad, N.; Nabid, M.R. Synthesis and Characterization of a Novel Internal Emulsifier Derived from Sunflower Oil for the Preparation of Waterborne Polyurethane and Their Application in Coatings. Prog. Org. Coat. 2017, 105, 303–309. [Google Scholar] [CrossRef]
  137. Babanejad, N.; Farhadian, A.; Omrani, I.; Nabid, M.R. Design, Characterization and in Vitro Evaluation of Novel Amphiphilic Block Sunflower Oil-Based Polyol Nanocarrier as a Potential Delivery System: Raloxifene-Hydrochloride as a Model. Mater. Sci. Eng. C 2017, 78, 59–68. [Google Scholar] [CrossRef]
  138. Omrani, I.; Farhadian, A.; Babanejad, N.; Shendi, H.K.; Ahmadi, A.; Nabid, M.R. Synthesis of Novel High Primary Hydroxyl Functionality Polyol from Sunflower Oil using Thiol-Ene Reaction and Their Application in Polyurethane Coating. Eur. Polym. J. 2016, 82, 220–231. [Google Scholar] [CrossRef]
  139. Hajirahimkhan, S.; Xu, C.C.; Ragogna, P.J. UV Curable Coatings of Modified Lignin. ACS Sustain. Chem. Eng. 2018, 6, 14685–14694. [Google Scholar] [CrossRef]
  140. Guo, L.; Huang, S.; Qu, J.Q. Synthesis and Properties of High-Functionality Hydroxyl-Terminated Polyurethane Dispersions. Prog. Org. Coat. 2018, 119, 214–220. [Google Scholar] [CrossRef]
  141. Valverde, C.; Lligadas, G.; Ronda, J.C. Hydroxyl Functionalized Renewable Polyesters Derived From 10-Undecenoic Acid: Polymer Structure and Postpolymerization Modification. Eur. Polym. J. 2018, 105, 68–78. [Google Scholar] [CrossRef]
  142. Cheng, Z.; Li, Q.T.; Yan, Z.; Liao, G.F.; Zhang, B.X.; Yu, Y.M.; Yi, C.F.; Xu, Z.S. Design and Synthesis of Novel Aminosiloxane Crosslinked Linseed Oil-Based Waterborne Polyurethane Composites and Its Physicochemical Properties. Prog. Org. Coat. 2019, 127, 194–201. [Google Scholar] [CrossRef]
  143. Lu, K.T.; Chang, J.P. Synthesis and Antimicrobial Activity of Metal-Containing Linseed Oil-Based Waterborne Urethane Oil Wood Coatings. Polymer 2020, 12, 663. [Google Scholar] [CrossRef]
  144. Chen, R.Q.; Zhang, C.Q.; Kessler, M.R. Anionic Waterborne Polyurethane Dispersion from a Bio-Based Ionic Segment. RSC Adv. 2014, 4, 35476–35483. [Google Scholar] [CrossRef]
  145. Openshaw, K. A Review of Jatropha curcas: An Oil Plant of Unfulfilled Promise. Biomass Bioenergy 2000, 19, 1–15. [Google Scholar] [CrossRef]
  146. Kumar, A.; Sharma, S. An Evaluation of Multipurpose Oil Seed Crop for Industrial Uses (Jatropha curcas L.): A Review. Ind. Crops Prod. 2008, 28, 1–10. [Google Scholar] [CrossRef]
  147. Ling, C.K.; Aung, M.M.; Rayung, M.; Abdullah, L.C.; Lim, H.N.; Noor, I.S.M. Performance of Ionic Transport Properties in Vegetable Oil-Based Polyurethane Acrylate Gel Polymer Electrolyte. ACS Omega 2019, 4, 2554–2564. [Google Scholar] [CrossRef]
  148. Saalah, S.; Abdullah, L.C.; Aung, M.M.; Salleh, M.Z.; Biak, D.R.A.; Basri, M.; Jusoh, E.R.; Mamat, S. Physicochemical Properties of Jatropha Oil-Based Polyol Produced by a Two Steps Method. Molecules 2017, 22, 551. [Google Scholar] [CrossRef] [PubMed]
  149. Saalah, S.; Abdullah, L.C.; Aung, M.M.; Salleh, M.Z.; Biak, D.R.A.; Basri, M.; Jusoh, E.R.; Mamat, S. Colloidal Stability and Rheology of Jatropha Oil-Based Waterborne Polyurethane (JPU) Dispersion. Prog. Org. Coat. 2018, 125, 348–357. [Google Scholar] [CrossRef]
  150. Yusof, N.S.B.; Sapuan, S.M.; Sultan, M.T.H.; Jawaid, M. Manufacturing Process Selection of “Green” Oil Palm Natural Fiber Reinforced Polyurethane Composites Using Hybrid TEA Criteria Requirement and AHP Method for Automotive Crash Box. J. Renew. Mater. 2020, 8, 647–660. [Google Scholar] [CrossRef]
  151. Su, Y.P.; Lin, H.; Zhang, S.T.; Yang, Z.H.; Yuan, T. One-Step Synthesis of Novel Renewable Vegetable Oil-Based Acrylate Prepolymers and Their Application in UV-Curable Coatings. Polymer 2020, 12, 1165. [Google Scholar] [CrossRef] [PubMed]
  152. Sittinun, A.; Pisitsak, P.; Manuspiya, H.; Thiangtham, S.; Chang, Y.H.; Ummartyotin, S. Utilization of Palm Olein-Based Polyol for Polyurethane Foam Sponge Synthesis: Potential as A Sorbent Material. J. Polym. Environ. 2020, 28, 3181–3191. [Google Scholar] [CrossRef]
  153. Ghosh, T.; Karak, N. Cashew Nut Shell Liquid Terminated Self-Healable Polyurethane as an Effective Anticorrosive Coating with Biodegradable Attribute. Prog. Org. Coat. 2020, 139, 105472. [Google Scholar] [CrossRef]
