You are on page 1of 22

European Polymer Journal 160 (2021) 110778

Contents lists available at ScienceDirect

European Polymer Journal


journal homepage: www.elsevier.com/locate/europolj

Biobased 2,5-furandicarboxylic acid (FDCA) and its emerging copolyesters’


properties for packaging applications
Surabhi Pandey a, Marie-Josée Dumont a, *, Valérie Orsat a, Denis Rodrigue b
a
Bioresource Engineering Department, McGill University, 21111 Lakeshore Rd., Ste-Anne-de-Bellevue, QC H9X 3V9, Canada
b
Chemical Engineering Department, Department of Chemical Engineering, Université Laval, 1065 Avenue de la Médecine, Québec City, Québec G1V 0A6, Canada

A R T I C L E I N F O A B S T R A C T

Keywords: The synthesis of 2,5-furandicarboxylic acid (FDCA) based polymers is gaining importance as a replacement for
2,5-furandicarboxylic acid petrochemical polymers due to their excellent thermomechanical and barrier properties. However, the industrial
Food and beverage packaging production of FDCA based homopolyesters is impeded due to several factors including poor optical property, low
Biodegradability
ductility, and slow crystallization rate. In order to tune the properties of FDCA-based homopolyesters, FDCA has
Poly(butylene adipate-co-furandicarboxylate)
Thermomechanical properties
been copolymerized with several aliphatic (adipic acid, succinic acid, lactic acid, sebacic acid, polyglycolic acid,
and polyethylene glycol) and cyclic (caprolactam and isohexides) comonomers. The present review focuses on
the current trends related to the synthesis of FDCA and its copolyesters. The thermomechanical and barrier
properties of copolyesters have been discussed concerning applications in the food and beverage packaging
sectors, with an emphasis on their biodegradability (hydrolytic and enzymatic) and compostability.

1. Introduction sustainable approach to produce polymeric materials [1,2]. These bio-


based polymers can be categorized into three categories namely: (a)
The replacement of fossil fuels with biomass has been regarded as a naturally occurring, (b) chemically synthesized, and (c) polymers

