Next Article in Journal
Engineering Characteristics of Cement Composites Containing a Chitosan-Based Polymer and Steel Slag Aggregates
Next Article in Special Issue
High Mechanical Properties of Stretching Oriented Poly(butylene succinate) with Two-Step Chain Extension
Previous Article in Journal
Rheology and Tack Properties of Biodegradable Isodimorphic Poly(butylene succinate)-Ran-Poly(ε-caprolactone) Random Copolyesters and Their Potential Use as Adhesives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress on Bio-Based Polyesters Derived from 2,5-Furandicarbonxylic Acid (FDCA)

1
Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, 1219 Zhongguan West Road, Zhenhai District, Ningbo 315201, China
2
Key Laboratory of Bio-Based Polymeric Materials Technology and Application of Zhejiang Province, 1219 Zhongguan West Road, Zhenhai District, Ningbo 315201, China
3
University of Chinese Academy of Sciences, No.19 A, Yuquan Road, Shijingshan District, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
Polymers 2022, 14(3), 625; https://doi.org/10.3390/polym14030625
Submission received: 19 January 2022 / Revised: 2 February 2022 / Accepted: 5 February 2022 / Published: 6 February 2022
(This article belongs to the Special Issue Development of High-Performance Biobased Polyesters)

Abstract

:
The big challenge today is the upgrading of sustainable materials to replace miscellaneous ones from petroleum resources. Thus, a generic bio-based building block lays the foundation of the huge bio-market to green economy. 2,5-Furandicarboxylic acid (FDCA), a rigid diacid derived from lignocellulose or fructose, represents a great potential as a contender to terephthalic acid (TPA). Recently, studies on the synthesis, modification, and functionalization of bio-based polyesters based on FDCA have attracted widespread attention. To apply furanic polyesters on engineering plastics, packaging materials, electronics, etc., researchers have extended the properties of basic FDCA-based homo-polyesters by directional copolymerization and composite preparation. This review covers the synthesis and performance of polyesters and composites based on FDCA with emphasis bedded on the thermomechanical, crystallization, barrier properties, and biodegradability. Finally, a summary of what has been achieved and the issues waiting to be addressed of FDCA-based polyester materials are suggested.

Graphical Abstract

1. Introduction

As concern regarding the pollution of petrochemicals spreads abroad, sustainability has been highlighted in the last decades: bio-based plastics from renewable resources have become a solution, trending within polymer research [1,2]. It is estimated that a total amount of 8000 Mt plastics has been produced worldwide, 79% of which has been discarded. Furthermore, the global production of plastics is rising stably, reaching 28,000 Mt in 2050 [3]. In this vein, the significance of recycling is highlighted as the European Parliament calls for a circular economy where waste prevention and recycling are encouraged. Currently, widely-applied bio-based polymers such as poly(lactic acid) (PLA) [4], polyhydroxyalkanoates (PHAs) [5], poly(glycolic acid) (PGA) [6], and polybutylene succinate (PBS) [7], which all belong to aliphatic polymers, lack rigidity from their innate structure, leading to weaker thermomechanical properties compared to oil-based poly(ethylene terephthalate) (PET), poly(butylene terephthalate) (PBT), polycarbonates (PC), etc. To move forward, bio-based polymer materials demand rigidity through polymerization or blending. Back in 2004, the U.S. Department of Energy presented “twelve platform bio-based chemicals”, among which FDCA was the only aromatic compound that was later employed heavily in the synthesis of polyester [8,9,10,11], polyamide [12,13,14,15], epoxy resin [16,17], and other polymers [18,19,20]. The large-scale investigations of polyesters have proven them to be promising bio-based building blocks that are expected to substitute TPA (Figure 1) [21,22,23]. Major corporations in the chemical industries such as DuPont and BASF have engaged in the mass production of FDCA [24]. In addition, the FDCA-based polyester of most interest, poly(ethylene 2,5-furancoate) (PEF), has been commercialized by Avantium with the YXY® plant-to-plastics process [25].
FDCA has many reasons to explain its growing influence in the realm of bio-polymers. The primary one is the sustainable origin of FDCA. The main route is the oxidation of 5-hydroxymethylfurfural (HMF), which is produced from lignocellulose or fructose [26]. The fast development of renewable biomass undoubtedly promotes the downstream industry as Avantium, ADM, Shell, and BASF have engaged in the pilot plant production of HMF. One life cycle assessment showed that the non-renewable energy and the greenhouse gas emissions of PEF production are cut down by about 50% compared to that of PET [22]. Besides the green origin of FDCA, the foremost advantage is its structural similarity to TPA. The rigidity of its five-membered ring enhances the thermomechanical properties of polymers that can distinguish itself from other long-chain bio-monomers. Although the common features of rigidity from the ring structure are shared between FDCA and TPA, the divergences in the ring size, polarity, and geometry make differences in the macro performance of polyesters based on them. More details can be demonstrated as the interatomic distance between carboxylic acid groups in FDCA is 4.830 Å while in TPA, it is 5.731 Å [27,28]. The carbon atoms of the carboxyl units are non-linear in FDCA with an angle of 129.4° while TPA showed a linear structure. The structure, along with the polarity endowed by the oxygen atom, determines that the furan ring is more difficult to flip than TPA. These divergences produce a “butterfly effect” that greatly influences the macro-performances of furanic polyesters, resulting in the huge gap in thermomechanical, barrier, and anti-UV properties compared to TPA-based analogues.
Several reviews on polyesters based on renewable resources have been published, however, the discussion dedicated to furan-based polyesters with counterintuitive and interesting properties still seems helpful. In this paper, the synthesis and basic properties of homo-polyesters, co-polyesters, and composites based on FDCA were collected and discussed. The industrial construction was also involved. At the same time, the authors attempt to clarify the basal conditions and probe the next step of the investigation into furan-polyesters.

2. Synthetic Routes of Furanic Polyesters

Three ways are widely accepted as polyester synthetic routes: melt polymerization, solution esterification, and ring-opening polymerization (ROP) (Scheme 1). Melt polymerization can be further split across two methods based on the starting materials: FDCA with direct polymerization or dimethyl furan-2,5-dicarboxylate (DMFD) from transesterification. Industrial production favors direct esterification to avoid an additional esterification step while transesterification is preferred in lab synthesis to produce polyesters with novel structures. In the early exploration of furanic polyester synthesis, solution polymerization was tried, but suffered from multiple steps of isolation and purification, low yield, and low molecular weight, etc. However, considering that sometimes the melting point of the product is too high, solution polymerization is an option. ROP have been reported more recently to obtain polyesters from cyclic esters or anhydrides [29,30]. Additionally, ROP is the mainstream in the industrial production of PLA. The first report of furanic polyester synthesis can be dated back to 1946 when the Celanese Corporation of America requested a patent application for FDCA transesterification synthesis of PEF [31]. Next, with many explorations and dedication into FDCA-based polyester synthesis, the investigations gathered momentum by around 2010 (Figure 2).

3. Homo-Polyesters from FDCA

3.1. PEF

As a top candidate in the race between bio-based and petroleum-based polyesters, PEF is definitely worth specific discussion. In 2009, Gandini et al. [9] reported a pioneer work where a variety of methods were involved in PEF synthesis. Following this, more reports of PEF properties demonstrated the potentiality and possibility of PEF: its good thermomechanical behaviors and excellent barrier property. In 2011, Gomes et al. [10] prepared PEF via transesterification. The glass transition temperature (Tg) was measured as 80 °C, higher than that of PET (75 °C). The melting temperature (Tm) was given as 215 °C, much lower than that of PET (260 °C). The authors suggested that PEF may replace PET based on the satisfied thermal performances. Zhou’s group [11] investigated the mechanical properties of PEF and found a tensile modulus of 2100 MPa and tensile strength of 66.7 MPa, which was close to that of PET (2000 MPa and 45 MPa). These early investigations of the thermomechanical properties of PEF greatly inspired more in-depth studies. PEF synthesized by ROP was proved feasible, and a weight–average molecular weight (Mw) was reached of 50,000 g/mol. The thermal properties were similar to those obtained via melt polymerization [32]. Knoop et al. [33] prepared PEF via transesterification and solid-state post-condensation (SSPC) to obtain high molecular weight PEF (number-average molecular weight: 83,000 g/mol). The tensile modulus and strength were determined as 2450 MPa and 35 MPa, respectively. In the SSPC process, the solid polyester phase was dispersed in a gaseous phase or liquid phase to continue the transesterification to avoid the unwanted thermal oxidative degradation caused by the long reaction time of traditional melt condensation. The temperature and reaction time were the crucial factors in SSPC: on one hand, the kinetics of chain end diffusion would be slowed down so the long time is required to obtain the designed molecular weight, which would be inefficient for industrial production. On the other hand, the high temperature demands the high crystallinity of polyesters to stay solid, which further shrinks the amorphous proportion, thus limiting the mobility of the reacting groups. In a word, the SSPC process exhibits high complexity so that the conclusions of related reports is sometimes controversial [34]. Banella et al. pursued targeted synthesis of PEF suitable for food-contact-grade packaging material by applying zinc acetate and aluminum acetylacetonate as catalysts. The results showed that both catalysts were operative in the PEF preparation, resulting in film production with high transparency and low discoloration (Figure 3i) [35]. Zhou et al. [36] synthesized white PEF with a fast crystallization rate using metal zinc as the catalyst via transesterification (Figure 3ii). They found that the new catalyst inhibited the discoloration and enhanced the crystallization rate. Besides the basic thermal and mechanical properties, Koros et al. [37,38] explored the barrier properties of PEF and outlined a variety of important conclusions: compared to PET, PEF had an 11 and 19 times higher O2 and CO2 barrier and the reason behind this was ascribed to the difficulty of furan-ring flipping due to the furan ring polarity and the structural nonlinearity [39]. Liu et al. [40] tried to explain the surprisingly good barrier properties of furan-based polyesters from the role of the furan ring’s polarity. Results showed that although PEF had a higher free volume than poly(ethylene-2,5-thiophenedicarboxylate), it still exhibited a better barrier property, which was attributed to the higher dipole moment of the furan ring. The basic thermal and mechanical properties of PEF are collected in Table 1.
The industrial production of PEF is guided by Avantium, who built a pilot plant in the Netherlands to produce PEF via its famous YXY® plant-to-plastics technology. They plan to open its flagship factory in 2023 with the expected capacity of five kilotons per year [43]. The PEF commercialization in Asia is under the plan of Mitsui (Japan) who had a partnership with Avantium. Recent news of ETH Zürich offering another option in the large-scale production of PEF besides melt polymerization: they reported the rapid synthesis of bottle-grade PEF from cyclic oligomers. Sulzer is working with them to adjust the process to adapt for an industrial application [44,45].

3.2. Furan-Polyesters with Aliphatic Diols

Aliphatic long-chain diols (C2–C18) employed in the synthesis of other poly(alkylene furanoates) have been widely reported [10,46,47,48]. Zhou et al. [11] successfully synthesized a series of homo-polyesters from C3 to C8 via direct polymerization, namely, poly(1,3-propylene 2,5-furandicarboxylate) (PPF), poly(1,4-butylene 2,5-furandicarboxylate) (PBF), poly(1,6-hexylene 2,5-furandicarboxylate) (PHF), and poly(octylene 2,5-furandicarboxylate) (POF). The Tg and Tm were measured as 21.8–89.9 °C and 148.2–210.4 °C while the tensile modulus and elongation at break ranged from 340 to 2070 MPa and from 4.2 to 210%. Later, Bikiaris et al. [42] further synthesized polyesters with longer diols such as C8, C9, C10, and C12. The characterization of these more ductile polyesters showed a Tg under 0 °C and inferior mechanical properties. When the length of the aliphatic chain increased, the Tg of these polyesters continuously declined, as listed in Table 1. Besides the synthesis methods above-mentioned, PAF preparation via enzyme polymerization was also reported [49]. However, most number-average molecular weights (Mn) of these polyesters were low (200–10,100 g/mol) and characterization of properties was absent.
In fact, PBT and PPT are famous engineering plastics, so studies on their furan-based homologs, PBF, and PPF have been broadly reported. Zhu et al. [21] investigated the thermal and mechanical properties of PBF and compared them to that of PBT, concluding that the good thermal stability and mechanical strength of PBF made it a possible alternative to PBT. When the Mw of PBF reached 65,000 g/mol, the Tg and Tm were 40 °C and 173 °C, respectively. Meanwhile, the tensile modulus was around 1000 MPa. Via wide angle X-ray scattering, the structure of PBF (Figure 4a) was found to be close to the α and β forms of PBT. Ma et al. [50] also contended that PBF may be a furan counterpart of PBT after investigating their crystallinity and found that the PBF filament could be drawn directly from the melt. Thiyagarajan et al. [51] investigated the thermal properties of PBF from 2,5-FDCA, 2,4-FDCA, and 3,4-FDCA. Results showed that no significant influence was found in Tg. The synthesis, thermal, and mechanical properties of PPF have also been reported [10,11], which showed that PPF has sufficient mechanical strength and thermal stability such as 1500 MPa for tensile modulus and 68 MPa for tensile strength [11] as well as the degradation temperature of 390 °C [10]. Papageorgiou et al. [41] made a comparison between PPF, poly(propylene terephthalate) (PPT), and poly(propylene 2,6-naphthalate) (PPN) in a thermal analysis and found that the thermal transition temperatures of PPF were similar to that of PPT. Additionally, the polarized light microscopy signaled that the size of the PPF crystal was smaller (Figure 4b) and the crystallization of PPF was slower than that of PPT. Zhao et al. [52] compared the barrier properties among PEF, PPF, and PBF theoretically and experimentally. They found that for PEF and PPF, there was little difference in barrier property, however, as the aliphatic chain increased (PBF), the oxygen permeability showed a significant increase while the CO2 permeability remained unchanged. The collection of the thermal and mechanical properties of these polyesters are listed in Table 1.