  154. Kathalewar, M.; Sabnis, A. Preparation of Novel CNSL-Based Urethane Polyol via Nonisocyanate Route: Curing with Melamine-Formaldehyde Resin and Structure-Property Relationship. J. Appl. Polym. Sci. 2014, 132, 41391. [Google Scholar] [CrossRef]
  155. Shrestha, M.L.; Ionescu, M.; Wan, X.M.; Bilic, N.; Petrovic, Z.S.; Upshaw, T. Biobased Aromatic-Aliphatic Polyols from Cardanol by Thermal Thiol-Ene Reaction. J. Renew. Mater. 2018, 6, 87–101. [Google Scholar] [CrossRef]
  156. Wang, H.R.; Zhou, Q.X. Synthesis of Cardanol-Based Polyols via Thiol-ene/Thiol-epoxy Dual Click-Reactions and Thermosetting Polyurethanes Therefrom. ACS Sustain. Chem. Eng. 2018, 6, 12088–12095. [Google Scholar] [CrossRef]
  157. Monica, F.D.; Kleij, A.W. From Terpenes to Sustainable and Functional Polymers. Polym. Chem. 2020, 11, 5109–5127. [Google Scholar] [CrossRef]
  158. Touaibia, M.; Boutekedjiret, C.; Perino, S. Natural Terpenes as Building Blocks for Green Chemistry; Springer: Singapore, 2019; pp. 171–195. [Google Scholar] [CrossRef]
  159. Zhu, Y.Q.; Romain, C.; Williams, C.K. Sustainable Polymers from Renewable Resources. Nature 2016, 540, 354–362. [Google Scholar] [CrossRef] [PubMed]
  160. Liu, G.F.; Wu, G.M.; Jin, C.; Kong, Z.W. Preparation and Antimicrobial Activity of Terpene-Based Polyurethane Coatings with Carbamate Group-Containing Quaternary Ammonium Salts. Prog. Org. Coat. 2015, 80, 150–155. [Google Scholar] [CrossRef]
  161. Firdaus, M.; Meier, M.A.R. Renewable Polyamides and Polyurethanes Derived from Limonene. Green Chem. 2013, 15, 370–380. [Google Scholar] [CrossRef]
  162. Gupta, R.K.; Ionescu, M.; Radojcic, D.; Wan, X.; Petrovic, Z.S. Novel Renewable Polyols Based on Limonene for Rigid Polyurethane Foams. J. Polym. Environ. 2014, 22, 304–309. [Google Scholar] [CrossRef]
  163. Bhr, M.; Bitto, A.; Mülhaupt, R. Cyclic Limonene Dicarbonate as a New Monomer for Non-Isocyanate Oligo- and Polyurethanes (NIPU) Based upon Terpenes. Green Chem. 2012, 14, 1447–1454. [Google Scholar] [CrossRef]
  164. Luc, C.; Xavier, F.; Serge, K. Ultrasonic and Catalyst Free Epoxidation of Limonene and Other Terpenes Using Dimethyl Dioxirane in Semi-Batch Conditions. ACS Sustain. Chem. Eng. 2018, 6, 12224–12231. [Google Scholar] [CrossRef]
  165. Wu, G.M.; Kong, Z.W.; Chen, J.; Huo, S.P.; Liu, G.F. Preparation and Properties of Waterborne Polyurethane/Epoxy Resin Composite Coating from Anionic Terpene-Based Polyol Dispersion. Prog. Org. Coat. 2014, 77, 315–321. [Google Scholar] [CrossRef]
  166. Maiti, S.; Ray, S.S.; Kundu, A.K. Rosin: A Renewable Resource for Polymers and Polymer Chemicals. Prog. Polym. Sci. 1989, 14, 297–338. [Google Scholar] [CrossRef]
  167. Wilbon, P.A.; Chu, F.X.; Tang, C.B. Progress in Renewable Polymers from Natural Terpenes, Terpenoids, and Rosin. Macromol. Rapid Commun. 2013, 34, 8–37. [Google Scholar] [CrossRef]
  168. Li, T.T.; Liu, X.Q.; Jiang, Y.H.; Ma, S.Q.; Zhu, J. Bio-Based Shape Memory Epoxy Resin Synthesized from Rosin Acid. Iran. Polym. J. 2016, 25, 1–9. [Google Scholar] [CrossRef]
  169. Xu, X.; Song, Z.Q.; Shang, S.B.; Cui, S.Q.; Rao, X.P. Synthesis and Properties of Novel Rosin-Based Water-Borne Polyurethane. Polym. Int. 2011, 60, 1521–1526. [Google Scholar] [CrossRef]
  170. Hsieh, C.C.; Chen, Y.C. Synthesis of Bio-Based Polyurethane Foam Modified with Rosin Using an Environmentally-Friendly Process. J. Clean. Prod. 2020, 276, 124203. [Google Scholar] [CrossRef]
  171. Liu, G.F.; Wu, G.M.; Chen, J.; Kong, Z.W. Synthesis, Modification and Properties of Rosin-Based Non-Isocyanate Polyurethanes Coatings. Prog. Org. Coat. 2016, 101, 461–467. [Google Scholar] [CrossRef]
  172. Xu, C.A.; Qu, Z.C.; Lu, M.G.; Meng, H.F.; Zhan, Y.J.; Chen, B.; Wu, K.; Shi, J. Effect of Rosin on the Antibacterial Activity against S. aureus and Adhesion Properties of UV-Curable Polyurethane/Polysiloxane Pressure-Sensitive Adhesive. Colloids Surf. A 2021, 614, 126146. [Google Scholar] [CrossRef]
  173. Vevere, L.; Fridrihsone, A.; Kirpluks, M.; Cabulis, U. A Review of Wood Biomass-Based Fatty Acids and Rosin Acids Use in Polymeric Materials. Polymer 2020, 12, 2706. [Google Scholar] [CrossRef] [PubMed]
  174. Zhang, L.S.; Jiang, Y.H.; Xiong, Z.; Liu, X.Q.; Na, H.N.; Zhang, R.Y.; Zhu, J. Highly Recoverable Resin-Based Shape Memory Polyurethenes. J. Mater. Chem. A. 2013, 1, 3263. [Google Scholar] [CrossRef]
  175. Zakzeski, J.; Bruijnincx, P.C.A.; Jongerius, A.L.; Weckhuysen, B.M. The Catalytic Valorization of Lignin for the Production of Renewable Chemicals. Chem. Rev. 2013, 110, 3552–3599. [Google Scholar] [CrossRef]