Abbreviations: εb, elongation at break; σb, strength at break; Ƞ, intrinsic viscosity; BDA, 1,4-butanediamine; BIFp, barrier improvement factor; CALB, Candida
antarctica lipase B; CBDO, 2,2,4,4-tetramethyl-1,3-cyclobutanediol; CHDM, 1,4-cyclohexanedimethanol; DBTO, dibutyltin(IV) oxide; DDCA, dodecanedioic acid;
DMFD, dimethyl-2,5-furandicarboxylate; E, Young’s modulus; E′ , storage modulus; FDCA, 2,5-furandicarboxylate; HT, hydrotalcite; ΔHm, melting enthalpy in the
first heating scan; La(acac)3, lanthanum(III) acetylacetonate; Mn, number average molecular weight; Mw, weight average molecular weight; MSP, minimum selling
price; PBA, poly(butylene adipate); PBA-co-PBF, poly(butylene adipate-co-furandicarboxylate); PBA-co-PBT, poly(butylene adipate-co-terephthalate); PBF, poly
(butylene furandicarboxylate); PBF-co-PEG, poly(ethylene glycol-co-butylene furandicarboxylate); PBS, poly(butylene succinate); PBS-co-PBA, poly(butylene suc­
cinate-co-adipate); PBS-co-PBF, poly(butylene succinate-co-furandicarboxylate); PBS-co-PBT, poly(butylene succinate-co-terephthalate); PBSe-co-PBF, poly(butylene
sebacate-co-furandicarboxylate); PCF, poly(1,4-cyclohexanedimethylene furandicarboxylate); PCL, poly(ε-caprolactone); PCL-co-PBF, poly(ε-caprolactone-co-
butylene furandicarboxylate); PCL-co-PHeF, poly(ε-caprolactone-co-hexamethylene furandicarboxylate); PCL-co-PPeF, poly(ε-caprolactone-co-pentylene fur­
andicarboxylate); PDeF, poly(decylene furandicarboxylate); PDG-co-PBF, poly(diglycolate-co-butylene furandicarboxylate); PDI, polymer dispersity index; PDIsF,
poly(decamethylene-co-isosorbide furandicarboxylate); PDoF, poly(dodecylene furandicarboxylate); PE, polyethylene; PEA-co-PEF, poly(ethylene adipate-co-fur­
andicarboxylate); PEBF, poly(2-ethyl-2-butyl-1,3-propylene furandicarboxylate); PEDF, poly(ethylene dodecanedioate-2,5-furandicarboxylate); PEF, poly(ethylene
furandicarboxylate); PESe, poly(ethylene sebacate); PESe-co-PEF, poly(ethylene sebacate-co-furandicarboxylate); PET, poly(ethylene terephthalate); PHA, poly
(hydroxyalkanoates); PHB, poly(hydroxybutyrate); PHeF, poly(hexamethylene furandicarboxylate); PHepF, poly(heptylene furandicarboxylate); PIsBF, poly(iso­
sorbide-co-butylene furandicarboxylate); PIsCBF, poly(isosorbide carbonate-co-butylene furandicarboxylate); PIs(CL)F, poly(ε-caprolactone-co-isosorbide fur­
andicarboxylate); PIsF, poly(isosorbide furandicarboxylate); PLA, poly(lactic acid); PLA-co-PBF, poly(lactic acid-co-butylene furandicarboxylate); PLA-co-PEF, poly
(lactic acid-co-ethylene furandicarboxylate); PNF, poly(nonylene furandicarboxylate); PNgF, poly(neopentyl glycol furandicarboxylate; PNS-co-PNF, poly(neopentyl
glycol succinate-co-furandicarboxylate); POF, poly(octylene furandicarboxylate); PP, polypropylene; PPeF, poly(pentylene furandicarboxylate); PPF, poly(propylene
furandicarboxylate); PPS, poly(propylene succinate); PPS-co-PBA, poly(propylene succinate-co-adipate); PPS-co-PPF, poly(propylene succinate-co-furandicarbox­
ylate); PVP, polyvinylpyrrolidone; ROP, ring opening polymerization; Sb2O3, antimony(III) oxide; SnOct2, stannous octoate; SSP, solid state polymerization; Tc,
crystallization temperature; Td max, maximum decomposition temperature; Tg, glass transition temperature in the second heating scan; Tm, melting temperature in the
first heating scan; TBT, titanium(IV) butoxide; TPA, terephthalic acid; TTIP, titanium(IV) isopropoxide; Zn(Ac)2, zinc acetate.
* Corresponding author.
E-mail address: marie-josee.dumont@mcgill.ca (M.-J. Dumont).

https://doi.org/10.1016/j.eurpolymj.2021.110778
Received 30 July 2021; Received in revised form 17 September 2021; Accepted 19 September 2021
Available online 21 September 2021
0014-3057/© 2021 Elsevier Ltd. All rights reserved.
S. Pandey et al. European Polymer Journal 160 (2021) 110778