3.3. Furan-Polyesters with Rigid Diols

One of the important issues of aliphatic long-chain polyesters is the low Tg, which dramatically reduces its application temperature range. To impart the desired stiffness to furan-polyesters, 1,4:3,6-dianhydrohexitols (DAHs), which have three stereoisomers according to the stereochemistry of the two hydroxyl groups, have been employed in many reports [53,54,55,56]. Among them, isosorbide (ISB) is the most widely studied and has been employed in commercial biopolymers as the obviously rigid structure and green origin. The polyester from FDCA and isosorbide (PDASF) or isoidide (PDAIF) were prepared via solution polycondensation. Their Tg reached 180 °C and 140 °C, respectively [10]. Lopez-Sanchez et al. [57] synthesized fully biomass-derived polyester coatings from FDCA, ISB, succinic acid, 1,3-BDO, and 1,5-PDO. ISB was introduced to increase Tg by approximately 40 °C. Terzopoulou et al. [58] synthesized poly(isosorbide furanoate) (PIsF) from DMFD and ISB via melt polycondensation for a food packaging application. Results showed that the Tg was 157 °C and Tdmax was 421 °C with an intrinsic viscosity of 0.39 dL/g. Chebbi et al. [59] synthesized poly(decamethylene-co-isosorbide 2,5-furandicarboxylate)s (PDIsFs) from DMFD, ISB, and 1,10-decanediol via melt polycondensation with Mn within the range of 11,500–25,400 g/mol. ISB units can obviously enhance the thermal performance (Tg of −1 °C to 21 °C, Td5% of 405 °C to 413 °C) and mechanical properties (tensile modulus ranging from 14 MPa to 559 MPa and elongation at break between 205–266%) of the obtained polyesters. Chen et al. [60] employed ISB to modify PBF. The Tg of copolyesters reached 107 °C with ISB moiety of 50%. In addition, the hydrolytic degradation was promoted by ISB content. Wang et al. [61] synthesized fully bio-based poly(isosorbide-co-butylene 2,5-furandicarboxylate) (PISBF) from DMFD, ISB, and 1,3-BDO. The Tg and Tdmax of PISBF was between 55 °C and 150 °C and 405 °C and 417 °C, respectively. Results showed that BDO toughened the poly(isosorbide 2,5-furandicarboxylate) (PIF) polyester chains where the elongation at break was higher than 46%. Kasmi et al. [62] synthesized poly(1,4-cyclohexanedimethanol-co-isosorbide 2,5-furandicarboxylate)s (PCIsFs) from DMFD, ISB, and CHDM with Tg ranging from 76.8 °C to 103.5 °C. The investigation of crystallization revealed that isosorbide furanoate were inserted in the crystals of PCF homo-polyester. Besides FDCA-based polyesters, ISB has been introduced in numerous polymers. Wu et al. [63] employed isoidide dicarboxylic acid (IIDCA) and diols derived from isohexide building blocks to synthesize novel bio-polyesters. The obtained polyester from IIDCA and ISB (PIsI) showed a Tg of 73 °C and Td5% of 274 °C with Mn of 2600 g/mol while polyester (PImI) from IIDCA and isomannide (IM) showed a Tg of 30 °C and Td5% of 267 °C with Mn of 1200 g/mol. For polyester (PIiI) from IIDMC and isoidide (II), the observed Tg and Td5% were highest at 80 °C and 286 °C with Mn of 2500 g/mol. The authors suggested that IIDCA can enhance the Tg by approximately 70 °C, as introduced into the polymer main chain. Jacquel et al. [64] modified PBS polymers with ISB (PBIS) or FDCA (PBSF) for biodegradable film applications. The thermal properties of the copolyesters obtained were compared to PBST. TPA, FDCA, and ISB all showed positive effects on Tg, among which ISB exhibited a 19 °C increase in Tg compared to pristine PBS (−30 °C). Furthermore, PBSFs were observed as compostable polyesters. Wu et al. [65] prepared poly(isoidide-2,5-bismethylene furan-2,5-dicarboxylate) (PXIIF) from extended isoidide, isoidide-2,5-dimethanol, via melt polycondensation. The Mn of the obtained polyesters was 6900 g/mol and later reached 30,300 g/mol after SSPC. The Tg from the DSC results was 94 °C and the Tm was 250 °C. Other cyclic compounds such as 1,4-cyclohexanedimethanol (CHDM) and 1,4-cyclohexanedicarboxylic acid (CHDA) were also preferred in polyester preparation [23,66,67,68,69]. Matos et al. [66] undertook a comparative study between poly(1,4-cyclohexylene 2,5-furandicarboxylate) (PCdF) and poly(1,4-cyclohexanedimethylene furandicarboxylate) (PCF). The high Tg of PCdF (175 °C) was attributed to the increasing stiffness of the linkage between cyclohexylene and the furan ring. Wang et al. [23] investigated the different steric conformation of PCF and found that the thermal and mechanical properties were heavily affected. When the trans moiety of CHDM increased, the Tm and Tg increased from 219 °C to 291 °C and 71 °C to 87 °C, respectively. Additionally, PCF-trans 98 exhibited the shortest crystallization half time at 7 s and the best mechanical properties (1820 MPa for tensile modulus and 52 MPa for tensile strength). Other cyclic monomers were also investigated and patented such as 2,2,4,4-tetramethylcyclobutane-1,3-diol [70], dianhydrohexitols [71], hydroquinone [72], and dichloro-2,3-o-isopropylidene L-tartrate [73]. Scheme 2 shows the structure of the main rigid diols above-mentioned. As shown in Table 2, the thermal and mechanical properties of polyesters with rigid diols are listed. It is worth mentioning that the mechanical information was absent for most of the references due to the brittleness of the main chain.

4. Co-Polyesters from FDCA

4.1. Balance between Rigidity and Flexibility

Taking PEF as an example again, as much interest has been garnered due to its promising properties compared to its terephthalate homolog. However, the brittleness and low crystallization rate limited the further application of PEF. Its modification is to soften the fragile chain and enhance the crystallizability, thus increasing the crystallization rate and the elongation at break [33,74]. As discussed previously, furan-polyesters with long-chain aliphatic diols exhibited improved elongation at break and crystallizability, nevertheless, their thermal transition temperature and mechanical properties deteriorated. It can be assumed that if PEF was softened by the incorporation of the linear diols, the originally good Tg, Tm, mechanical strength, and modulus would decline to substandard. Ma et al. [75] combined two frequently-used diols, ethylene glycol (EG) and 1,4-butylene glycol (BG), to prepare the poly[(ethylene 2,5-furandicarboxylate)-co-(butylene 2,5-furandicarboxylate)] (PEF-PBF). The thermal characteristics showed that the Tg decreased as the BG content increased. They confirmed that diols with longer chains exhibited higher reactivity with FDCA while the mechanical property was absent.
The cyclic monomers were found to be well balanced between the required high Tg and the need to toughen the PEF chain. In the report by Hong et al. [76], CHDM was employed in the synthesis of PEF via transesterification. Results showed that the elongation at break was increased to 79% as CHDM content increased, thus confirming that CHDM improved the ductility of the backbone. Naturally, the crystallization rate was also enhanced upon the introduction of CHDM as the additional strain-induced crystallization was observed in the tensile test. Then, after SSPC section, the polyesters obtained were successfully processed as bottles, which exhibited better oxygen permeability and less acetaldehyde content compared to PET-containing bottles. Liu et al. [77] prepared poly(ethylene-co-1,4-cyclohexanedimethylene 2,5-furandicarboxylate)s (PECFs) via transesterification and the products exhibited enhanced toughness with satisfying thermal and mechanical properties. The crystallizability and mechanical properties increased with the increasing CHDM moiety. Poly(ethylene 2,5-furandicarboxylate-co-ethylene 1,4-cyclohexanedicarboxylate)s (PEFCs) were also synthesized and investigated [78]. The thermal characteristics showed that the Tg decreased with the increasing content of CHDA and all samples showed thermal stability at 350 °C. Results showed that PEFC-30 behaved as a toughened polyester, since tensile modulus was 800 MPa and the elongation at break was more than 600%. Poly(butylene-co-1,4-cyclohexanedimethylene-2,5-furandicarboxylic acid) (PBCFs) were prepared via direct esterification. Thermal characteristics showed that the Tg (45.7–87.5 °C) and Tm (140.1–264.9 °C) of the polyesters obtained increased with the increasing CHDM moiety. The crystallinity and crystal structure were also affected by CHDM content as when the CHDM moiety increased, the polyesters became semi-crystalline from completely amorphous [79]. Jia et al. [80] employed CHDM to modify PPF. The co-polyesters showed satisfying good mechanical properties, although the barrier performance weakened a little. Terpolyesters based on FDCA, ISB, EG, and CHDM (PEICF) were prepared via direct esterification [81]. The polyesters obtained were all completely amorphous since no Tm was observed. The Tg increased linearly as the increasing isosorbide (ISB) moiety and the maximum one reached 182.4 °C when the ISB content was 100%. Poly(propylene-co-1,4-cyclohexanedimethylene 2,5-furandicarboxylate)s (PPCFs) showed enhanced Tg (59–73 °C) than PPF (56 °C) and comparable mechanical properties as elongation at break was over 150% for all co-polyesters [80]. 2,2,4,4-Tetramethyl-1,3-cyclobutanediol (CBDO) was another unique cyclic diol that has been employed in PEF modification [82]. The poly(ethylene-co-2,2,4,4-tetramethyl-1,3-cyclobutanediol 2,5-furandicarboxylate) (PETF) showed satisfying properties: Tg was around 90 °C and tensile modulus reached 3000 MPa with elongation at break of 10%. Additionally, the co-polyesters films showed high transparency with seven times and five times better barrier property for CO2 and O2 compared to PET. Zhou et al. [83] prepared poly(ethylene 2,5-furandicarboxylate)-poly(ethylene glycol) (PEFEGs) based on PEF and poly(ethylene glycol) (PEG). The Tg of the polyesters obtained was from 78 °C to 85 °C with an elongation at break up to 60%. With the increasing content of PEG, the shape fixity increased. They also tried to synthesize a series of furfural-based quadri-copolyesters based on FDCA, BDO, succinic anhydride (SA), and tricyclic diacid (TCDA) as an alternative to PBT. The polyesters obtained showed a satisfied Tg (36.7 °C) and tensile strength (46.9 MPa) with excellent ductility (elongation at break of 830%) [84]. Gao et al. [85] employed alicyclic (1R,3S)-1,2,2-trimethylcyclopentane-1,3-dimethanol (TCDM), which was derived from camphor in the polyester synthesis with linear or aromatic diacids. Polyesters based on FDCA showed a high Tg of 98 °C. Liu et al. [86] synthesized a series of novel polyesters with Tg higher than 100 °C from FDCA, BDO, and bis [4-(2-hydroxyethoxy) phenyl] sulfone (BHEPS). When the BHEPS content reached 65%, the Tg was 104.7 °C with Mw of 28,500 g/mol, tensile strength of 82 MPa, and elongation at break of 98%. Table 3 summarizes the properties of these polyesters, as discussed above.