  176. Rad, E.R.; Vahabi, H.; de Anda, A.R.; Saeb, M.R.; Thomas, S. Bio-epoxy resins with inherent flame retardancy. Prog. Org. Coat. 2020, 135, 608–612. [Google Scholar] [CrossRef]
  177. Ma, X.; Chen, J.; Zhu, J.; Yan, N. Lignin-Based Polyurethane: Recent Advances and Future Perspectives. Macromol. Rapid Commun. 2020, 42, 2000492. [Google Scholar] [CrossRef]
  178. Domenek, S.; Louaifi, A.; Guinault, A.; Baumberger, S. Potential of Lignins as Antioxidant Additive in Active Biodegradable Packaging Materials. J. Polym. Environ. 2013, 21, 692–701. [Google Scholar] [CrossRef]
  179. El Salamouny, S.; Shapiro, M.L.K.S.; Shepard, B.M. Black Tea and Lignin as Ultraviolet Protectants for the Beet Armyworm Nucleopolyhedrovirus. J. Entomol. Sci. 2009, 44, 50–58. [Google Scholar] [CrossRef]
  180. Liu, L.N.; Qian, M.B.; Song, P.A.; Huang, G.B.; Yu, Y.M.; Fu, S.Y. Fabrication of Green Lignin-Based Flame Retardants for Enhancing the Thermal and Fire Retardancy Properties of Polypropylene/Wood Composites. ACS Sustain. Chem. Eng. 2016, 4, 2422–2431. [Google Scholar] [CrossRef]
  181. Ren, L.F.; Zhao, Y.X.; Qiang, T.T.; He, Q.Q. Synthesis of a Biobased Waterborne Polyurethane with Epichlorohydrin-Modified Lignin. J. Dispers. Sci. Technol. 2019, 40, 1499–1506. [Google Scholar] [CrossRef]
  182. Du, J.H.; Wang, H.; Huang, Z.Y.; Liu, X.C.; Yin, X.S.; Wu, J.X.; Lin, W.J.; Lin, X.F.; Yi, G.B. Construction and mechanism study of lignin-based polyurethane with high strength and high self-healing properties. Int. J. Biol. Macromol. 2023, 248, 125925. [Google Scholar] [CrossRef] [PubMed]
  183. Liang, J.; Zhao, Z.; Zhou, T.; Fu, Y.; Tian, G.; Wang, Z. Fabrication of Bio-Based Flexible Polyurethane Foam with Biodegradation by Etherification of Kraft Lignin. ACS Appl. Polym. Mater. 2024, 6, 4441–4448. [Google Scholar] [CrossRef]
  184. Quan, P.; Kiziltas, A.; Gondaliya, A.; Siahkamari, M.; Nejad, M.; Xie, X. Kraft Lignin with Improved Homogeneity Recovered Directly from Black Liquor and Its Application in Flexible Polyurethane Foams. ACS Omega 2022, 7, 16705–16715. [Google Scholar] [CrossRef]
  185. Chen, S.; Zhang, W.; Ye, Y.; Ying, X.; Huang, J.; Li, X. Introduction of Aminated Sodium Lignosulfonate as a Chain Extender for Preparation of High-Performance Waterborne Polyurethane. Int. J. Adhes. Adhes. 2023, 125, 103415. [Google Scholar] [CrossRef]
  186. Cheradame, H.; Detoisien, M.; Gandini, A.; Pla, F.; Roux, G. Polyurethane from Kraft Lignin. Br. Polym. J. 1989, 21, 269–275. [Google Scholar] [CrossRef]
  187. Li, H.; Liang, Y.; Li, P. Conversion of Biomass Lignin to High-Value Polyurethane: A Review. J. Bioresour. Bioprod. 2020, 5, 163–179. [Google Scholar] [CrossRef]
  188. Griffini, G.; Passoni, V.; Suriano, R.; Levi, M.; Turni, S. Polyurethane Coatings Based on Chemically Unmodified Fractionated Lignin. ACS Sustain. Chem. Eng. 2015, 3, 1145–1154. [Google Scholar] [CrossRef]
  189. Wang, Y.Y.; Wyman, C.E.; Cai, C.M.; Ragauskas, A.J. Lignin-Based Polyurethanes from Unmodified Kraft Lignin Fractionated by Sequential Precipitation. ACS Appl. Polym. Mater. 2019, 1, 1672–1679. [Google Scholar] [CrossRef]
  190. Zhang, Y.; Liao, J.; Fang, X.; Bai, F.D.; Qiao, K.; Wang, L.M. Renewable High Performance Polyurethane Bioplastics Derived from Lignin-Poly(Ε-Caprolactone). ACS Sustain. Chem. Eng. 2017, 5, 4276–4284. [Google Scholar] [CrossRef]
  191. Laurichesse, S.; Avérous, L. Chemical Modification of Lignins: Towards Biobased Polymers. Prog. Polym. Sci. 2014, 39, 1266–1290. [Google Scholar] [CrossRef]
  192. Lora, J.H.; Glasser, W.G. Recent Industrial Applications of Lignin: A Sustainable Alternative to Nonrenewable Materials. J. Polym. Environ. 2002, 10, 39–48. [Google Scholar] [CrossRef]
  193. Sadeghifar, H.; Cui, C.; Argyropoulos, D.S. Toward Thermoplastic Lignin Polymers. Part I. Selective Masking of Phenolic Hydroxyl Groups in Kraft Lignins via Methylation and Oxypropylation Chemistries. Ind. Eng. Chem. Res. 2012, 51, 16713–16720. [Google Scholar] [CrossRef]
  194. Zhang, Y.M.; Yan, R.; Ngo, T.D.; Zhao, Q.; Duan, J.C.; Du, X.W.; Wang, Y.L.; Liu, B.J.; Sun, Z.Y.; Hu, W. Ozone Oxidized Lignin-Based Polyurethane with Improved Properties. Eur. Polym. J. 2019, 117, 114–122. [Google Scholar] [CrossRef]