fermented by microorganisms [3]. The production routes for biobased Table 2


polyesters, such as polylactic acid (PLA), poly(butylene succinate) Applications of furanic based homo-/copolyesters for food and beverage
(PBS), and poly-(hydroxyalkanoates) (PHA) have already been well packaging.
established [4–8]. However, their thermomechanical and barrier prop­ Polymers Applications Reference
erties are inferior to poly(ethylene terephthalate) (PET), which is b
PEF, PBF, Find applications in food and beverage [83,101,161]
considered the most suitable polymer for beverage packaging. The lack and PPF packaging, frozen food trays manufacturing, [156]
of aromatic ring structures in the architecture of the current biobased and mainly for bottle manufacturing. They
polymers might be the one reasonable hypothesis for their inferior exhibit better barrier properties than
commercially adopted bioplastic such as PET,
properties (summarized in Table 1). Therefore, the exploration of
PBS, PLA and PBA-co-PBT, but are not
monomers having rigid aromatic ring structures derived from biomass biodegradable
has become significant for the development of the next generation of a
PBA-co-PBF Can be used as a biobased replacement for PBA- [73]
biobased polymers [9]. co-PBT and is 100% compostable and
Sugar-derived monomers consisting of cyclic structures impart high enzymatically degradable
a
PIsCBF Potential use in food packaging and [155]
glass transition temperatures (Tg), as well as excellent gas and water manufacturing of heat resistant containers
vapor barrier properties when polymerized into polyesters. Amongst a
PPS-co-PPF Exhibit superior barrier properties compared to [72]
them, 2,5-furandicarboxylic acid (FDCA) has been included in the top- PLA and PBA-co-PBT, and offers competitive
15 list of molecules by the US Department of Energy (DOE) having the ability for food packaging and agricultural films
b
PNF Have been found suitable for low moisture [106]
potential to replace petrochemical monomers such as terephthalic acid
foods, fatty foods, alcoholic beverages, and
(TPA) [10]. Its structural similarity to TPA makes it suitable for the acidic foods
production of polyesters without a significant decrease in Tg. The FDCA a
PBF-co-DGA Suitable for processed fruits and vegetable [145]
molecule is composed of a furan ring with two carboxylic acid groups, under modified atmospheric packaging
b
which can be converted into succinic acid, 2,5-furandicarbaldehyde, P(CHDM)F These are high molecular weight polymers [108]
suitable for beverage packaging
2,5-dihydroxymethylfuran, among others. The polycondensation reac­
tion of FDCA with ethylene glycol produces poly(ethylene furanoate) a: biodegradable bioplastics, b: non-biodegradable bioplastics.
(PEF), which has superior mechanical properties, improved gas barrier
properties, and reduced non-renewable energy consumption as the minimum selling price (MSP) was estimated to be around $1490/
compared to PET [11–13]. Moreover, FDCA has been copolymerized ton. They showed that the main processing costs are associated with the
with aliphatic and cyclic acid comonomers to improve biodegradability. HMF formation stage ($650/ton) and the long residence time (4.3 h) for
These copolyesters have demonstrated outstanding thermomechanical HMF oxidation. The group further suggested that the reduction in resi­
and gas barrier properties making them suitable candidates for food and dence time (up to 40%) and feedstock cost (by 10%) could drop the MSP
beverage packaging (Table 2). to $1310/ton, which could compete with the current production process
The most common route for FDCA synthesis is through oxidation of of TPA ($1445/ton) [30]. A techno-economic analysis reported MSP of
5-hydroxymethylfurfural (HMF), the latter being produced by carbo­ HMF as $1715/ton from fructose as feedstock, which was further