4.2. Introduction of Biodegradability

Most FDCA homo-polyesters lack biodegradability. The introduction of monomers such as ε-caprolactone (CL), adipic acid (AA), and succinate to furan-polyesters, is a simple way to impart biodegradability to furanic polyesters. The co-polyesters based on these monomers and TPA have shown good biodegradability as poly(butylene adipate-coterephthalate) (PBAT) [87,88] and poly(butylene succinate-co-terephthalate) (PBST) [89,90].
CL is a famous building block of semi-crystalline polyester poly(ε-caprolactone) (PCL), which has been widely applied in the biomedical field due to its superior biodegradability, biocompatibility, and drug permeability [91,92,93]. Poly(ethylene 2,5-furandicarboxylate-co-ε-caprolactone) (PEFCL) was synthesized and the thermomechanical properties were discussed [94]. The Tg was found to decrease with the increase in CL moiety. The mechanical modulus and strength decreased slightly while elongation at break enhanced notably when the CL content continued growing. However, information on biodegradability was absent. Kasmi et al. [95] investigated CL-based furan-polyesters from 1,5-pentanediol and 1,6-hexanediol via ROP. The poly(pentylene 2,5-furandicarboxylate-co-caprolactone) (PPeCFs) and poly(hexamethylene 2,5-furandicarboxylate-co-caprolactone) (PHeCFs) were thermally stable at 310 and 360 °C, respectively. PPeF and PHeF degraded very slowly, which remained invariant after 25 days of testing, while the incorporation of over 40 mol% CL content obviously accelerated degradation as 32% and 13% weight loss was achieved by PHeCF 50/50 and PHeCF 60/40, respectively. Zheng et al. [96] synthesized an elastomer, poly(butylene 2,5-furandicarboxylate-ε-caprolactone) (PBFCL) via ROP with randomly distributed rigid and soft segments. The polyesters obtained showed good mechanical strength (25 MPa) and thermal stability (Td,5% > 320 °C). The CL content ensured the biodegradability of polyesters, however, the crystalline part hindered the enzymatic attack. The results confirmed that by simply introducing the CL units to the polyester, biodegradability could be obtained. Morales-Huerta et al. [97] prepared poly(ε-caprolactone-co-butylene 2,5-furandicarboxylate) by enzymatic ROP with Lipase Candida Antarctica (CALB) as the catalyst. The advantages of enzymatic ROP can be concluded as mild conditions, no volatile by-products, and low coloration with the Mw between 22,000–50,000 g/mol. The BF content increased the thermal transition temperature of co-polyesters while slowing down the crystallization rate. The result clarified that the co-polyesters obtained remained unchangeable in the hydrolytic degradation while exhibited notable weight loss (ranging from 20% to 40%) in the presence of lipases. Hu et al. [98] prepared poly(butylene furandicarboxylate-co-ε-caprolactone) (PBFCL) co-polyesters where the results showed that after 48 days with enzymes, the weight loss could be over 30% and when the CL content was more than 40 mol%, the co-polyesters exhibited good recoverability.
Succinate and adipate are also common monomers for co-polymerization of furan-polyesters. The aliphatic-aromatic co-polyesters employing succinate such as PBST and poly(butylene succinate-co-2,5-furandicarboxylate) (PBSF) exhibited the desired biodegradability and good thermomechanical properties [99,100,101]. Poly(butylene 2,5-furandicarboxylate-co-succinate) (coPBFxSy) was synthesized via ROP [102]. Enzymatic degradation revealed that the succinate moiety could improve the degradation process while PBF remained stable under the same conditions. Poly(1,3-propylene 2,5-furandicarboxylate-co-1,3-propylene succinate) (PPFPS) synthesized via polycondensation and azeotropic distillation have been investigated and a kinetic model for furan-polyesters was suggested [103]. Results showed that the rate of polymerization would be promoted by the elevation of temperature as expected, while at around 230 °C, the polymerization rate was decreased due to the increasing monomer evaporation. Wu et al. [104] also employed succinate in the polymerization of PBF. Results showed that PBSF was less crystallizable than poly(butylene succinate), but still had fast crystallization behavior in isothermal and non-isothermal characterization. The elongation at break, tensile modulus, and tensile strength were 230–758%, 330–680 MPa, and 18–56 MPa, respectively. The information of degradation was not discussed. Later, the same group [105,106] synthesized and discussed the degradation behavior of poly(butylene adipate-co-butylene furandicarboxylate) (PBAF) and poly(butylene succinate-co-butylene furandicarboxylate) (PBSF). Results showed that in the 110 day composting test, all samples accomplished more than 90% biodegradability. Unsurprisingly, higher BF content led to slower degradation as the alkaline conditions would accelerate the weight loss more than acidic and neutral conditions due to the better water solubility and diffusibility of degradation products. Compared to their TPA-based homologs, the furan-polyesters were easier to degrade. Hu et al. [107] synthesized a series of poly(propylene succinate-co-furandicarboxylate) (PPSF) ranging from poly(propylene succinate) (PPS) to PPF. It is clear that degradation would be improved with the increase in succinate content as PPSF60 exhibited the lowest weight loss of 3.5% and the weight loss of PPSF10 was 14.8%. Besides, the co-polyesters obtained showed good mechanical performance and excellent barrier property, which determined its potential for green packaging applications. Adipate has been successfully introduced in the synthesis of TPA-based plastics that reached the perfect combination of mechanical property and biodegradability such as poly(butylene adipate-co-butylene terephthalate), which has been commercialized as Ecoflex@ and Easter-bio@ [108,109]. Therefore, it is of interest and promise to introduce adipate into FDCA-based polyesters targeting biodegradability and tough products. Papadopoulos et al. [110] prepared co-polyesters based on PEF and poly(ethylene adipate) (PEAd) with good thermal stability as Td was over 300 °C. Results showed that when the content of ethylene adipate was over 90%, the co-polyesters could degrade completely after a month. Zhou et al. [111] investigated poly(butylene adipate-co-butylene furandicarboxylate (PBAFs), which showed biodegradability with FDCA content at a range of 0–50 mol%. When the FDCA moiety rose to 75%, the co-polyesters were un-degradable. Wu et al. [112] conducted a comparative study between PBF and PBAF. Results showed that PBAFs were all semi-crystalline polyesters and manifested high-elastic deformation with BF content under 50 mol%. With higher BF moiety, PBSFs performed as non-rigid plastics (42–110 MPa and 30–42 MPa for tensile modulus and strength) while PBF behaved as tough plastics (1900 MPa and 56 MPa for tensile modulus and strength). Both showed high elongation at break (>250%).
Other biodegradable monomers have also been investigated in the polycondensation of furan-polyesters. Hu et al. [113] synthesized poly(butylene furandicarboxylate-co-glycolate) (PBFGA) via melt polycondensation. The polyesters obtained showed good thermal stability (>320 °C), mechanical properties (500–1260 MPa for tensile modulus and 16–44 MPa for tensile strength), and over 220% for elongation at break while exhibiting a superior gas barrier property to the commercial PBAT films. Furthermore, the enzymatic degradation test showed an obvious weight loss when glycolic acid content was more than 20 mol%. Soccio et al. [114] synthesized poly(butylene 2,5-furanoate/diglycolate) P(BFxDGy) random co-polyesters via melt polycondensation. Results showed that the samples exhibited higher thermal stability (Tmax > 380 °C) than PBF (Tmax = 366 °C). The degradation rate of the co-polyesters depended strictly on the composition and crystallinity, where P(BF60DG40) showed more than 40% weight loss after 62 days while P(BF90DG10) showed about 10% weight loss under the same conditions. The mechanical properties deteriorated with the increasing DG units. Wu et al. [115] prepared co-polyesters based on lactic acid and ethylene glycol (PEFL). The Td ranged from 298 °C to 398 °C since more EF content would enhance the thermal stability. After 55 days in PBS buffer solution and in soil, the PEFL20/80 exhibited weight loss of 24% and 65%, respectively. Jia et al. [116] completely synthesized amorphous co-polyesters from EG and dodecanedioic acid (DDCA) via transesterification. The elongation at break was over 200% for all samples except for PEDF-10. PEDF-60 exhibited the best biodegradability of over 5% weight loss while PEF and PEDF-20 showed no weight loss during the 28 day test. Polyester based on 2,5-furandicarboxylic acid and a very-long aliphatic monomer, 1,20-eicosanediol, was synthesized and investigated [117]. The incubation result showed that after 21 days, the poly(1,20-eicosanediyl 2,5-furandicarboxylate) lost about 11% weight in the presence of Porcine pancreas enzyme. Weinberger et al. [118] investigated the PEF enzymatic degradation and found that the influence of particle size became smaller when the molecular weight was higher and they optimized the hydrolysis conditions such as in the 1 M phosphate buffer, pH 8, at 65 °C with Humicola insolens, where PEF reached 100% hydrolysis. Zhou et al. [119] synthesized a series of biodegradable polyesters with FDCA, BDO, and oxabicyclic units via ROP. The polyesters obtained showed biodegradability within seven days with less than 30 mol% FDCA units. Furthermore, the polyesters exhibited good mechanical (34 MPa of tensile strength) and barrier properties. Koo et al. [120] synthesized poly(butylene adipate-co-furanoate) (PBAF) and compared the properties with PBAT. PBAF showed instant elastic recovery in the stretching test while PBAT malfunctioned. In the enzymatic degradation, only the PBAF with 50% FDCA went through hydrolysis due to the specificity of the enzyme.

5. Composites of Furan-Polyesters

5.1. Inorganic and Organic Fillers

Polymer composites promoted by the advances of nanotechnology have been regarded as an indispensable part of polymer sciences. In fact, nanocomposites have been widely applied in the biomedical, electronic, automotive fields, etc. [121,122]. As furan-polyesters with such great potential have manifested, the field has been further extended by the evaluation of furan-contained composites. In fact, carbon-based composites of polymers have been reported to enhance the thermal, mechanical, and electrical properties [123,124,125,126]. Multi-walled carbon nanotubes (MWCNTs) and graphene oxide (GO) have been employed in the preparation of PEF nanocomposites, which showed faster crystallization than the neat material due to the higher nucleation density [127]. Generally speaking, a good dispersion of nano-fillers results in elevated nucleation sites, thus explaining the increasing crystallinity and crystallization rate. The crystal structures were different: only the α-crystal phase was observed in nanocomposites rather than both α- and β-crystals, which appeared in pristine PEF. However, the thermal stability of the nanocomposites weakened as those containing GO showed a Td of 370 °C. Terzopoulou et al. [128] prepared poly(propylene 2,5-furan dicarboxylate)/graphene nanocomposites and investigated the thermal decomposition kinetics and mechanisms. The thermal analysis showed that the graphene had no obvious influence on degradation, but resulted in a tiny increase in activation energy while the degradation mechanism was not affected by the presence of graphene. Poly(trimethylene 2,5-furanoate) (PTF) was also prepared as nanocomposites with few-layer graphene (FLG) via in situ polymerization [129]. Compared to the pristine PTF, the nano-reinforced composites exhibited improved mechanical properties and thermal stability. The crystallization behavior was not affected by the FLG. Achilias et al. [130] synthesized PEF composites with SiO2 and TiO2. The effect of nano-fillers on the SSPC was investigated and the presence of additives resulted in higher transesterification rate constants and lower activation energies. To enhance the slow crystallization of PBF, Zhou et al. [131] synthesized PBF/TiO2 nanocomposites via in situ polymerization with rutile/anatase TiO2. The crystallization rate of the composites obtained was less than 1 min at a TiO2 content of 7‰. Additionally, the mechanical properties were proven as the elongation at break and impact strength increased by 118% and 200%.
The well-dispersed clay on the polyester matrix could reinforce the thermal, mechanical, and barrier properties, but showed a possible negative impact on molecular weight and the thermal stability of the composites [132,133,134]. Organo-modified montmorillonite (OMMT) clays were employed in the preparation of PEF nanocomposites via the solvent casting method and the crystallization of the composite was improved due to the nucleation effect of the clays [135]. Additionally, the thermal stability was significantly enhanced since the initial degradation temperature was increased by 20 °C. Aluminosilicate clays were utilized in the preparation of PPF composites to enhance the crystallization rate of the pristine PPF [136]. The samples presented good thermal stability (>300 °C) and the β-scission was found to be the dominating mechanism in the degradation process. To improve the low crystallization rate of PEF, PEF/nanocrystalline cellulose composites were prepared via solvent casting and the sonication procedures were utilized to well-disperse the cellulose nanocrystals. The thermal measurements revealed a slight improvement in the thermal degradation of nanocomposites [137]. There are many reports highlighting the significance of nanocellulose in the reinforcement materials since they have high crystallinity and strong hydrogen bonds that apply greater barrier properties [138,139,140]. Further investigations have shown that PEF/cellulose composites prepared via twin screw extrusion exhibited faster crystallization, as expected [141]. Composites based on poly(butylene 2,5-furanoate), poly(butylene 2,5-furanoate)-co-(butylene diglycolate), and bacterial cellulose were investigated by Matos et al. [142]. Results showed that the Tg was in the range from −25 °C to 46 °C and Tm was between 61 °C and 174 °C. The composites obtained exhibited highly reinforced mechanical properties such as 1240 MPa for tensile modulus, 14 MPa for tensile strength, and elongation at break up to 240%. In addition, the barrier property of composites was superior to that of PBF. Other nano-clays such as organically modified sepiolite-based and montmorillonite-based PEF composites were prepared via melt extrusion [143]. The thermal stability was improved by the clays used while the thermal and rheological properties were unaffected.