  195. Wang, Y.Y.; Scheidemantle, B.; Ragauskas, A.J. Polyurethanes Based on Unmodified and Refined Technical Lignins: Correlation between Molecular Structure and Material Properties. Biomacromolecules 2021, 22, 2129–2136. [Google Scholar] [CrossRef]
  196. Hu, L.S.; Guang, C.Y.; Liu, Y.; Su, Z.Q.; Gong, S.D.; Yao, Y.J.; Wang, Y.P. Adsorption Behavior of Dyes from An Aqueous Solution onto Composite Magnetic Lignin Adsorbent. Chemosphere 2020, 246, 125757. [Google Scholar] [CrossRef]
  197. Zhang, Y.; Li, T.T.; Lou, C.W.; Lin, J.H. Facile Method for Tent Fabrics with Eco-Friendly/Durable Properties Using Waterborne Polyurethane/Lignin: Preparation and Evaluation. J. Ind. Text. 2022, 51, 4149S–4166S. [Google Scholar] [CrossRef]
  198. Oveissi, F.; Naficy, S.; Le, T.Y.L.; Fletcher, D.F.; Dehghani, F. Tough and Processable Hydrogels Based on Lignin and Hydrophilic Polyurethane. ACS Appl. Bio Mater. 2018, 1, 2073–2081. [Google Scholar] [CrossRef] [PubMed]
  199. Castro, E.D.; Von Pinho, E.V.D.; de Lima, A.E.; Labory, C.R.G.; dos Santos, H.O.; Alves, E.; Guerra, A.F.D. Physiological Quality, Lignin and the Ultrastructural Characterization of Soybean Seeds. Acta Sci. Agron. 2024, 46, e63621. [Google Scholar] [CrossRef]
  200. Zhang, J.M.; Hori, N.; Takemura, A. Thermal and Time Regularities during Oilseed Rape Straw Liquefaction Process to Produce Bio-Polyol. J. Clean. Prod. 2020, 277, 124015. [Google Scholar] [CrossRef]
  201. Chen, F.G.; Lu, Z.M. Liquefaction of Wheat Straw and Preparation of Rigid Polyurethane Foam from the Liquefaction Products. J. Appl. Polym. Sci. 2010, 111, 508–516. [Google Scholar] [CrossRef]
  202. Wang, T.P.; Zhang, L.H.; Li, D.; Yin, J.; Wu, S.; Mao, Z.H. Mechanical Properties of Polyurethane Foams Prepared from Liquefied Corn Stover with PAPI. Bioresour. Technol. 2008, 99, 2265–2268. [Google Scholar] [CrossRef] [PubMed]
  203. D’Souza, J.; Yan, N. Producing Bark-based Polyols through Liquefaction: Effect of Liquefaction Temperature. ACS Sustain. Chem. Eng. 2013, 1, 534–540. [Google Scholar] [CrossRef]
  204. Ye, L.Y.; Zhang, J.M.; Zhao, J.; Tu, S. Liquefaction of Bamboo Shoot Shell for the Production of Polyols. Bioresour. Technol. 2014, 153, 147–153. [Google Scholar] [CrossRef]
  205. Hu, S.J.; Wan, C.X.; Li, Y.B. Production and Characterization of Biopolyols and Polyurethane Foams from Crude Glycerol Based Liquefaction of Soybean Straw. Bioresour. Technol. 2012, 103, 227–233. [Google Scholar] [CrossRef]
  206. Serrano, L.; Rincon, E.; Garcia, A.; Rodriguez, J.; Briones, R. Bio-Degradable Polyurethane Foams Produced by Liquefied Polyol from Wheat Straw Biomass. Polymer 2020, 12, 2646. [Google Scholar] [CrossRef]
  207. Li, Y.Y.; Jing, W.W.; Wang, J.H.; Li, J.F. Elucidating the Relationship Between Structure and Property of Waterborne Polyurethane-Cellulose Nanocrystals Nanocomposite Films. Sci. Adv. Mater. 2020, 12, 1213–1224. [Google Scholar] [CrossRef]
  208. Mahdieh, A.; Motasadizadeh, H.; Yeganeh, H.; Nyström, B.; Dinarvand, R. Redox-Responsive Waterborne Polyurethane Nanocarriers for Targeted Doxorubicin Delivery. Int. J. Pharm. 2022, 628, 122275. [Google Scholar] [CrossRef] [PubMed]
  209. Cai, J.; Murugadoss, V.; Jiang, J.; Gao, X.; Lin, Z.; Huang, M.; Jiang, G.; Alsareii, S.A.; Algadi, H.; Kathiresan, M. Waterborne Polyurethane and Its Nanocomposites: A Mini-Review for Anti-Corrosion Coating, Flame Retardancy, and Biomedical Applications. Adv. Compos. Hybrid Mater. 2022, 5, 641–650. [Google Scholar] [CrossRef]
  210. Wang, Y.X.; Zhang, L.N. High-Strength Waterborne Polyurethane Reinforced with Waxy Maize Starch Nanocrystals. J. Nanosci. Nanotechnol. 2008, 8, 5831–5838. [Google Scholar] [CrossRef] [PubMed]
  211. Kashcheyeva, E.I.; Gismatulina, Y.A.; Mironova, G.F.; Gladysheva, E.K.; Budaeva, V.V.; Skiba, E.A.; Zolotuhin, V.N.; Shavyrkina, N.A.; Kortusov, A.N.; Korchagina, A.A. Properties and Hydrolysis Behavior of Celluloses of Different Origin. Polymer 2022, 14, 3899. [Google Scholar] [CrossRef] [PubMed]
  212. Reis, R.S.; Souza, D.D.H.S.; Marques, M.D.F.V.; da Luz, F.S.; Monteiro, S.N. Novel Bionanocomposite of Polycaprolactone Reinforced with Steam-Exploded Microfibrillated Cellulose Modified with ZnO. J. Mater. Res. Technol. 2021, 13, 1324–1335. [Google Scholar] [CrossRef]
  213. Stanzione, M.; Oliviero, M.; Cocca, M.; Errico, M.E.; Gentile, G.; Avella, M.; Lavorgna, M.; Buonocore, G.G.; Verdolotti, L. Tuning of polyurethane foam mechanical and thermal properties using ball-milled cellulose. Carbohydr. Polym. 2020, 231, 115772. [Google Scholar] [CrossRef]