hydrate dehydration. The production of HMF has been extensively reduced to $1446/ton by changing the feedstock to glucose [31].
studied by several authors using feedstocks ranging from simple sugars Replacing simple sugars with food wastes could be a sustainable option
to complex carbohydrates [14–21]. However, the direct production of for HMF production from environmental and economic perspectives.
FDCA from biomass is still limited to fructose, glucose, sucrose, high Studies are now emerging where, for instance, starch (bread wastes,
fructose corn syrup, jerusalem artichoke, and starch [22–27]. The in­ cooked rice, pasta, and noodles) and cellulose-rich food wastes (fruit
dustrial production of FDCA has not reached its full potential due to its peels) have been used to synthesize HMF [14,15,32].
low yield and high production cost as the current average for FDCA Sustainability has become an integral part of the field of polymer
synthesis is $2300/kg. To make the process economically viable, FDCA sciences; therefore, this review focuses on providing condensed infor­
production cost should be below $1000/ton [28,29]. Motagamwala mation related to the synthesis of FDCA using heterogeneous catalysts,
et al. (2018) described a process to produce FDCA from fructose where base-free systems, and biomass-derived feedstocks. Herein, the effect of
the addition of aliphatic and cyclic monomers in the FDCA polymeric
chain has been discussed in relation to biodegradability and compost­
Table 1
ability aspects. Finally, this review presents the effect of copolymeri­
Examples of examples of food and beverage packaging and their limitations.
zation on thermomechanical, optical, and gas barrier properties of FDCA
Polymers Limitations Reference based polymers keeping food and beverage packaging in retrospect.
a
PLA Thermally instable, difficult to heat seal, [157]
brittle, low melt strength, high water vapor 2. Industries involved in FDCA production
and oxygen permeability making it unsuitable
for oxygen and moisture sensitive beverage
packaging The downstream processing and catalytic selectivity are major crit­
c
PBAT Comparable properties to LDPE, but is not [158] ical factors for deciding the efficiency of the FDCA production process.
biobased as it uses terephthalic acid as a Avantium and BASF, under a joint venture (Synvina), successfully
comonomer improved the separation technology and catalyst selectivity to produce
a
PHA and PHB Stiffness, brittleness, thermal instability and [159]
poor impact resistance restricting their
FDCA with the MSP of $1000/ton. In early 2020, Avantium signed a
applications in food packaging memorandum of understanding (MOU) to locate a pilot plant in the
a
Starch and Not suitable for hydrophilic foods, low water [160] Netherlands with a production goal of 5 kilotons using the "YXY" tech­
b
cellulose based vapor barrier which is responsible for poor nology. This technology involves the dehydration of carbohydrates in an
blends processability, brittleness, vulnerability to
alcohol medium to form alkoxymethylfurfural or methox­
degradation, limited long-term stability and
poor mechanical properties ymethylfurfural and methyl levulinate, and their subsequent oxidation
d
PET Highly used for bottle manufacturing and [139] to FDCA using acetic acid. This is followed by the polymerization of
beverages packaging, but it is not biobased or FDCA with ethylene glycol to form PEF [29]. Several other companies
biodegradable including Dupont, Archer Daniels Midland Company, Origin Materials
a: biodegradable or compostable bioplastics, b: non-biodegradable bioplastics, c: and Eastman Company, AVA Biochem, and the VTT research center have
biodegradable synthetic plastics, and d: non-biodegradable synthetic plastics.