5.2. Blends of Furanoates

It is well-known that reactive blending, which is applied to promote the miscibility between furanoates and other polymers at high temperature, is an efficient way to obtain homogenous polyester blends [144]. Poulopoulou et al. [145] prepared a series of polyester blends based on PEF, PBF, and PPF and their terephthalate homologs via reactive blending. The blends of PEF/PPF exhibited dynamic homogeneity and miscibility, but PPF/PBF and PEF/PBF showed immiscibility. The temperature and time were proven to be the crucial factor to control the miscibility of furan-polyester blends. Additionally, the composition and the catalysts play an essential part in the blending process. They further investigated the five different PBF blends with PLA, PET, PPT, PC, and poly(butylene 2,6-naphthalenedicarboxylate). The samples exhibited immiscibility except for PET/PBF. Moreover, they suggested that reactive blending can improve the homogeneity of immiscible blends such as PBN/PBF blends [146]. Interestingly, when the backbone of the two polyesters differed by a single methylene unit, the blends would probably be dynamically homogeneous, while being heterogeneous if differing by two methylene units [147]. They introduced poly(ethylene terephthalate-co-ethylene furanoate) to the PEF/PET blends and the miscibility of the blends was improved. Furthermore, they assumed that any furan-polyester blends with PC were essentially immiscible [148,149]. Long et al. [150] prepared PBF via transesterification and then blended it with PLA to make a composite in which PBF was dispersed from the sphere to fibril phase. Results showed that the elongation at break of the blends was 17 times better than the neat PLA, while the mechanical modulus and strength stayed the same. In addition, the impact strength was also enhanced compared to that of the pristine PBF and PLA. They further investigated the retroreflection in PLA/PBF blends and found that the separated phase or semi-crystallization were the main structure in the blends, which was compared to the tapetum lucidum in the cat’s eye by the authors [151].

6. Conclusions

Making materials from renewable resources is what a sustainable economy requires. Standing out from the bio-based polymers, furan-polyesters are on the way from academia to industry. While one of the crucial issues is the price of raw materials, FDCA, as produced from the oxidation of 5-hydroxymethylfurfural and mainly catalyzed by noble metals from biomass (fructose and lignin), cannot meet the requirement for industrial preparation. Therefore, efficient, economic, and cheap catalysts for FDCA production are urgently needed. As reviewed above, polyester materials based on FDCA present encouraging properties as they satisfy the thermomechanical property, which lays the groundwork for a broad application range, excellent barrier property for packaging materials, and biodegradability for a cyclic economy. Nevertheless, furanic polyesters still suffer from discoloration due to the substandard purification of FDCA and decarboxylation or other side reactions. An investigation of colorless transparent furanic polyesters was reported in 2021 [152]. Another is the unexpected susceptibility of PEF to UV irradiation, which showed obvious degradation signs such as discoloration, deteriorating thermal properties, and chain cross-linking [153]. In summary, with further synthetic advances and structural manipulation for designated properties, FDCA-based polyester materials are inspiring in a variety of areas.