  214. Klemm, D.; Kramer, F.; Moritz, S.; Lindstrom, T.; Ankerfors, M.; Gray, D.; Dorris, A. Nanocelluloses: A New Family of Nature-Based Materials. Angew. Chem. Int. Ed. 2011, 50, 5438–5466. [Google Scholar] [CrossRef]
  215. Lee, K.Y.; Aitomaki, Y.; Berglund, L.A.; Oksman, K.; Bismarck, A. On the Use of Nanocellulose as Reinforcement in Polymer Matrix Composites. Compos. Sci. Technol. 2014, 105, 15–27. [Google Scholar] [CrossRef]
  216. Azeredo, H.M.C.; Rosa, M.F.; Mattoso, L.H.C. Nanocellulose in Bio-Based Food Packaging Applications. Ind. Crops Prod. 2017, 97, 664–671. [Google Scholar] [CrossRef]
  217. Jorfi, M.; Foster, E.J. Recent Advances in Nanocellulose for Biomedical Applications. J. Appl. Polym. Sci. 2015, 132, 41719. [Google Scholar] [CrossRef]
  218. Chen, R.D.; Huang, C.F.; Hsu, S.H. Composites of Waterborne Polyurethane and Cellulose Nanofibers for 3D Printing and Bioapplications. Carbohydr. Polym. 2019, 212, 75–88. [Google Scholar] [CrossRef] [PubMed]
  219. Dutta, G.K.; Karak, N. Waste Brewed Tea Leaf Derived Cellulose Nanofiber Reinforced Fully Bio-Based Waterborne Polyester Nanocomposite as An Environmentally Benign Material. RSC Adv. 2019, 9, 20829–20840. [Google Scholar] [CrossRef] [PubMed]
  220. Choi, S.M.; Lee, M.W.; Shin, E.J. One-Pot Processing of Regenerated Cellulose Nanoparticles/Waterborne Polyurethane Nanocomposite for Eco-Friendly Polyurethane Matrix. Polymers 2019, 11, 356. [Google Scholar] [CrossRef] [PubMed]
  221. Kong, L.L.; Xu, D.D.; He, Z.X.; Wang, F.Q.; Gui, S.H.; Fan, J.L.; Pan, X.Y.; Dai, X.H.; Dong, X.Y.; Liu, B.X.; et al. Nanocellulose-Reinforced Polyurethane for Waterborne Wood Coating. Molecules 2019, 24, 3151. [Google Scholar] [CrossRef] [PubMed]
  222. Zhou, X.; Zhang, X.; Wang, D.; Fang, C.Q.; Lei, W.Q.; Huang, Z.G.; Song, Y.H.; He, X.Y.; Huang, Y.W. Preparation and Characterization of Waterborne Polyurethane/Cellulose Nanocrystal Composite Membrane from Recycling Waste Paper. J. Renew. Mater. 2020, 8, 631–645. [Google Scholar] [CrossRef]
  223. Lei, W.Q.; Zhou, X.; Fang, C.Q.; Song, Y.H.; Li, Y.G. Eco-Friendly Waterborne Polyurethane Reinforced with Cellulose Nanocrystal from Office Waste Paper by Two Different Methods. Carbohydr. Polym. 2019, 209, 299–309. [Google Scholar] [CrossRef] [PubMed]
  224. Zhang, S.D.; Sun, K.; Liu, H.; Chen, X.Y.; Zheng, Y.J.; Shi, X.Z.; Zhang, D.B.; Mi, L.W.; Liu, C.T.; Shen, C.Y. Enhanced Piezoresistive Performance of Conductive WPU/CNT Composite Foam Through Incorporating Brittle Cellulose Nanocrystal. Chem. Eng. J. 2020, 387, 124045. [Google Scholar] [CrossRef]
  225. Zhang, P.B.; Lu, Y.D.; Fan, M.M.; Jiang, P.P.; Dong, Y.M. Modified Cellulose Nanocrystals Enhancement to Mechanical Properties and Water Resistance of Vegetable Oil-Based Waterborne Polyurethane. J. Appl. Polym. Sci. 2019, 136, 48228. [Google Scholar] [CrossRef]
  226. Cheng, L.S.; Ren, S.B.; Lu, X.N. Application of Eco-Friendlywaterborne Polyurethane Composite Coating Incorporated with Nano Cellulose Crystalline and Silver Nano Particles on Wood Antibacterial Board. Polymers 2020, 12, 407. [Google Scholar] [CrossRef]
  227. Hormaiztegui, M.E.V.; Daga, B.; Aranguren, M.I.; Mucci, V. Bio-Based Waterborne Polyurethanes Reinforced with Cellulose Nanocrystals as Coating Films. Prog. Org. Coat. 2020, 144, 105649. [Google Scholar] [CrossRef]
  228. Hu, W.; Chen, S.; Yang, J.; Li, Z.; Wang, H. Functionalized Bacterial Cellulose Derivatives and Nanocomposites. Carbohydr. Polym. 2014, 101, 1043–1060. [Google Scholar] [CrossRef]
  229. Urbina, L.; Alonso-Varona, A.; Saralegi, A.; Palomares, T.; Eceiza, A.; Corcuera, M.A.; Retegi, A. Hybrid and Biocompatible Cellulose/Polyurethane Nanocomposites with Water-Activated Shape Memory Properties. Carbohydr. Polym. 2019, 216, 86–96. [Google Scholar] [CrossRef]
  230. Zia, F.; Zia, K.M.; Zuber, M.; Kamal, S.; Aslam, N. Starch Based Polyurethanes: A Critical Review. Carbohydr. Polym. 2015, 134, 784–798. [Google Scholar] [CrossRef]
  231. Tai, N.L.; Ghasemlou, M.; Adhikari, R.; Adhikari, B. Starch-based isocyanate- and non-isocyanate polyurethane hybrids: A review on synthesis, performance and biodegradation. Carbohydr. Polym. 2021, 265, 118029. [Google Scholar] [CrossRef]