2
S. Pandey et al. European Polymer Journal 160 (2021) 110778

also made successive breakthroughs in developing their technologies for imperative to elucidate the initial step of sugars conversion into HMF.
the commercial production of FDCA [33]. A two-step process for the The current industrial production of HMF involves the thermo­
conversion of sucrose, glucose, and fructose was introduced by Petro­ chemical acid-catalyzed dehydration of fructose via the loss of three
bras. This process consists of converting sugars into HMF as the first water molecules to yield HMF through a cyclic or acyclic route (Scheme
reaction step over an ion exchange resin, followed by their oxidation to 1) [47,48]. Glucose, on the other hand, requires an additional step of
FDCA [34]. Each technology has its challenges that impede the large- isomerization to fructose before HMF dehydration. According to Antal
scale production of FDCA. Most of the microbiological processes yield et al., the dehydration of HMF from sucrose proceeds via a cyclic route
a low amount of FDCA for prolonged reaction times. However, the through the formation of a fructofuranosyl-cation intermediate [48]. An
fermentation pathway is the greenest option compared to the use of alternate route was followed by Moreau et al. where 3-deoxyhexosulose
toxic solvents and high temperatures. Using homogenous catalytic sys­ was formed as an intermediate by elimination of 3-hydroxyl group. The
tems such as Co(OAc)2/Mn(OAc)2/HBr and Co(OAc)2/Zn(OAc)2/NaBr presence of furfural obtained by retro-aldo cleavage suggested that the
under high pressure and air oxidation of HMF generated approximately acyclic route was prevalent during the HMF synthesis from fructose
60% FDCA yield. Pure oxygen and t-BuOOH are the commonly used [47]. Since the acid-catalyzed carbohydrate is a complex reaction, un­
oxidant during FDCA production [35]. However, the practical applica­ wanted reactions, such as rehydration, fragmentation, reverse-aldo re­
tion of homogeneous catalysts in HMF to FDCA oxidation suffers major action, and condensation, occur in parallel. For instance,
drawbacks including low FDCA yield, byproduct formation, and diffi­ dihydroxyacetone and glyceraldehyde appear from reverse-aldo reac­
culty in separation and catalyst recovery [36]. Nevertheless, a majority tion of fructose, formic acid (FA), and levulinic acid (LA) as the rehy­
of companies rely on chemical processes for FDCA production and are dration products of HMF, and condensation leads to humin formation
now focusing on developing new techniques using heterogeneous cata­ [48]. When complex carbohydrates are present in the feedstock, such as
lysts and green solvent systems [37]. for food wastes, the formation of HMF undergoes three steps including
Metal-supported heterogeneous catalysts, for instance, gold (Au), the hydrolysis of complex di-/polysaccharides to glucans, their isomer­
platinum (Pt), ruthenium (Ru), and palladium (Pd) have been commonly ization to fructose, and dehydration to HMF. The dehydration and hy­
used for HMF to FDCA oxidation. Pt supported on carbonaceous mate­ drolysis steps are favored by Bronsted acids through the protonation of
rials such as Pt/C and Pt-Pb/C under 1 bar O2 pressure gave 81 and 99% the hydroxyl groups of sugars, while the isomerization is catalyzed by
FDCA yield, respectively [38]. The increase in FDCA yield by Pd doping Lewis acids which facilitate the 1,2-hydride shift by accepting electrons
led to the development of bimetallic catalysts for HMF oxidation as re­ [15].
ported by several authors [39–41]. A range of Pd catalysts have been Oxidation of HMF to FDCA under different catalytic systems
synthesized by doping palladium nanoparticles with C, PVP, and metal The oxidation of HMF to FDCA can occur through chemical, enzy­
oxides as support materials. PVP stabilized Pd nanoparticles (Pd/PVP) matic, or electrochemical processes. Amongst them, the chemical
catalyst gave similar yield as obtained with Pt (99% FDCA yield) transformation seems to be most promising at the industrial level owing
[42,43]. The support materials used for metal deposition could also to high yields, product purity, and reaction rate [49]. The chemical
impact the FDCA yield, for example, Au nanoparticles doped on CeO2 conversion of HMF to FDCA is achieved using high temperature in
and TiO2 afforded higher FDCA yield (>99%) than Fe2O3 and C [44]. presence of additives, (such as base including NaHCO3, Na2CO3, NaOH,
Ru-based catalysts were initially used for HMF oxidation to DFF rather K2CO3) while the biocatalytic transformation occurs at relatively mild
than FDCA synthesis. The mechanism of HMF to FDCA oxidation over Ru conditions. However, the yield obtained by the latter is not appreciable
catalysts and the impact of base materials such as La2O3, MgO, and HT [50]. The oxidation route of HMF to FDCA is dependent on the type of
have been well researched [45,46]. reaction system used and can proceed through two routes (Scheme 2).
The first pathway (Route 1) involves the oxidation of the aldehyde group
3. Chemistry involved in FDCA production to form 5-hydroxymethyl-2-furan carboxylic acid (HFCA); while the
second route follows the alcoholic group oxidation to 2,5-furan dicar­
To understand the reaction pathway for FDCA production, it is boxaldehyde (DFF). These intermediate products are then transformed

Scheme 1. Reaction pathways for acid-catalyzed dehydration of glucose and fructose [47,48].

3
S. Pandey et al. European Polymer Journal 160 (2021) 110778

Scheme 2. Reaction mechanism for HMF oxidation to FDCA [49].