Author Contributions

Software, Y.J.; writing-original draft preparation, X.F.; writing-review and editing, J.W.; visualization, X.Z. and Z.J.; Conceptualization, supervision and funding acquisition, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors greatly thank the financial support from the Zhejiang Provincial Key Scientific Research Programs (2021C01061), Zhejiang Provincial Natural Science Foundation of China (LR20E030001, LQ19E030009, and LGG21B040001), National Natural Science Foundation of China (NSFC 21975270), National Key R&D Program of China (2021YFB3700300), and Ningbo 2025 Key Scientific Research Programs (2018B10013).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vilela, C.; Sousa, A.F.; Fonseca, A.C.; Serra, A.C.; Coelho, J.F.J.; Freire, C.S.R.; Silvestre, A.J.D. The quest for sustainable polyesters—Insights into the future. Polym. Chem. 2014, 5, 3119–3141. [Google Scholar] [CrossRef]
  2. Fei, F.; Wang, J.G.; Zhu, J.; Wang, X.Z.; Liu, X.Q. Bio-based poly(ethylene 2,5-furancoate): No longer an alternative, but an irreplaceable one in polymer industry. ACS Sustain. Chem. Eng. 2020, 8, 8471–8485. [Google Scholar] [CrossRef]
  3. Geyer, R.; Jambeck, J.R.; Law, K.L. Production, use, and fate of all plastics ever made. Sci. Adv. 2017, 3, e1700782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Rhim, J.W. Preparation and characterization of vacuum sputter silver coated PLA film. LWT-Food Sci. Technol. 2013, 54, 477–484. [Google Scholar] [CrossRef]
  5. Sudesh, K.; Abe, H.; Doi, Y. Synthesis, structure and properties of polyhydroxyalkanoates: Biological polyesters. Prog. Polym. Sci. 2000, 25, 1503–1555. [Google Scholar] [CrossRef]
  6. Takahashi, K.; Taniguchi, I.; Miyamoto, M.; Kimura, Y. Melt/solid polycondensation of glycolic acid to obtain high-molecular-weight poly(glycolic acid). Polymer 2000, 41, 8725–8728. [Google Scholar] [CrossRef]
  7. Ye, H.M.; Wang, R.D.; Liu, J.; Xu, J.; Guo, B.H. Isomorphism in poly(butylene succinate-co-butylene fumarate) and its application as polymeric nucleating agent for poly(butylene succinate). Macromolecules 2012, 45, 5667–5675. [Google Scholar] [CrossRef]
  8. Gandini, A.; Coelho, D.; Gomes, M.; Reis, B.; Silvestre, A. Materials from renewable resources based on furan monomers and furan chemistry: Work in progress. J. Mater. Chem. 2009, 19, 8656–8664. [Google Scholar] [CrossRef] [Green Version]
  9. Gandini, A.; Silvestre, A.J.D.; Neto, C.P.; Sousa, A.F.; Gomes, M. The furan counterpart of poly(ethylene terephthalate): An alternative material based on renewable resources. J. Polym. Sci. Polym. Chem. 2009, 47, 295–298. [Google Scholar] [CrossRef]
  10. Gomes, M.; Gandini, A.; Silvestre, A.J.D.; Reis, B. Synthesis and characterization of poly(2,5-furan dicarboxylate)s based on a variety of diols. J. Polym. Sci. Polym. Chem. 2011, 49, 3759–3768. [Google Scholar] [CrossRef]
  11. Jiang, M.; Liu, Q.; Zhang, Q.; Ye, C.; Zhou, G. A series of furan-aromatic polyesters synthesized via direct esterification method based on renewable resources. J. Polym. Sci. Polym. Chem. 2012, 50, 1026–1036. [Google Scholar] [CrossRef]
  12. Grosshardt, O.; Fehrenbacher, U.; Kowollik, K.; Tuebke, B.; Dingenouts, N.; Wilhelm, M. Synthesis and properties of polyester and polyamide based on furan 2,5 dycarboxylic acid. Chem. Ing. Tech. 2009, 81, 1829–1835. [Google Scholar]
  13. Jiang, Y.; Maniar, D.; Woortman, A.J.; Alberda van Ekenstein, G.O.; Loos, K. Enzymatic polymerization of furan-2,5-dicarboxylic acid-based furanic-aliphatic polyamides as sustainable alternatives to polyphthalamides. Biomacromolecules 2015, 16, 3674–3685. [Google Scholar] [CrossRef]
  14. Luo, K.; Wang, Y.; Yu, J.; Zhu, J.; Hu, Z. Semi-bio-based aromatic polyamides from 2,5-furandicarboxylic acid: Toward high-performance polymers from renewable resources. Rsc Adv. 2016, 6, 87013–87020. [Google Scholar] [CrossRef]
  15. Cao, M.; Zhang, C.; He, B.; Huang, M.; Jiang, S. Synthesis of 2,5-furandicarboxylic acid-based heat-resistant polyamides under existing industrialization process. Macromol. Res. 2017, 25, 722–729. [Google Scholar] [CrossRef]
  16. Hu, F.; La Scala, J.J.; Sadler, J.M.; Palmese, G.R. Synthesis and characterization of thermosetting furan-based epoxy systems. Macromolecules 2014, 47, 3332–3342. [Google Scholar] [CrossRef]
  17. Deng, J.; Liu, X.Q.; Li, C.; Jiang, Y.H.; Zhu, J. Synthesis and properties of a bio-based epoxy resin from 2,5-furandicarboxylic acid (FDCA). Rsc Adv. 2015, 5, 15930–15939. [Google Scholar] [CrossRef]
  18. Bao, F.; Song, Y.; Liu, Q.; Song, C.; Liu, C.; Wang, J.; Jian, X.; Xiao, J. Partial bio-based poly (aryl ether ketone) derived from 2,5-furandicarboxylic acid with enhanced processability. Polym. Degrad. Stab. 2019, 161, 309–318. [Google Scholar] [CrossRef]
  19. Lu, Z.; Xiaolan, L.; Yusheng, Q.; Yebo, L. A novel 2,5-furandicarboxylic acid-based bis(cyclic carbonate) for the synthesis of biobased non-isocyanate polyurethanes. Rsc Adv. 2016. [Google Scholar]
  20. María Nelly García, G.; Pål, B.; Marinella, L.; Stefano, T. Development and life cycle assessment of polyester binders containing 2,5-furandicarboxylic acid and their polyurethane coatings. J. Polym. Environ. 2018. [Google Scholar]
  21. Zhu, J.H.; Cai, J.L.; Xie, W.C.; Chen, P.H.; Gazzano, M.; Scandola, M.; Gross, R.A. Poly(butylene 2,5-furan dicarboxylate), a biobased alternative to PBT: Synthesis, physical properties, and crystal structure. Macromolecules 2013, 46, 796–804. [Google Scholar] [CrossRef]
  22. Eerhart, A.J.J.E.; Faaij, A.P.C.; Patel, M.K. Replacing fossil based PET with biobased PEF; process analysis, energy and GHG balance. Energy Environ. Sci. 2012, 5, 6407–6422. [Google Scholar] [CrossRef]
  23. Wang, J.G.; Liu, X.Q.; Jia, Z.; Sun, L.Y.; Zhu, J. Highly crystalline polyesters synthesized from furandicarboxylic acid (FDCA): Potential bio-based engineering plastic. Eur. Polym. J. 2018, 109, 379–390. [Google Scholar] [CrossRef]
  24. Van Putten, R.J.; van der Waal, J.C.; de Jong, E.; Rasrendra, C.B.; Heeres, H.J.; de Vries, J.G. Hydroxymethylfurfural, a versatile platform chemical made from renewable resources. Chem. Rev. 2013, 113, 1499–1597. [Google Scholar] [CrossRef] [PubMed]
  25. de Jong, E.; Dam, M.A.; Sipos, L.; Gruter, G.-J.M. Furandicarboxylic Acid (FDCA), A Versatile Building Block for a Very Interesting Class of Polyesters. In Biobased Monomers, Polymers and Materials; Smith, P.B., Gross, R., Eds.; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2012; pp. 1–13. [Google Scholar]
  26. Albonetti, S.; Lolli, A.; Morandi, V.; Migliori, A.; Lucarelli, C.; Cavani, F. Conversion of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid over Au-based catalysts: Optimization of active phase and metal–support interaction. Appl. Catal. B 2015, 163, 520–530. [Google Scholar] [CrossRef]
  27. Martuscelli, E.; Pedone, C. The crystal and molecular structure of furane-α,α’-dicarboxylic acid. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 1968, 24, 175–179. [Google Scholar] [CrossRef]
  28. Bailey, M.; Brown, C.J. The crystal structure of terephthalic acid. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater. 1967, 22, 387–391. [Google Scholar] [CrossRef]
  29. Zhang, H.; Zhou, G.; Jiang, M.; Zhang, H.; Wang, H.; Wu, Y.; Wang, R. Bio-based polyesters with high glass-transition temperatures and gas barrier properties derived from renewable rigid tricyclic diacid or tetracyclic anhydride. Macromolecules 2020, 53, 5475–5486. [Google Scholar] [CrossRef]
  30. Yu, X.; Jia, J.; Xu, S.; Lao, K.U.; Sanford, M.J.; Ramakrishnan, R.K.; Nazarenko, S.I.; Hoye, T.R.; Coates, G.W.; DiStasio, R.A. Unraveling substituent effects on the glass transition temperatures of biorenewable polyesters. Nat. Commun. 2018, 9, 2880. [Google Scholar] [CrossRef]
  31. Drewitt, J.G.; Lincoln, J. Improvement in polymers. Br. Patent 621,971, 12 November 1946. [Google Scholar]
  32. Morales-Huerta, J.C.; de Ilarduya, A.M.; Munoz-Guerra, S. Poly(alkylene 2,5-furandicarboxylate)s (PEF and PBF) by ring opening polymerization. Polymer 2016, 87, 148–158. [Google Scholar] [CrossRef] [Green Version]
  33. Knoop, R.J.I.; Vogelzang, W.; van Haveren, J.; van Es, D.S. High molecular weight poly(ethylene-2,5-furanoate); critical aspects in synthesis and mechanical property determination. J. Polym. Sci. Polym. Chem. 2013, 51, 4191–4199. [Google Scholar] [CrossRef]
  34. Gantillon, B.; Spitz, R.; McKenna, T.E. The solid state postcondensation of PET: A review of the physical and chemical processes taking place in the solid state. Angew. Makromol. Chem. 2004, 289, 88–105. [Google Scholar]
  35. Banella, M.B.; Bonucci, J.; Vannini, M.; Marchese, P.; Lorenzetti, C.; Celli, A. Insights into the synthesis of poly(ethylene 2,5-furandicarboxylate) from 2,5-furandicarboxylic acid: Steps toward environmental and food safety excellence in packaging applications. Ind. Eng. Chem. Res. 2019, 58, 8955–8962. [Google Scholar] [CrossRef]
  36. Qu, X.L.; Zhou, G.Y.; Wang, R.; Zhang, H.Y.; Wang, Z.P.; Jiang, M.; Tang, J. Insights into high molecular weight poly(ethylene 2,5-furandicarboxylate) with satisfactory appearance: Roles of in-situ catalysis of metal zinc. J. Ind. Eng. Chem. 2021, 99, 422–430. [Google Scholar] [CrossRef]
  37. Burgess, S.K.; Karvan, O.; Johnson, J.R.; Kriegel, R.M.; Koros, W.J. Oxygen sorption and transport in amorphous poly(ethylene furanoate). Polymer 2014, 55, 4748–4756. [Google Scholar] [CrossRef]
  38. Burgess, S.K.; Kriegel, R.M.; Koros, W.J. Carbon dioxide sorption and transport in amorphous poly(ethylene furanoate). Macromolecules 2015, 48, 2184–2193. [Google Scholar] [CrossRef]
  39. Burgess, S.K.; Leisen, J.E.; Kraftschik, B.E.; Mubarak, C.R.; Kriegel, R.M.; Koros, W.J. Chain mobility, thermal, and mechanical properties of poly(ethylene furanoate) compared to poly(ethylene terephthalate). Macromolecules 2014, 47, 1383–1391. [Google Scholar] [CrossRef]
  40. Sun, L.; Wang, J.; Mahmud, S.; Jiang, Y.; Zhu, J.; Liu, X. New insight into the mechanism for the excellent gas properties of poly(ethylene 2,5-furandicarboxylate) (PEF): Role of furan ring’s polarity. Eur. Polym. J. 2019, 118, 642–650. [Google Scholar] [CrossRef]
  41. Papageorgiou, G.Z.; Papageorgiou, D.G.; Tsanaktsis, V.; Bikiaris, D.N. Synthesis of the bio-based polyester poly(propylene 2,5-furan dicarboxylate). Comparison of thermal behavior and solid state structure with its terephthalate and naphthalate homologues. Polymer 2015, 62, 28–38. [Google Scholar] [CrossRef]
  42. Tsanaktsis, V.; Papageorgiou, G.Z.; Bikiaris, D.N. A facile method to synthesize high-molecular-weight biobased polyesters from 2,5-furandicarboxylic acid and long-chain diols. J. Polym. Sci. Polym. Chem. 2015, 53, 2617–2632. [Google Scholar] [CrossRef]
  43. Avantium. Avantium Technology & Markets Day: “Entering the Commercialization Phase for PEF”. Available online: https://www.avantium.com/press-releases/avantium-technology-markets-day-entering-the-commercialization-phase-for-pef/ (accessed on 25 July 2018).
  44. Maltsev, A. PEF Challenges PET to Battle. Available online: https://ethz.ch/en/news-and-events/eth-news/news/2018/07/pef-for-pet.html (accessed on 3 August 2018).
  45. Rosenboom, J.G.; Hohl, D.K.; Fleckenstein, P.; Storti, G.; Morbidelli, M. Bottle-grade polyethylene furanoate from ring-opening polymerisation of cyclic oligomers. Nat. Commun. 2018, 9, 2701. [Google Scholar] [CrossRef] [Green Version]