  232. Chemelli, A.; Gomernik, F.; Thaler, F.; Huber, A.; Hirn, U.; Bauer, W.; Spirk, S. Cationic Starches in Paper-Based Applications—A Review on Analytical Methods. Carbohydr. Polym. 2020, 235, 115964. [Google Scholar] [CrossRef] [PubMed]
  233. Azeredo, H.M.C.D. Nanocomposites for Food Packaging Applications. Food Res. Int. 2009, 42, 1240–1253. [Google Scholar] [CrossRef]
  234. Tsupphayakorn-aek, P.; Suwan, A.; Chaisit, T.; Tulyapitak, T.; Phinyocheep, P.; Pilard, J.F.; Saetung, N.; Saetung, A. A new UV-curable biodegradable waterborne polyurethane-acrylate based on natural rubber blended with cassava starch. J. Appl. Polym. 2023, 140, e53787. [Google Scholar] [CrossRef]
  235. Lee, S.J.; Kim, B.K. Covalent Incorporation of Starch Derivative into Waterborne Polyurethane for Biodegradability. Carbohydr. Polym. 2012, 87, 1803–1809. [Google Scholar] [CrossRef]
  236. Tai, N.L.; Adhikari, R.; Shanks, R.; Halley, P.; Adhikari, B. Flexible Starch-Polyurethane Films: Effect of Mixed Macrodiol Polyurethane Ionomers on Physicochemical Characteristics and Hydrophobicity. Carbohydr. Polym. 2018, 197, 312–325. [Google Scholar] [CrossRef] [PubMed]
  237. Chang, P.R.; Ai, F.J.; Chen, Y.; Dufresne, A.; Huang, J. Effects of Starch Nanocrystal-Graft-Polycaprolactone on Mechanical Properties of Waterborne Polyurethane-Based Nanocomposites. J. Appl. Polym. Sci. 2008, 111, 619–627. [Google Scholar] [CrossRef]
  238. Chen, G.J.; Wei, M.; Chen, J.H.; Huang, J.; Dufresne, A.; Chang, P.R. Simultaneous Reinforcing and Toughening: New Nanocomposites of Waterborne Polyurethane Filled with Low Loading Level of Starch Nanocrystals. Polymer 2008, 49, 1860–1870. [Google Scholar] [CrossRef]
  239. Lee, D.S.; Hur, P.; Kim, B.K. Chemical Hybridization of Waterborne Polyurethane with Beta-Cyclodextrin by Sol-Gel Reaction. Prog. Org. Coat. 2017, 111, 107–111. [Google Scholar] [CrossRef]
  240. Konieczny, J.; Loos, K. Bio-Based Polyurethane Films Using White Dextrins. J. Appl. Polym. Sci. 2019, 136, 47454. [Google Scholar] [CrossRef]
  241. Manna, S.; Seth, A.; Gupta, P.; Nandi, G.; Dutta, R.; Jana, S.; Jana, S. Chitosan Derivatives as Carriers for Drug Delivery and Biomedical Applications. ACS Biomater. Sci. Eng. 2023, 9, 2181–2202. [Google Scholar] [CrossRef]
  242. Aoun, R.B.; Trabelsi, N.; Abdallah, M.; Mourtzinos, I.; Mhamdi, R. Towards a Greener Future: Exploring the Challenges of Extraction of Chitin and Chitosan as Bioactive Polysaccharides. Mater. Today Commun. 2024, 39, 108761. [Google Scholar] [CrossRef]
  243. Shaabani, A.; Sedghi, R. Preparation of Chitosan Biguanidine/PANI-Containing Self-Healing Semi-Conductive Waterborne Scaffolds for Bone Tissue Engineering. Carbohydr. Polym. 2021, 264, 118045. [Google Scholar] [CrossRef] [PubMed]
  244. Bankoti, K.; Rameshbabu, A.P.; Datta, S.; Maity, P.P.; Goswami, P.; Datta, P.; Ghosh, S.K.; Mitra, A.; Dhara, S. Accelerated Healing of Full Thickness Dermal Wounds by Macroporous Waterborne Polyurethane-Chitosan Hydrogel Scaffolds. Mater. Sci. Eng. C 2017, 81, 133–143. [Google Scholar] [CrossRef]
  245. Lin, L.; Tu, Y.; Li, Z.; Wu, H.; Mao, H.; Wang, C. Synthesis and Application of Multifunctional Lignin-Modified Cationic Waterborne Polyurethane in Textiles. Int. J. Biol. Macromol. 2024, 262, 130063. [Google Scholar] [CrossRef]
  246. Liu, Y.X.; Zou, Y.L.; Wang, J.; Wang, S.; Liu, X.F. A Novel Cationic Waterborne Polyurethane Coating Modified by Chitosan Biguanide Hydrochloride with Application Potential in Medical Catheters. J. Appl. Polym. Sci. 2020, 138, e50290. [Google Scholar] [CrossRef]
  247. Lin, T.W.; Hsu, S.H. Self-Healing Hydrogels and Cryogels from Biodegradable Polyurethane Nanoparticle Crosslinked Chitosan. Adv. Sci. 2020, 7, 1901388. [Google Scholar] [CrossRef]
  248. Gupta, A.; Kim, B.S. Shape Memory Polyurethane Biocomposites Based on Toughened Polycaprolactone Promoted by Nano-Chitosan. Nanomaterials 2019, 9, 225. [Google Scholar] [CrossRef]
  249. Kabir, I.I.; Sorrell, C.C.; Mofarah, S.S.; Yang, W.; Yuen, A.C.Y.; Nazir, M.T.; Yeoh, G.H. Alginate/Polymer-Based Materials for Fire Retardancy: Synthesis, Structure, Properties and Applications. Polym. Rev. 2021, 61, 357–414. [Google Scholar] [CrossRef]
  250. Wang, X.; Zhang, Y.; Liang, H.; Zhou, X.; Fang, C.; Zhang, C.; Luo, Y. Synthesis and Properties of Castor Oil-Based Waterborne Polyurethane/Sodium Alginate Composites with Tunable Properties. Carbohydr. Polym. 2019, 208, 391–397. [Google Scholar] [CrossRef]