to formylfurancarboxylic acid (FFCA), which gets oxidized to FDCA concentration favored the HFCA route to FDCA formation, with no DFF
[30]. detected. When the solute pH dropped below 3.8, FDCA was protonated
While homogeneous catalysts, such as [Co/Mn/Br] and [Co/Zn/Br], which facilitated its direct precipitation. Being the least expensive of all
provide high FDCA yields during HMF oxidation, the non-recyclability the noble metals, ruthenium supported on MnCo2O4 gave a 99.1% FDCA
of these catalysts has limited their use in various reaction systems yield from base-free oxidation of HMF in water. The catalyst exhibited
[35,51]. Therefore, heterogeneous catalytic systems have gathered excellent recyclability without the loss of active species [56].
much attention in recent years. Several authors have reported the use of Moreover, the use of ionic liquids (ILs) as solvent is advantageous for
supported heterogeneous noble metal catalysts, such as Au, Pt, Pd, and HMF to FDCA oxidation due to the high FDCA solubility. The catalyst
Ru and their alloys, for the conversion of carbohydrates into FDCA under inactivation due to FDCA deposition is minimized by the use of ILs.
basic environment [23,25,42,49]. Supported Au nanoparticles (NPs) Besides, non-noble metal catalysts, such as Fe-Zr-O and Fe-Zr-Ce-O, were
have shown excellent performance for the oxidation of HMF to FDCA investigated to reduce the cost of HMF oxidation using [BMIM]Cl in a
under aerobic conditions. For instance, CeO2 supported Au NPs pro­ base free reaction system, yielding 61 and 44 mol% of FDCA yield,
duced high FDCA yield (99 mol%) from HMF oxidation using highly respectively after 24 h [59,60]. The FDCA yield in [BMIM]Cl was
concentrated NaOH (4 equivalent of HMF) at 130 ◦ C under 1 MPa air enhanced to 89 mol% by using heteropolyacids at 140 ◦ C, and 1 MPa O2
pressure [52]. TiO2/Au was less efficient, leading to a yield of only 65 in 6 h [61]. Furthermore, Motagamwala et al. (2018) used equal parts of
mol% FDCA from HMF at a high NaOH to HMF ratio (>20) [53]. A GVL and water to synthesize FDCA (93 mol%) from fructose-derived
comparative study on carbon-supported Au, Pt, and Pd NPs reported HMF using base-free conditions. The high yield obtained was due to
that Pt/C and Pd/C were more efficient for HMF oxidation (71–79 mol% the enhanced solubility of FDCA in the solvent system, which decreased
FDCA) than Au/C NPs at a constant HMF to KOH molar ratio of 1:2, 7 bar with decreasing temperature, facilitating FDCA crystallization below
O2, 295 K and 7 h of reaction time [40]. 4 ◦ C [30].
Although the basic environment for HMF oxidation generally pro­ Few studies have focused on the synthesis and application of mag­
vides high FDCA yields, these are characterized by the high base usage netic catalysts. Zhang et al. studied the catalytic effect of magnetic
and O2 pressure, which can triggers a parallel reaction for HMF degra­ palladium nanocatalyst for efficient conversion of HMF to FDCA (93 mol
dation. Other compounds, such as FA, LA, and humins, can be formed, %) in water at low K2CO3 concentration (0.5 equivalent mole ratio of
which depict the instability of HMF under high pH conditions. More­ HMF) under aerobic conditions [49]. The reaction time and energy
over, the product formed in a basic environment is FDCA salt, which consumption of HMF to FDCA oxidation were substantially reduced
cannot be directly used for the synthesis of polymers, thus needs to be using metal ferrites, MFe2O4 (M = Cu, Mn, Mg, and Co) as catalysts in
separated using strong mineral acids [54]. presence of tert-butyl hydroperoxide as oxidant and acetonitrile as sol­
Base-free HMF oxidation using heterogeneous catalysts has been vent. The highest FDCA yield was reported as 85 mol% at 100 ◦ C for 5 h
extensively investigated by various researchers [55–58]. Hydrotalcite using Mn as the active species [62]. Furthermore, a two-step one-pot
supported Au NPs showed 99% FDCA yield from HMF in water at 95 ◦ C conversion of fructose to FDCA (60 mol%) was achieved using nano-
and 7 h under aerobic conditions. The basicity of metallic supports is a Fe3O4-CoOx catalyst and tert-butyl hydroperoxide as the oxidant in a
crucial factor to be considered during HMF oxidation to FDCA. DMSO solvent system [63].
Compared to neutral or acidic supports, such as Au/SiO2 and Au/Al2O3 Although FDCA could be produced from HMF with nearly quanti­
which yield almost no FDCA, basic supports, such as MgO and hydro­ tative yields (ranging from 60 to 99 mol%), the cost of HMF is high.
talcites, achieved full HMF conversion leading to high FDCA yields [55]. Carbohydrates are much cheaper and more abundant than HMF [44,64].
However, the instability of hydrotalcites due to high Mg2+ leaching and Therefore, it is more attractive to carry out the oxidative conversion of
MgO dissolution during FDCA formation limits their use and their carbohydrates into FDCA by one-pot reaction over multiple functional
recyclability [57]. The incorporation of a second metal with gold using catalysts combing acidic and metal sites. There have been few studies on
carbon nanotubes as support (AuPd/CNT) enhanced the basicity of Au- direct carbohydrate conversion into FDCA. For example, Kroger and co-
based catalysts. It was reported that 100% HMF conversion can be workers used a biphasic system composed of water/methyl isobutyl
achieved in 12 h using nascent O2 as the oxidant [58]. Another study ketone (MIBK) for one-pot conversion of fructose to FDCA using a pol­
investigated the effect of the support material (MgO–MgF2) basicity on ytetrafluorethylene membrane reactor. The fructose was firstly trans­
the reaction pathway for HMF to FDCA oxidation [28]. The study formed into HMF using Lewatit SPC 108 as the solid acid catalyst, the
showed that increasing the MgO content enhanced the FDCA production HMF was then extracted in the MIBK phase, followed by its oxidation
due to the formation of basic Mg(OH)2 in water. Higher MgO into FDCA (25%) over metal catalysts. The low FDCA yield and