  46. Tsanaktsis, V.; Vouvoudi, E.; Papageorgiou, G.Z.; Papageorgiou, D.G.; Chrissafis, K.; Bikiaris, D.N. Thermal degradation kinetics and decomposition mechanism of polyesters based on 2,5-furandicarboxylic acid and low molecular weight aliphatic diols. J. Anal. Appl. Pyrolysis 2015, 112, 369–378. [Google Scholar] [CrossRef]
  47. Papageorgiou, G.Z.; Papageorgiou, D.G.; Terzopoulou, Z.; Bikiaris, D.N. Production of bio-based 2,5-furan dicarboxylate polyesters: Recent progress and critical aspects in their synthesis and thermal properties. Eur. Polym. J. 2016, 83, 202–229. [Google Scholar] [CrossRef]
  48. Sousa, A.F.; Vilela, C.; Fonseca, A.C.; Matos, M.; Freire, C.S.R.; Gruter, G.-J.M.; Coelho, J.F.J.; Silvestre, A.J.D. Biobased polyesters and other polymers from 2,5-furandicarboxylic acid: A tribute to furan excellency. Polym.Chem. 2015, 6, 5961–5983. [Google Scholar] [CrossRef]
  49. Jiang, Y.; Woortman, A.J.J.; Van Ekenstein, G.O.R.A.; Loos, K. A biocatalytic approach towards sustainable furanic–aliphatic polyesters. Polym. Chem. 2015, 6, 5198–5211. [Google Scholar] [CrossRef] [Green Version]
  50. Ma, J.P.; Yu, X.F.; Xu, J.; Pang, Y. Synthesis and crystallinity of poly(butylene 2,5-furandicarboxylate). Polymer 2012, 53, 4145–4151. [Google Scholar] [CrossRef]
  51. Thiyagarajan, S.; Vogelzang, W.; Knoop, R.J.I.; Frissen, A.E.; van Haveren, J.; van Es, D.S. Biobased furandicarboxylic acids (FDCAs): Effects of isomeric substitution on polyester synthesis and properties. Green Chem. 2014, 16, 1957–1966. [Google Scholar] [CrossRef] [Green Version]
  52. Zhao, M.; Zhang, C.L.; Yang, F.; Weng, Y.X. Gas barrier properties of furan-based polyester films analyzed experimentally and by molecular simulations. Polymer 2021, 233, 124200. [Google Scholar] [CrossRef]
  53. Fenouillot, F.; Rousseau, A.; Colomines, G.; Saint-Loup, R.; Pascault, J.P. Polymers from renewable 1,4:3,6-dianhydrohexitols (isosorbide, isomannide and isoidide): A review. Prog. Polym. Sci. 2010, 35, 578–622. [Google Scholar] [CrossRef]
  54. Okada, M.; Tsunoda, K.; Tachikawa, K.; Aoi, K. Biodegradable polymers based on renewable resources. IV. Enzymatic degradation of polyesters composed of 1,4:3.6-dianhydro-D-glucitol and aliphatic dicarboxylic acid moieties. J. Appl. Polym. Sci. 2000, 77, 338–346. [Google Scholar] [CrossRef]
  55. Caouthar, A.; Roger, P.; Tessier, M.; Chatti, S.; Blais, J.C.; Bortolussi, M. Synthesis and characterization of new polyamides derived from di(4-cyanophenyl)isosorbide. Eur. Polym. J. 2007, 43, 220–230. [Google Scholar] [CrossRef]
  56. Chatti, S.; Schwarz, G.; Kricheldorf, H.R. Cyclic and noncyclic polycarbonates of isosorbide (1,4:3,6-dianhydro-D-glucitol). Macromolecules 2006, 39, 9064–9070. [Google Scholar] [CrossRef]
  57. Lomelí-Rodríguez, M.; Corpas-Martínez, J.R.; Willis, S.; Mulholland, R.; Lopez-Sanchez, J.A. Synthesis and Characterization of Renewable Polyester Coil Coatings from Biomass-Derived Isosorbide, FDCA, 1,5-Pentanediol, Succinic Acid, and 1,3-Propanediol. Polymers 2018, 10, 600. [Google Scholar] [CrossRef] [Green Version]
  58. Terzopoulou, Z.; Kasmi, N.; Tsanaktsis, V.; Doulakas, N.; Bikiaris, D.N.; Achilias, D.S.; Papageorgiou, G.Z. Synthesis and Characterization of Bio-Based Polyesters: Poly(2-methyl-1,3-propylene-2,5-furanoate), Poly(isosorbide-2,5-furanoate), Poly(1,4-cyclohexanedimethylene-2,5-furanoate). Materials 2017, 10, 801. [Google Scholar] [CrossRef] [Green Version]
  59. Chebbi, Y.; Kasmi, N.; Majdoub, M.; Cerruti, P.; Scarinzi, G.; Malinconico, M.; Dal Poggetto, G.; Papageorgiou, G.Z.; Bikiaris, D.N. Synthesis, Characterization, and Biodegradability of Novel Fully Biobased Poly(decamethylene-co-isosorbide 2,5-furandicarboxylate) Copolyesters with Enhanced Mechanical Properties. ACS Sustain. Chem. Eng. 2019, 7, 5501–5514. [Google Scholar] [CrossRef]
  60. Chen, J.; Lin, Y.; Chen, Y.; Koning, C.E.; Wu, J.; Wang, H. Low-crystallinity to highly amorphous copolyesters with high glass transition temperatures based on rigid carbohydrate-derived building blocks. Polym. Int. 2021, 70, 536–545. [Google Scholar] [CrossRef]
  61. Wang, X.; Wang, Q.; Liu, S.; Wang, G. Synthesis and characterization of poly(isosorbide-co-butylene 2,5-furandicarboxylate) copolyesters. Eur. Polym. J. 2019, 115, 70–75. [Google Scholar] [CrossRef]
  62. Kasmi, N.; Majdoub, M.; Papageorgiou, G.Z.; Bikiaris, D.N. Synthesis and crystallization of new fully renewable resources-based copolyesters: Poly(1,4-cyclohexanedimethanol-co-isosorbide 2,5-furandicarboxylate). Polym. Degrad. Stab. 2018, 152, 177–190. [Google Scholar] [CrossRef]
  63. Wu, J.; Eduard, P.; Jasinska-Walc, L.; Rozanski, A.; Noordover, B.A.J.; van Es, D.S.; Koning, C.E. Fully Isohexide-Based Polyesters: Synthesis, Characterization, and Structure–Properties Relations. Macromolecules 2013, 46, 384–394. [Google Scholar] [CrossRef]
  64. Jacquel, N.; Saint-Loup, R.; Pascault, J.-P.; Rousseau, A.; Fenouillot, F. Bio-based alternatives in the synthesis of aliphatic–aromatic polyesters dedicated to biodegradable film applications. Polymer 2015, 59, 234–242. [Google Scholar] [CrossRef]
  65. Wu, J.; Eduard, P.; Thiyagarajan, S.; Noordover, B.A.; van Es, D.S.; Koning, C.E. Semi-aromatic polyesters based on a carbohydrate-derived rigid diol for engineering plastics. ChemSusChem 2015, 8, 67–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Matos, M.; Sousa, A.F.; Silvestre, A.J.D. Improving the thermal properties of poly(2,5-furandicarboxylate)s using cyclohexylene moieties: A comparative study. Macromol. Chem. Phys. 2016, 218, 1600492. [Google Scholar] [CrossRef]
  67. Celli, A.; Marchese, P.; Sisti, L.; Dumand, D.; Sullalti, S.; Totaro, G. Effect of 1,4-cyclohexylene units on thermal properties of poly(1,4-cyclohexylenedimethylene adipate) and similar aliphatic polyesters. Polym. Int. 2013, 62, 1210–1217. [Google Scholar] [CrossRef]
  68. Celli, A.; Marchese, P.; Sullalti, S.; Berti, C.; Barbiroli, G. Eco-friendly poly(butylene 1,4-cyclohexanedicarboxylate): Relationships between stereochemistry and crystallization behavior. Macromol. Chem. Phys. 2011, 212, 1524–1534. [Google Scholar] [CrossRef]
  69. Wang, J.G.; Liu, X.Q.; Jia, Z.; Sun, L.Y.; Zhang, Y.J.; Zhu, J. Modification of poly(ethylene 2,5-furandicarboxylate) (PEF) with 1, 4-cyclohexanedimethanol: Influence of stereochemistry of 1,4-cyclohexylene units. Polymer 2018, 137, 173–185. [Google Scholar] [CrossRef]
  70. Carman, H.S.; Killman, J.I.; Crawford, E.D.; Jenkins, J.C.; Eastman Chemical Company. Polyester Compositions Containing Furandicarboxylic Acid or an Ester Thereof, and 2,2,4,4-tetramethyl-1,3-cyclobutanediol. U.S. Patent 2,013,055,862, 18 April 2013. [Google Scholar]
  71. Ghosh, T.; Mahajan, K.; Narayan-Sarathy, S.; Balgacem, M.N.; Gopalakrishnan, P. 2,5-Furan Dicarboxylic Acid-based Polyesters Prepared from Biomass. U.S. Patent 20,130,171,397A1, 4 July 2013. [Google Scholar]
  72. Zaidi, S.; Soares, M.J.; Bougarech, A.; Thiyagarajan, S.; Guigo, N.; Abid, S.; Abid, M.; Silvestre, A.J.D.; Sousa, A.F. Unravelling the para- and ortho-benzene substituent effect on the glass transition of renewable wholly (hetero-)aromatic polyesters bearing 2,5-furandicarboxylic moieties. Eur. Polym. J. 2021, 150, 110413. [Google Scholar] [CrossRef]
  73. Bengs, H.; Schoenfeld, A.; Boehm, G.; Weis, S.; Clauss, J. Biodegradable Polymers Based on Natural and Renewable Raw Materials Especially Isosorbite. U.S. Patent 6,342,300 B1, 29 January 2002. [Google Scholar]
  74. van Berkel, J.G.; Guigo, N.; Kolstad, J.J.; Sipos, L.; Wang, B.; Dam, M.A.; Sbirrazzuoli, N. Isothermal crystallization kinetics of poly (ethylene 2,5-furandicarboxylate). Macromol. Mater. Eng. 2015, 300, 466–474. [Google Scholar] [CrossRef]
  75. Ma, J.P.; Pang, Y.; Wang, M.; Xu, J.; Ma, H.; Nie, X. The copolymerization reactivity of diols with 2,5-furandicarboxylic acid for furan-based copolyester materials. J. Mater. Chem. 2012, 22, 3457–3461. [Google Scholar] [CrossRef]
  76. Hong, S.; Min, K.D.; Nam, B.U.; Park, O.O. High molecular weight bio furan-based co-polyesters for food packaging applications: Synthesis, characterization and solid-state polymerization. Green Chem. 2016, 18, 5142–5150. [Google Scholar] [CrossRef]
  77. Wang, J.G.; Liu, X.Q.; Zhang, Y.J.; Liu, F.; Zhu, J. Modification of poly(ethylene 2,5-furandicarboxylate) with 1,4-cyclohexanedimethylene: Influence of composition on mechanical and barrier properties. Polymer 2016, 103, 1–8. [Google Scholar] [CrossRef]
  78. Wang, X.S.; Wang, Q.Y.; Liu, S.Y.; Wang, G.Y. Biobased copolyesters: Synthesis, structure, thermal and mechanical properties of poly(ethylene 2,5-furandicarboxylate-co-ethylene 1,4-cyclohexanedicarboxylate. Polym. Degrad. Stab. 2018, 154, 96–102. [Google Scholar] [CrossRef]
  79. Diao, L.C.; Su, K.M.; Li, Z.H.; Ding, C.K. Furan-based co-polyesters with enhanced thermal properties: Poly(1,4-butylene-co-1,4-cyclohexanedimethylene-2,5-furandicarboxylic acid). Rsc Adv. 2016, 6, 27632–27639. [Google Scholar] [CrossRef]
  80. Jia, Z.; Wang, J.G.; Sun, L.Y.; Liu, F.; Zhu, J.; Liu, X.Q. Copolyesters developed from bio-based 2,5-furandicarboxylic acid: Synthesis, sequence distribution, mechanical, and barrier properties of poly(propylene-co-1,4-cyclohexanedimethylene 2,5-furandicarboxylate)s. J. Appl. Polym. Sci. 2019, 136, 47291. [Google Scholar] [CrossRef]
  81. Kim, T.; Koo, J.M.; Ryu, M.H.; Jeon, H.; Kim, S.M.; Park, S.A.; Oh, D.X.; Park, J.; Hwang, S.Y. Sustainable terpolyester of high T-g based on bio heterocyclic monomer of dimethyl furan-2,5-dicarboxylate and isosorbide. Polymer 2017, 132, 122–132. [Google Scholar] [CrossRef]
  82. Wang, J.; Liu, X.; Zhu, J.; Jiang, Y. Copolyesters based on 2,5-furandicarboxylic acid (FDCA): Effect of 2,2,4,4-tetramethyl-1,3-cyclobutanediol units on their properties. Polymers 2017, 9, 305. [Google Scholar] [CrossRef] [Green Version]
  83. Wang, G.Q.; Jiang, M.; Zhang, Q.; Wang, R.; Zhou, G.Y. Biobased multiblock copolymers: Synthesis, properties and shape memory performance of poly(ethylene 2,5-furandicarboxylate)-b-ly(ethylene glycol). Polym. Degrad. Stab. 2017, 144, 121–127. [Google Scholar] [CrossRef]
  84. Zhang, H.; Jiang, M.; Wu, Y.; Li, L.; Wang, Z.; Wang, R.; Zhou, G. Development of completely furfural-based renewable polyesters with controllable properties. Green Chem. 2021, 23, 2437–2448. [Google Scholar] [CrossRef]
  85. Pang, C.; Jiang, X.; Yu, Y.; Chen, L.; Ma, J.; Gao, H. Copolymerization of natural camphor-derived rigid diol with various dicarboxylic acids: Access to biobased polyesters with various properties. ACS Macro Lett. 2019, 8, 1442–1448. [Google Scholar] [CrossRef]
  86. Shen, A.; Wang, J.; Ma, S.; Fei, X.; Zhang, X.; Zhu, J.; Liu, X. Completely amorphous high thermal resistant copolyesters from bio-based 2,5-furandicarboxylic acid. J. Appl. Polym. Sci. 2021, 138, 50627. [Google Scholar] [CrossRef]
  87. Witt, U.; Einig, T.; Yamamoto, M.; Kleeberg, I.; Deckwer, W.D.; Muller, R.J. Biodegradation of aliphatic-aromatic copolyesters: Evaluation of the final biodegradability and ecotoxicological impact of degradation intermediates. Chemosphere 2001, 44, 289–299. [Google Scholar] [CrossRef]