  251. Tan, S.H.; Chua, D.A.C.; Tang, J.R.J.; Bonnard, C.; Leavesley, D.; Liang, K. Design of Hydrogel-Based Scaffolds for in Vitro Three-Dimensional Human Skin Model Reconstruction. Acta Biomater. 2022, 153, 13–37. [Google Scholar] [CrossRef]
  252. Zia, F.; Zia, K.M.; Zuber, M.; Ahmad, H.B.; Muneer, M. Glucomannan Based Polyurethanes: A Critical Short Review of Recent Advances and Future Perspectives. Int. J. Biol. Macromol. 2016, 87, 229–236. [Google Scholar] [CrossRef]
  253. Alonso-Sande, M.; Teijeiro-Osorio, D.; Remunan-Lopez, C.; Alonso, M.J. Glucomannan, a Promising Polysaccharide for Biopharmaceutical Purposes. Eur. J. Pharm. Biopharm. 2009, 72, 453–462. [Google Scholar] [CrossRef] [PubMed]
  254. Yang, G.; Huang, Q.; Zhang, L.; Zhou, J.; Gao, S. Miscibility and Properties of Blend Materials from Waterborne Polyurethane and Carboxymethyl Konjac Glucomannan. J. Appl. Polym. Sci. 2010, 92, 77–83. [Google Scholar] [CrossRef]
  255. Zhang, S.; Xiang, A.M.; Tian, H.F.; Rajulu, A.V. Water-Blown Castor Oil-Based Polyurethane Foams with Soy Protein as a Reactive Reinforcing Filler. J. Polym. Environ. 2018, 26, 15–22. [Google Scholar] [CrossRef]
  256. Madbouly, S.A.; Lendlein, A. Degradable Polyurethane/Soy Protein Shape-Memory Polymer Blends Prepared Via Environmentally-Friendly Aqueous Dispersions. Macromol. Mater. Eng. 2012, 297, 1213–1224. [Google Scholar] [CrossRef]
  257. Chan-Chan, L.H.; Gonzalez-Garcia, G.; Vargas-Coronado, R.F.; Cervantes-Uc, J.M.; Hernandez-Sanchez, F.; Marcos-Femandez, A.; Cauich-Rodriguez, J.V. Characterization of model compounds and poly(amide-urea) urethanes based on amino acids by FTIR, NMR and other analytical techniques. Eur. Polym. J. 2017, 92, 27–39. [Google Scholar] [CrossRef]
  258. Liang, C.; Gracida-Alvarez, U.R.; Gallant, E.T.; Gillis, P.A.; Marques, Y.A.; Abramo, G.P.; Hawkins, T.R.; Dunn, J.B. Material Flows of Polyurethane in the United States. Environ. Sci. Technol. 2021, 55, 14215–14224. [Google Scholar] [CrossRef] [PubMed]
  259. Pham, C.T.; Nguyen, B.T.; Nguyen, H.T.T.; Kang, S.J.; Kim, J.; Lee, P.C.; Hoang, D.Q. Comprehensive Investigation of the Behavior of Polyurethane Foams Based on Conventional Polyol and Oligo-Ester-Ether-Diol from Waste Poly(ethylene terephthalate): Fireproof Performances, Thermal Stabilities, and Physicomechanical Properties. ACS Omega 2021, 5, 33053–33063. [Google Scholar] [CrossRef] [PubMed]
  260. Molero, C.; De Lucas, A.; Romero, F.; Rodríguez, J.F. Influence of the Use of Recycled Polyols Obtained by Glycolysis on the Preparation and Physical Properties of Flexible Polyurethane. J. Appl. Polym. Sci. 2008, 109, 617–626. [Google Scholar] [CrossRef]
  261. Tan, S.Q.; Abraham, T.; Ference, D.; Macosko, C.W. Rigid Polyurethane Foams from a Soybean Oil-Based Polyol. Polymer 2011, 52, 2840–2846. [Google Scholar] [CrossRef]
  262. Lu, Y.S.; Larock, R.C. Soybean Oil-Based, Aqueous Cationic Polyurethane Dispersions: Synthesis and Properties. Prog. Org. Coat. 2010, 69, 31–37. [Google Scholar] [CrossRef]
  263. Gandhi, T.S.; Patel, M.R.; Dholakiya, B.Z. Synthesis of Cashew Mannich Polyol Via a Three Step Continuous Route and Development of PU Rigid Foams with Mechanical, Thermal and Fire Studies. J. Polym. Eng. 2015, 35, 533–544. [Google Scholar] [CrossRef]
  264. Zhang, C.; Bhoyate, S.; Ionescu, M.; Kahol, P.K.; Gupta, R.K. Highly Flame Retardant and Bio-Based Rigid Polyurethane Foams Derived from Orange Peel Oil. Polym. Eng. Sci. 2018, 58, 2078–2087. [Google Scholar] [CrossRef]
Figure 1. The chemical structure of PU.
Figure 1. The chemical structure of PU.
Polymers 16 01889 g001
Figure 3. Hydrolysis of PET to TPA. (A) Alkaline solid-state PET hydrolysis into TPA by vapor-assisted aging under different relative humidity levels and solvent vapors. Reproduced with permission [62]. Copyright 2020, Wiley. (B) The overall plan of the upgrading and utilization of a waste PET biological refinery. Reproduced with permission [46]. Copyright 2019, American Chemical Society.
Figure 3. Hydrolysis of PET to TPA. (A) Alkaline solid-state PET hydrolysis into TPA by vapor-assisted aging under different relative humidity levels and solvent vapors. Reproduced with permission [62]. Copyright 2020, Wiley. (B) The overall plan of the upgrading and utilization of a waste PET biological refinery. Reproduced with permission [46]. Copyright 2019, American Chemical Society.