4
S. Pandey et al. European Polymer Journal 160 (2021) 110778

cumbersome FDCA separation from the by-products were few draw­ alcohol in presence of a catalyst with an excess of diols to obtain olig­
backs in the process [65]. Ribeiro and Schuchardt prepared a bifunc­ omers, followed by the polycondensation reaction. In general, dimethyl-
tional catalyst by the encapsulation of Co(acac)3 in sol–gel silica, 2,5-furan dicarboxylate (DMFD), an ester of FDCA, has been used as the
combining the acidic and redox ability, and studied the one-pot con­ starting material. However, for industrial purposes, direct esterification
version of fructose to FDCA, obtaining fructose conversion of 72% with a is preferred [68,69]. The esterification or transesterification reactions
selectivity of 99%. Compared to the method reported by Kröger and co- are generally carried out in a temperature range of 150–180 ◦ C, while
workers, both FDCA yield and selectivity were greatly improved. the polycondensation is conducted at a higher temperature range of
However, the reaction was carried out at 165 ◦ C and a pressure of 20 bar, 180–230 ◦ C. By the end of the esterification step, the mixture becomes
which are difficult conditions to reach [66]. transparent after the Weissenberg effect, which later turns into a molten
Another system was reported by Zhang and co-workers where a two- product in the polycondensation step. During the polycondensation re­
step method for the conversion of fructose to FDCA was used [67]. The action, diols are produced as the side product. To evaporate the excess of
HMF synthesis from fructose was carried out in isopropanol catalyzed by diols, the process is carried out under vacuum, thus preventing polymer
HCl. The isopropanol was recollected for the next run by evaporation, oxidation, and facilitating the formation of a high molecular weight
followed by HMF oxidation to FDCA (83%) in water using Au/HT (Mw) polymer [70].
catalyst. The deposition of humins over the catalyst’s surface led to the Solid-state polymerization (SSP) is an additional step to further in­
deactivation of the catalyst. Using Jerusalem artichoke tuber (fructose crease the Mw of the melt, by heating the polymer between its glass
units being the major component) as the feedstock led to an overall transition (Tg) and melting (Tm) temperatures [71]. The most commonly
FDCA yield of 52%. A recent work involved the use of a triphasic reactor used organometallic catalyst during melt polymerization is titanium(IV)
to convert sugars to FDCA in a one-pot process [23]. The triphasic sys­ butoxide (TBT), which accelerates the polymerization rate of dicar­
tem is composed of tetraethylammonium bromide (TEAB) or water boxylic acids [23,63,67]. Other catalysts commonly used during the
(phase I)–methyl isobutyl ketone (MIBK) (phase II)–water (phase III). In polycondensation step are titanium(IV) isopropoxide (TTIP), antimony
the designed triphasic setup, sugars (fructose or glucose) were first (III) oxide (Sb2O3), and zinc acetate (Zn(Ac)2) [75–78].
dehydrated to HMF in Phase I. HMF was then extracted, purified, and Several investigations have been reported on the use of combined
transferred to Phase III via a bridge (Phase II). Finally, HMF was catalytic systems to improve the reactivity and polymerization rate. A
oxidized to FDCA over Au/HT catalyst in phase III. Overall, FDCA yields white fibrous polymer with the film-forming property was prepared by
of 78% and 50% were achieved with fructose and glucose feedstock, melt transesterification of DMFD with 1,6-hexanediol in presence of
respectively. Phase II plays multiple roles as a bridge for HMF extraction, calcium acetate and Sb2O3 as the catalyst. The number average molec­
transportation, and purification. The same group synthesized Fe3O4@­ ular weight (Mn) was low (below 10,000 g/mol), while the distribution
SiO2–SO3H acid catalyst which dehydrated fructose to HMF in DMSO as was relatively high (polydispersity index = 2.5) [79]. A patent reported
the solvent system. The HMF was then oxidized to FDCA with t-BuOOH the use of Sn/Ti catalysts to prepare polyesters by reacting FDCA with