  88. Shah, A.A.; Eguchi, T.; Mayumi, D.; Kato, S.; Shintani, N.; Kamini, N.R.; Nakajima-Kambe, T. Purification and properties of novel aliphatic-aromatic co-polyesters degrading enzymes from newly isolated Roseateles depolymerans strain TB-87. Polym. Degrad. Stab. 2013, 98, 609–618. [Google Scholar] [CrossRef]
  89. Nakajimakambe, T.; Toyoshima, K.; Saito, C.; Takaguchi, H.; Akutsushigeno, Y.; Sato, M.; Miyama, K.; Nomura, N.; Uchiyama, H. Rapid monomerization of poly(butylene succinate)-co-(butylene adipate) by Leptothrix sp. J. Biosci. Bioeng. 2009, 108, 513–516. [Google Scholar] [CrossRef] [PubMed]
  90. Kasuya, K.; Ishii, N.; Inoue, Y.; Yazawa, K.; Tagaya, T.; Yotsumoto, T.; Kazahaya, J.; Nagai, D. Characterization of a mesophilic aliphatic–aromatic copolyester-degrading fungus. Polym. Degrad. Stab. 2009, 94, 1190–1196. [Google Scholar] [CrossRef]
  91. Wu, D.; Archana, S.; Rajiv, S.; Minna, H. Nano-graphene oxide functionalized bioactive poly(lactic acid) and poly(ε-caprolactone) nanofibrous scaffolds. Materials 2018, 11, 566. [Google Scholar] [CrossRef] [Green Version]
  92. Kashif, M.; Yun, B.M.; Lee, K.S.; Chang, Y.W. Biodegradable shape-memory poly(ε-caprolactone)/polyhedral oligomeric silsequioxane nanocomposites: Sustained drug release and hydrolytic degradation. Mater. Lett. 2016, 166, 125–128. [Google Scholar] [CrossRef]
  93. Woodruff, M.A.; Hutmacher, D.W. The return of a forgotten polymer—Polycaprolactone in the 21st century. Prog. Polym. Sci. 2010, 35, 1217–1256. [Google Scholar] [CrossRef] [Green Version]
  94. Wang, X.S.; Liu, S.Y.; Wang, Q.Y.; Li, J.G.; Wang, G.Y. Synthesis and characterization of poly(ethylene 2,5-furandicarboxylate-co-epsilon-ecaprolactone) copolyesters. Eur. Polym. J. 2018, 109, 191–197. [Google Scholar] [CrossRef]
  95. Kasmi, N.; Wahbi, M.; Papadopoulos, L.; Terzopoulou, Z.; Guigo, N.; Sbirrazzuoli, N.; Papageorgiou, G.Z.; Bikiaris, D.N. Synthesis and characterization of two new biobased poly(pentylene 2,5-furandicarboxylate-co-caprolactone) and poly(hexamethylene 2,5-furandicarboxylate-co-caprolactone) copolyesters with enhanced enzymatic hydrolysis properties. Polym. Degrad. Stab. 2019, 160, 242–263. [Google Scholar] [CrossRef]
  96. Zheng, M.Y.; Zang, X.L.; Wang, G.X.; Wang, P.L.; Lu, B.; Ji, J.H. Poly(butylene 2,5-furandicarboxylate-epsilon-caprolactone): A new bio-based elastomer with high strength and biodegradability. Express Polym. Lett. 2017, 11, 611–621. [Google Scholar] [CrossRef]
  97. Morales-Huerta, J.C.; de Ilarduya, A.M.; Munoz-Guerra, S. Blocky poly(-caprolactone-co-butylene 2,5-furandicarboxylate) copolyesters via enzymatic ring opening polymerization. J. Polym. Sci. Polym. Chem. 2018, 56, 290–299. [Google Scholar] [CrossRef]
  98. Hu, H.; Zhang, R.; Ying, W.B.; Kong, Z.; Wang, K.; Wang, J.; Zhu, J. Biodegradable elastomer from 2,5-furandicarboxylic acid and ε-caprolactone: Effect of crystallization on elasticity. ACS Sustain. Chem. Eng. 2019, 7, 17778–17788. [Google Scholar] [CrossRef]
  99. Li, F.X.; Xu, X.J.; Hao, Q.H.; Li, Q.B.; Yu, J.Y.; Cao, A.M. Effects of comonomer sequential structure on thermal and crystallization behaviors of biodegradable poly(butylene succinate-co-butylene terephthalate)s. J. Polym. Sci. B Polym. Phys. 2006, 44, 1635–1644. [Google Scholar] [CrossRef]
  100. Sun, Y.J.; Wu, L.B.; Bu, Z.Y.; Li, B.G.; Li, N.X.; Dai, J.M. Synthesis and thermomechanical and rheological properties of biodegradable long-chain branched poly(butylene succinate-co-butylene terephthalate) copolyesters. Ind. Eng. Chem. Res. 2014, 53, 10380–10386. [Google Scholar] [CrossRef]
  101. Luo, S.L.; Li, F.X.; Yu, J.Y. The thermal, mechanical and viscoelastic properties of poly(butylene succinate-co-terephthalate) (PBST) copolyesters with high content of BT units. J. Polym. Res. 2011, 18, 393–400. [Google Scholar] [CrossRef]
  102. Moraleshuerta, J.C.; Ciulik, C.; De Ilarduya, A.M.; Munozguerra, S. Fully bio-based aromatic–aliphatic copolyesters: Poly(butylene furandicarboxylate-co-succinate)s obtained by ring opening polymerization. Polym. Chem. 2017, 8, 748–760. [Google Scholar] [CrossRef] [Green Version]
  103. Lomeli-Rodriguez, M.; Martin-Molina, M.; Jimenez-Pardo, M.; Nasim-Afzal, Z.; Cauet, S.I.; Davies, T.E.; Rivera-Toledo, M.; Lopez-Sanchez, J.A. Synthesis and kinetic modeling of biomass-derived renewable polyesters. J. Polym. Sci. Polym. Chem. 2016, 54, 2876–2887. [Google Scholar] [CrossRef]
  104. Wu, L.; Mincheva, R.; Xu, Y.; Raquez, J.M.; Dubois, P. High molecular weight poly(butylene succinate-co-butylene furandicarboxylate) copolyesters: From catalyzed polycondensation reaction to thermomechanical properties. Biomacromolecules 2012, 13, 2973–2981. [Google Scholar] [CrossRef]
  105. Peng, S.; Wu, L.; Li, B.G.; Dubois, P. Hydrolytic and compost degradation of biobased PBSF and PBAF copolyesters with 40–60 mol% BF unit. Polym. Degrad. Stab. 2017, 146, 223–228. [Google Scholar] [CrossRef]
  106. Peng, S.B.; Wu, B.S.; Wu, L.B.; Li, B.G.; Dubois, P. Hydrolytic degradation of biobased poly(butylene succinate-co-furandicarboxylate) and poly(butylene adipate-co-furandicarboxylate) copolyesters under mild conditions. J. Appl. Polym. Sci. 2017, 134, 44674. [Google Scholar] [CrossRef]
  107. Hua, H.; Zhang, R.Y.; Wang, J.G.; Ying, W.B.; Zhu, J. Fully bio-based poly(propylene succinate-co-propylene furandicarboxylate) copolyesters with proper mechanical, degradation and barrier properties for green packaging applications. Eur. Polym. J. 2018, 102, 101–110. [Google Scholar] [CrossRef]
  108. Witt, U.; Yamamoto, M.; Seeliger, U.; Müller, R.J.; Warzelhan, V. Biodegradable polymeric materials—not the origin but the chemical structure determines biodegradability. Angew. Chem. Int. Ed. 1999, 38, 1438–1442. [Google Scholar] [CrossRef]
  109. Witt, U.; Muller, R.J.; Deckwer, W.D. Biodegradation of polyester copolymers containing aromatic-compounds. J. Macromol. 1995, A32, 851–856. [Google Scholar] [CrossRef]
  110. Papadopoulos, L.; Magaziotis, A.; Nerantzaki, M.; Terzopoulou, Z.; Papageorgiou, G.Z.; Bikiaris, D.N. Synthesis and characterization of novel poly(ethylene furanoate-co-adipate) random copolyesters with enhanced biodegradability. Polym. Degrad. Stab. 2018, 156, 32–42. [Google Scholar] [CrossRef]
  111. Zhou, W.D.; Wang, X.W.; Yang, B.; Xu, Y.; Zhang, W.; Zhang, Y.J.; Ji, J.H. Synthesis, physical properties and enzymatic degradation of bio-based poly(butylene adipate-co-butylene furandicarboxylate) copolyesters. Polym. Degrad. Stab. 2013, 98, 2177–2183. [Google Scholar] [CrossRef]
  112. Wu, B.S.; Xu, Y.T.; Bu, Z.Y.; Wu, L.B.; Li, B.G.; Dubois, P. Biobased poly(butylene 2,5-furandicarboxylate) and poly(butylene adipate-co-butylene 2,5-furandicarboxylate)s: From synthesis using highly purified 2,5-furandicarboxylic acid to thermo-mechanical properties. Polymer 2014, 55, 3648–3655. [Google Scholar] [CrossRef]
  113. Hu, H.; Zhang, R.Y.; Wang, J.G.; Ying, W.B.; Shi, L.; Yao, C.K.; Kong, Z.Y.; Wang, K.; Zhu, J. A mild method to prepare high molecular weight poly(butylene furandicarboxylate-co-glycolate) copolyesters: Effects of the glycolate content on thermal, mechanical, and barrier properties and biodegradability. Green Chem. 2019, 21, 3013–3022. [Google Scholar] [CrossRef]
  114. Soccio, M.; Costa, M.; Lotti, N.; Gazzano, M.; Siracusa, V.; Salatelli, E.; Manaresi, P.; Munari, A. Novel fully biobased poly(butylene 2,5-furanoate/diglycolate) copolymers containing ether linkages: Structure-property relationships. Eur. Polym. J. 2016, 81, 397–412. [Google Scholar] [CrossRef]
  115. Wu, H.L.; Wen, B.B.; Zhou, H.; Zhou, J.D.; Yu, Z.L.; Cui, L.Y.; Huang, T.; Cao, F. Synthesis and degradability of copolyesters of 2,5-furandicarboxylic acid, lactic acid, and ethylene glycol. Polym. Degrad. Stab. 2015, 121, 100–104. [Google Scholar] [CrossRef]
  116. Jia, Z.; Wang, J.G.; Sun, L.Y.; Zhu, J.; Liu, X.Q. Fully bio-based polyesters derived from 2,5-furandicarboxylic acid (2,5-FDCA) and dodecanedioic acid (DDCA): From semicrystalline thermoplastic to amorphous elastomer. J. Appl. Polym. Sci. 2018, 135, 46076. [Google Scholar] [CrossRef]
  117. Soares, M.J.; Dannecker, P.K.; Vilela, C.; Bastos, J.; Meier, M.A.R.; Sousa, A.F. Poly(1,20-eicosanediyl 2,5-furandicarboxylate), a biodegradable polyester from renewable resources. Eur. Polym. J. 2017, 90, 301–311. [Google Scholar] [CrossRef]
  118. Weinberger, S.; Canadell, J.; Quartinello, F.; Yeniad, B.; Arias, A.; Pellis, A.; Guebitz, G.M. Enzymatic degradation of poly(ethylene 2,5-furanoate) powders and amorphous films. Catalysts 2017, 7, 318. [Google Scholar] [CrossRef] [Green Version]
  119. Zhang, H.; Jiang, M.; Wu, Y.; Li, L.; Wang, Z.; Wang, R.; Zhou, G. Development of high-molecular-weight fully renewable biopolyesters based on oxabicyclic diacid and 2,5-furandicarboxylic acid: Promising as packaging and medical materials. ACS Sustain. Chem. Eng. 2021, 9, 6799–6809. [Google Scholar] [CrossRef]
  120. Kim, H.; Kim, T.; Choi, S.; Jeon, H.; Oh, D.X.; Park, J.; Eom, Y.; Hwang, S.Y.; Koo, J.M. Remarkable elasticity and enzymatic degradation of bio-based poly(butylene adipate-co-furanoate): Replacing terephthalate. Green Chem. 2020, 22, 7778–7787. [Google Scholar] [CrossRef]
  121. Paul, D.R.; Robeson, L.M. Polymer nanotechnology: Nanocomposites. Polymer 2008, 49, 3187–3204. [Google Scholar] [CrossRef] [Green Version]
  122. Huang, Z.M.; Zhang, Y.Z.; Kotaki, M.; Ramakrishna, S. A review on polymer nanofibers by electrospinning and their applications in nanocomposites. Compos. Sci. Technol. 2003, 63, 2223–2253. [Google Scholar] [CrossRef]
  123. Litchfield, D.W.; Baird, D.G. The role of nanoclay in the generation of poly(ethylene terephthalate) fibers with improved modulus and tenacity. Polymer 2008, 49, 5027–5036. [Google Scholar] [CrossRef]
  124. Kim, H.; Abdala, A.A.; Macosko, C.W. Graphene–polymer nanocomposites. Macromolecules 2010, 43, 6515–6530. [Google Scholar] [CrossRef]
  125. Ray, S.S.; Okamoto, M. Polymer/layered silicate nanocomposites: A review from preparation to processing. Prog. Polym. Sci. 2003, 28, 1539–1641. [Google Scholar]
  126. Moniruzzaman, M.; Winey, K.I. Polymer nanocomposites containing carbon nanotubes. Macromolecules 2006, 39, 5194–5205. [Google Scholar] [CrossRef]
  127. Lotti, N.; Munari, A.; Gigli, M.; Gazzano, M.; Tsanaktsis, V.; Bikiaris, D.N.; Papageorgiou, G.Z. Thermal and structural response of in situ prepared biobased poly(ethylene 2,5-furan dicarboxylate) nanocomposites. Polymer 2016, 103, 288–298. [Google Scholar] [CrossRef]
  128. Terzopoulou, Z.; Tarani, E.; Kasmi, N.; Papadopoulos, L.; Chrissafis, K.; Papageorgiou, D.G.; Papageorgiou, G.Z.; Bikiaris, D.N. Thermal decomposition kinetics and mechanism of in-situ prepared bio-based poly(propylene 2,5-furan dicarboxylate)/graphene nanocomposites. Molecules 2019, 24, 1717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Paszkiewicz, S.; Janowska, I.; Pawlikowska, D.; Szymczyk, A.; Irska, I.; Lisiecki, S.; Stanik, R.; Gude, M.; Piesowicz, E. New functional nanocomposites based on poly(trimethylene 2,5-furanoate) and few layer graphene prepared by in situ polymerization. Express Polym. Lett. 2018, 12, 530–542. [Google Scholar] [CrossRef]
  130. Achilias, D.S.; Chondroyiannis, A.; Nerantzaki, M.; Adam, K.V.; Terzopoulou, Z.; Papageorgiou, G.Z.; Bikiaris, D.N. Solid state polymerization of poly(ethylene furanoate) and its nanocomposites with SiO2 and TiO2. Macromol. Mater. Eng. 2017, 302, 1700012. [Google Scholar] [CrossRef]