Polymers 16 01889 g003
Figure 4. Alcoholysis of PET produces DMT. (A) The main depolymerization reaction of PET in supercritical methanol to DMT. Reproduced with permission [64]. Copyright 2015, Elsevier. (B) 1H spectra recorded by AP-PET and CP-PET in deuterated CDCl3/HFIP; The right figure shows PET degradation products were ethylene glycol and terephthalic acid, which indicated that the outer chain fracture occurred. Reproduced with permission [50]. Copyright 2019, Wiley.
Figure 4. Alcoholysis of PET produces DMT. (A) The main depolymerization reaction of PET in supercritical methanol to DMT. Reproduced with permission [64]. Copyright 2015, Elsevier. (B) 1H spectra recorded by AP-PET and CP-PET in deuterated CDCl3/HFIP; The right figure shows PET degradation products were ethylene glycol and terephthalic acid, which indicated that the outer chain fracture occurred. Reproduced with permission [50]. Copyright 2019, Wiley.
Polymers 16 01889 g004
Figure 5. The reaction scheme of PC waste and transesterified oils (CO-TMP, RO-TMP and CCO-TMP) for recycled polyols. The inset shows the acyl structure of the TMP monoester present in transesterification oils. Reproduced with permission [122]. Copyright 2018, Tech Science Press.
Figure 5. The reaction scheme of PC waste and transesterified oils (CO-TMP, RO-TMP and CCO-TMP) for recycled polyols. The inset shows the acyl structure of the TMP monoester present in transesterification oils. Reproduced with permission [122]. Copyright 2018, Tech Science Press.
Polymers 16 01889 g005
Table 1. Comparison of different recycling methods of waste PET.
Table 1. Comparison of different recycling methods of waste PET.
MethodsDegradation ReagentsProductsPotential ApplicationsRef.
GlycolysisDiethylene glycol Crude glycero
Pentaerytheritol Trimethyloylpropane
Ethylene glycol
Neopentyl glycol
Supercritical ethanol-ionic liquid
Polyol
BHET
DET
Oligomer
PU foams
for oil sorption
PU adhesive
Polyester-polyol
PU coatings
[42,43,44,45,51,52,53,54]
HydrolysisDeionized water
NaOH
TPA, EGAromatics Aromatic-derived compounds[46]
AlcoholysisIsooctyl alcohol
Methanol
DMT
EG
Polyvinyl chloride[47,48]
BiodegradationSakaiensis bacterium
Hydrolases
MHET, BHET
TPA, EG
PU foams[49,50]
Table 2. Comparison of different chemical recycling methods of waste PU.
Table 2. Comparison of different chemical recycling methods of waste PU.
MethodsDegradation ReagentsProductsPotential ApplicationsRefs.
HydrolysisCO2-water
KOH-water
Polyol
Diamines
Synthesis of PU
Translated
into isocyanate
[76,77,78]
GlycolysisDiethylene glycol
Ethylene glycol
Glycerol
Diglycerol
Pentaerythritol
Crude glycerine
Polyol
Aromatic-carbamates
Synthesis of new flexible PU foam
Thermoplastic PU and toluenediamines
[79,80,81,82,87]
AmmonolysisAlkanolamines
Dibutylamine ethanolamine
Polyol
Aromatic amines
Alkanolamine derivatives
Synthesis of new PUs melamine resins, epoxy resins, polyester, polycarbonates[99,100]
PhosphorolysisPhosphonic acids
Phosphate
Phosphoric acid
Ethyl ester
Phosphorus
Chlorine element oligomer
Flame retardant polyurethane or PVC materials[104,105,106]
HydroglycolysisWater and glycolsPolyol
Intermediate chemicals
Synthesis of new PUs[107,108,109]
Table 3. The overview of the main bio-based feedstocks involved in molecular structure and properties of PU products.
Table 3. The overview of the main bio-based feedstocks involved in molecular structure and properties of PU products.
Bio-Based Feedstocks Molecular StructureProductsPropertiesRef.
Vegetable oilPolymers 16 01889 i001PU elastomer
WPU
Mechanical properties
Thermal resistance
Physicomechanical stability
Good adhesion
[128,129,134]
Cashew nut shell liquidPolymers 16 01889 i002PU foamsGood mechanical, thermal and fire properties[110,156]
TerpenePolymers 16 01889 i003PU
elastomer
Good mechanical properties[161,162]
RosinPolymers 16 01889 i004PU elastomerExcellent mechanical rigidity, Heat resistance, shape memory property[155,168,172]
LigninPolymers 16 01889 i005PU foamsAntibacterial, oxidation and ultraviolet resistance
Flame retardant
[183,184,185]
CellulosePolymers 16 01889 i006WPUBiodegradability, mechanical properties, renewability
Thermal properties
[221,229]
StarchPolymers 16 01889 i007PU filmsGood transparency, thermal properties
Mechanical properties
[165,237,238]
ChitosanPolymers 16 01889 i008WPU
PU films
Self-healing properties
Good shape recovery
[197,247]
Sodium alginatePolymers 16 01889 i009WPUHydrophobicity
Mechanical properties
Thermal stability
[250]
GlucomannanPolymers 16 01889 i010WPUOutstanding mechanical properties
Good compatibility
[253,254]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pu, M.; Fang, C.; Zhou, X.; Wang, D.; Lin, Y.; Lei, W.; Li, L. Recent Advances in Environment-Friendly Polyurethanes from Polyols Recovered from the Recycling and Renewable Resources: A Review. Polymers 2024, 16, 1889. https://doi.org/10.3390/polym16131889

AMA Style

Pu M, Fang C, Zhou X, Wang D, Lin Y, Lei W, Li L. Recent Advances in Environment-Friendly Polyurethanes from Polyols Recovered from the Recycling and Renewable Resources: A Review. Polymers. 2024; 16(13):1889. https://doi.org/10.3390/polym16131889

Chicago/Turabian Style

Pu, Mengyuan, Changqing Fang, Xing Zhou, Dong Wang, Yangyang Lin, Wanqing Lei, and Lu Li. 2024. "Recent Advances in Environment-Friendly Polyurethanes from Polyols Recovered from the Recycling and Renewable Resources: A Review" Polymers 16, no. 13: 1889. https://doi.org/10.3390/polym16131889

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