  131. Zhou, G.; Li, L.; Jiang, M.; Wang, G.; Wang, R.; Wu, G.; Zhou, G. Renewable poly(butene 2, 5-furan dicarboxylate) nanocomposites constructed by TiO2 nanocubes: Synthesis, crystallization, and properties. Polym. Degrad. Stab. 2021, 189, 109591. [Google Scholar] [CrossRef]
  132. Pavlidou, S.; Papaspyrides, C.D. A review on polymerelayered silicate nanocomposites. Prog. Polym. Sci. 2008, 33, 1119–1198. [Google Scholar] [CrossRef]
  133. Stoeffler, K.; Lafleur, P.G.; Denault, J. Thermal decomposition of various alkyl onium organoclays: Effect on polyethylene terephthalate nanocomposites’ properties. Polym. Degrad. Stab. 2008, 93, 1332–1350. [Google Scholar] [CrossRef]
  134. Yuan, X.; Li, C.; Guan, G.; Xiao, Y.; Zhang, D. Thermal degradation investigation of poly(ethylene terephthalate)/fibrous silicate nanocomposites. Polym. Degrad. Stab. 2008, 93, 466–475. [Google Scholar] [CrossRef]
  135. Martino, L.; Niknam, V.; Guigo, N.; van Berkel, J.G.; Sbirrazzuoli, N. Morphology and thermal properties of novel clay-based poly(ethylene 2,5-furandicarboxylate) (PEF) nanocomposites. Rsc Adv. 2016, 6, 59800–59807. [Google Scholar] [CrossRef]
  136. Papadopoulos, L.; Terzopoulou, Z.; Bikiaris, D.N.; Patsiaoura, D.; Chrissafis, K.; Papageorgiou, D.G.; Papageorgiou, G.Z. Synthesis and characterization of in-situ-prepared nanocomposites based on poly(propylene 2,5-furan dicarboxylate) and aluminosilicate clays. Polymers 2018, 10, 937. [Google Scholar] [CrossRef] [Green Version]
  137. Codou, A.; Guigo, N.; van Berkel, J.G.; de Jong, E.; Sbirrazzuoli, N. Preparation and characterization of poly(ethylene 2,5-furandicarboxylate/nanocrystalline cellulose composites via solvent casting. J. Polym. Eng. 2017, 37, 869–878. [Google Scholar] [CrossRef]
  138. Nair, S.S.; Zhu, J.Y.; Deng, Y.; Ragauskas, A.J. High performance green barriers based on nanocellulose. Sustain. Chem. Processes 2014, 2, 1–7. [Google Scholar] [CrossRef] [Green Version]
  139. Abdul Khalil, H.; Bhat, Y.; Ireana, A.F. Green composites from sustainable cellulose nanofibrils: A review. Carbohydr. Polym. 2012, 87, 963–979. [Google Scholar] [CrossRef]
  140. Heux, L.; Chauve, G.; Bonini, C. Nonflocculating and chiral-nematic self-ordering of cellulose microcrystals suspensions in nonpolar solvents. Langmuir 2000, 16, 8210–8212. [Google Scholar] [CrossRef]
  141. Codou, A.; Guigo, N.; van Berkel, J.G.; de Jong, E.; Sbirrazzuoli, N. Preparation and crystallization behavior of poly(ethylene 2,5-furandicarboxylate)/cellulose composites by twin screw extrusion. Carbohydr. Polym. 2017, 174, 1026–1033. [Google Scholar] [CrossRef]
  142. Matos, M.; F. Sousa, A.; H.C.S. Silva, N.; S.R. Freire, C.; Andrade, M.; Mendes, A.; J.D. Silvestre, A. Furanoate-based nanocomposites: A case study using poly(butylene 2,5-furanoate) and poly(butylene 2,5-furanoate)-co-(butylene diglycolate) and bacterial cellulose. Polymers 2018, 10, 810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Martino, L.; Guigo, N.; Berkel, J.G.V.; Sbirrazzuoli, N. Influence of organically modified montmorillonite and sepiolite clayson the physical properties of bio-based poly(ethylene 2,5-furandicarboxylate). Compos. Part B-Eng. 2016, 110, 96–105. [Google Scholar] [CrossRef]
  144. Safapour, S.; Seyed-Esfahani, M.; Auriemma, F.; de Ballesteros, O.R.; Vollaro, P.; Di Girolamo, R.; De Rosa, C.; Khosroshahi, A. Reactive blending as a tool for obtaining poly(ethylene terephthalate)-based engineering materials with tailored properties. Polymer 2010, 51, 4340–4350. [Google Scholar] [CrossRef]
  145. Poulopoulou, N.; Kasmi, N.; Bikiaris, D.N.; Papageorgiou, D.G.; Floudas, G.; Papageorgiou, G.Z. Sustainable polymers from renewable resources: Polymer blends of furan-based polyesters. Macromol. Mater. Eng. 2018, 303, 1800153. [Google Scholar] [CrossRef]
  146. Poulopoulou, N.; Kantoutsis, G.; Bikiaris, D.N.; Achilias, D.S.; Kapnisti, M.; Papageorgiou, G.Z. Biobased engineering thermoplastics: Poly(butylene 2,5-furandicarboxylate) blends. Polymers 2019, 11, 937. [Google Scholar] [CrossRef] [Green Version]
  147. Poulopoulou, N.; Pipertzis, A.; Kasmi, N.; Bikiaris, D.N.; Papageorgiou, D.G.; Floudas, G.; Papageorgiou, G.Z. Green polymeric materials: On the dynamic homogeneity and miscibility of furan-based polyester blends. Polymer 2019, 174, 187–199. [Google Scholar] [CrossRef]
  148. Poulopoulou, N.; Smyrnioti, D.; Nikolaidis, G.N.; Tsitsimaka, I.; Christodoulou, E.; Bikiaris, D.N.; Charitopoulou, M.A.; Achilias, D.S.; Kapnisti, M.; Papageorgiou, G.Z. Sustainable plastics from biomass: Blends of polyesters based on 2,5-furandicarboxylic acid. Polymers 2020, 12, 225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Poulopoulou, N.; Kasmi, N.; Siampani, M.; Terzopoulou, Z.N.; Bikiaris, D.N.; Achilias, D.S.; Papageorgiou, D.G.; Papageorgiou, G.Z. Exploring next-generation engineering bioplastics: Poly(alkylene furanoate)/poly(alkylene terephthalate) (PAF/PAT) blends. Polymers 2019, 11, 556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Long, Y.; Zhang, R.; Huang, J.; Wang, J.; Jiang, Y.; Hu, G.H.; Yang, J.; Zhu, J. Tensile property balanced and gas barrier improved PLA by blending with bio-based poly(butylene 2,5-furan dicarboxylate). ACS Sustain. Chem. Eng. 2017, 5, 9244–9253. [Google Scholar] [CrossRef]
  151. Long, Y.; Zhang, R.Y.; Huang, J.C.; Wang, J.G.; Zhang, J.W.; Rayand, N.; Hu, G.H.; Yang, J.; Zhu, J. Retroreflection in binary bio-based PLA/PBF blends. Polymer 2017, 125, 138–143. [Google Scholar] [CrossRef]
  152. Cai, Q.Q.; Bai, T.W.; Zhang, H.J.; Yao, X.X.; Ling, J.; Zhu, W.P. Catalyst-free synthesis of polyesters via conventional melt polycondensation. Mater. Today 2021, 51, 155–164. [Google Scholar] [CrossRef]
  153. Maaskant, E.; van Es, D.S. Unexpected susceptibility of poly(ethylene furanoate) to UV irradiation: A warning light for furandicarboxylic acid? ACS Macro Lett. 2021, 10, 1616–1621. [Google Scholar] [CrossRef]
Figure 1. Different polyesters derived from 2,5-FDCA.
Figure 1. Different polyesters derived from 2,5-FDCA.
Polymers 14 00625 g001
Scheme 1. Synthesis of FDCA-based polyesters from melt polymerization, solution polymerization, and ROP.
Scheme 1. Synthesis of FDCA-based polyesters from melt polymerization, solution polymerization, and ROP.
Polymers 14 00625 sch001
Figure 2. The history of PEF and synthesis of other furanic polyesters.
Figure 2. The history of PEF and synthesis of other furanic polyesters.
Polymers 14 00625 g002
Figure 3. (i) 70 μm films of PEF (a) TBT 400 ppm; (b) ZnAcO 400 ppm; (c) Al(acac) 350 ppm [35]; (ii) Appearance comparison of PET catalyzed by metal Zn and PEFs catalyzed by metal Zn, TBT, and Fe [36].
Figure 3. (i) 70 μm films of PEF (a) TBT 400 ppm; (b) ZnAcO 400 ppm; (c) Al(acac) 350 ppm [35]; (ii) Appearance comparison of PET catalyzed by metal Zn and PEFs catalyzed by metal Zn, TBT, and Fe [36].
Polymers 14 00625 g003
Figure 4. (a) Views of the PBF structure [21]; (b) PLM images of spherulites of (a) PPF grown at 130 °C, (b) PPF grown at 135 °C, (c) PPF grown at 145 °C, (d) PPN grown at 165 °C, (e) PPN grown at 175 °C, (f) PPN grown at 190 °C, (g) PPT grown at 208 °C, (h) PPT grown at 212 °C, (i) PPT grown at 214 °C with a scale bar of 50 μm [41].
Figure 4. (a) Views of the PBF structure [21]; (b) PLM images of spherulites of (a) PPF grown at 130 °C, (b) PPF grown at 135 °C, (c) PPF grown at 145 °C, (d) PPN grown at 165 °C, (e) PPN grown at 175 °C, (f) PPN grown at 190 °C, (g) PPT grown at 208 °C, (h) PPT grown at 212 °C, (i) PPT grown at 214 °C with a scale bar of 50 μm [41].
Polymers 14 00625 g004
Scheme 2. Structure of rigid diols in polyester synthesis.
Scheme 2. Structure of rigid diols in polyester synthesis.
Polymers 14 00625 sch002
Table 1. Thermal and mechanical properties of furan-polyesters with aliphatic diols via different synthetic methods.
Table 1. Thermal and mechanical properties of furan-polyesters with aliphatic diols via different synthetic methods.
PolyesterSynthetic MethodTg
/°C
Tm
/°C
Tdmax
/°C
Tensile Modulus (GPa)Tensile Strength (MPa)Mn (103 g/mol)Ref.
PEFTransesterification77–802143982.453522.4–83[10,33]
PEFDirect esterification80–89210.44072.166.7105.3[11,35]
PEFROP80215----[32]
PPFDirect esterification53–57.91803961.668.213.9–60.2[11,41]
PBFDirect esterification30.5–40171392–4280.7–1.15.5–32.98.0–23.2[11,21]
PHFDirect esterification28.1148.23890.535.532.1[11]
POFDirect esterification21.8148.63910.320.320.7[11]
POFTransesterification−5140-0.326.534.6[42]
Poly(nonylene 2,5-furanoate) (PNF)Transesterification−3069-0.22140.0[42]
Poly(decylene 2,5-furanoate) (PDeF)Transesterification−8116-0.21136.7[42]
Poly(dodecylene 2,5-furanoate) (PDoF)Transesterification−22111-0.21139.4[42]
Table 2. Thermal and mechanical properties of furan-polyesters with rigid diols via different synthetic methods.
Table 2. Thermal and mechanical properties of furan-polyesters with rigid diols via different synthetic methods.
PolyesterSynthetic MethodTg
/°C
Tm
/°C
Td5%
/°C
Tdmax
/°C
Tensile Modulus (MPa)Tensile Strength (MPa)Mn (103 g/mol)Ref.
PDASFSolution polycondensation180--450--13.8[10]
PDAIFSolution polycondensation140--396--5.7[10]
PIsFTransesterification157--421---[58]
PDIsFsTransesterification−1–2164–111405–413439–44414–5590.7–2011.5–25.4[59]
PBIFTransesterification45–105160339–357365–384--24.4–31.0[60]
PISBFTransesterification55–151-370–376405–4171470639.3–19.1[61]
PCIsFsTransesterification75–103220–257363–375402–409---[62]
PIsITransesterification73-274310/382--2.6[63]
PBISDirect esterification−28–−1189–109 64–8717–2345.7–53.5[64]
PXIIFTransesterification/SSPC94250 411--30.3[65]
PCFTransesterification71–87219–291 377–4031.69–1.8244–528.3–26.2[23,66]
PCdFTransesterification175- 380--6.5[66]
Table 3. Thermal and mechanical properties of furan co-polyesters based on PEF.
Table 3. Thermal and mechanical properties of furan co-polyesters based on PEF.
PolyesterSynthetic MethodTg
/°C
Tm
/°C
Td5%
/°C
Tdmax
/°C
Tensile Modulus (GPa)Tensile Strength (MPa)Mn (103 g/mol)Ref.
PECFTransesterification80–85206–225365–404393–4172.3–3.260–7227.6–32.0[76,77]
PPCFTransesterification59–73193–232378–382412–4211.9–2.179–8835–41[80]
PETFTransesterification90.9–91.1-368–369400–4033.1–3.397–98(Mv = 51–56)[82]
PEFCDirect esterification32.3–76.1-365–390394–4210.8–1.818–4918.8–22.7[78]
PEFEGDirect esterification78.5–84.9177–209345–360--11–2734.0–46.8[83]
PEICFDirect esterification90–119-----16.4–19.6[81]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fei, X.; Wang, J.; Zhang, X.; Jia, Z.; Jiang, Y.; Liu, X. Recent Progress on Bio-Based Polyesters Derived from 2,5-Furandicarbonxylic Acid (FDCA). Polymers 2022, 14, 625. https://doi.org/10.3390/polym14030625

AMA Style

Fei X, Wang J, Zhang X, Jia Z, Jiang Y, Liu X. Recent Progress on Bio-Based Polyesters Derived from 2,5-Furandicarbonxylic Acid (FDCA). Polymers. 2022; 14(3):625. https://doi.org/10.3390/polym14030625

Chicago/Turabian Style

Fei, Xuan, Jinggang Wang, Xiaoqin Zhang, Zhen Jia, Yanhua Jiang, and Xiaoqing Liu. 2022. "Recent Progress on Bio-Based Polyesters Derived from 2,5-Furandicarbonxylic Acid (FDCA)" Polymers 14, no. 3: 625. https://doi.org/10.3390/polym14030625

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