over nano-Fe3O4–CoOx catalyst. The use of earth-abundant metals and different monomers, such as ethylene glycol, 1,3-propanediol, and 1,4-
magnetic catalyst provide more practical approach from recyclability butanediol monomers in the first step, with subsequent SSP to in­
and economic point of view [63]. crease their Mw [70]. Kato & Kasai reported the use of TBT and mag­
In a nutshell, the direct conversion of carbohydrates to FDCA has nesium acetate during the ester exchange step to produce PEF and PBF,
come a long way, however the current technologies need significant where PEF was reported to have excellent heat resistance and thermo­
improvement for large scale production. Here are some of the challenges mechanical properties. However, the reaction time was too long (7.5 h)
associated with the existing catalytic systems: to achieve high Mw [80]. The combined effect of La(acac)3 and TBT
proved to be more reactive than TBT alone, to catalyze the poly­
1. The feedstocks mainly getting attention for FDCA production are condensation of PBF. However, no significant effect on the polymer
fructose and glucose, while food and lignocellulosic wastes have not thermal stability was observed [81]. During SSP kinetics of PEF, it was
been well explored yet. observed that Sn-based catalysts gave higher Mw products than Ti-based
2. The separation of FDCA from the catalytic mixture is also an catalysts. Moreover, it was reported that TTIP showed higher reaction
important point which still needs to be focused upon. Many of the rates, while dibutyltin (IV) oxides (DBTO) enhanced the crystallinity
current studies use excess of base for HMF to FDCA oxidation which [82,83].
requires an additional neutralization due to the production of FDCA
salt. This results in a less green and cost intensive process. Therefore, 4.1.2. Ring-opening polymerization
much more effort is needed to develop more environmentally Most of the FDCA-based polyesters have been produced by poly­
friendly catalytic systems that can effectively promote the base-free condensation using organometallic catalysts. Nevertheless, researchers
oxidation of HMF into FDCA. have synthesized FDCA copolymers with lactide, isohexides, and
3. Preparation of multifunctional catalysts is an important area of caprolactam as comonomers using entropically driven ring-opening
research as they have shown to reduce the amount of side reactions. polymerization (ROP) to obtain high Mw polymers. Being a greener
The selection of the active sites on the catalyst is essential for the one- option, ROP has the advantages of minimum polymer coloration, low
pot conversion of carbohydrates to FDCA. melt viscosity, and shorter reaction time without byproduct formation.
4. In comparison to catalytic activity of noble metals, the non-noble ROP is triggered by the rupture of the aromatic molecule to form a linear
metals are much less studied. From the viewpoint of practical polymer in presence of suitable catalysts. The synthesis involves high
application, transition metal catalysts such as Co, Fe, Mn can be a dilution condensation to form cyclic oligomers followed by cyclo­
more suitable choice in terms of cost, stability, and recyclability, for depolymerization. During the former step, the acid dichloride reacts
the aerobic oxidation of HMF to FDCA. with diols under high dilution concentrations (<0.005 M) to form
instant cyclic oligomers in the presence of DABCO as the catalyst. The
4. FDCA based homopolyesters and their properties crucial step is to remove the linear oligomers formed during the reaction
using flash chromatography. The most commonly used catalyst for ROP
4.1. Mode of polymerization is stannous octoate (SnOct2) [84–86]. The Mw of PEF synthesized using
ROP lied between 50 and 60 kg/mol. However, more research is needed
4.1.1. Melt polycondensation to explore the polymer properties [87].
The reaction proceeds through direct esterification or trans­
esterification of FDCA or its corresponding ester, with polyhydroxy

5
S. Pandey et al. European Polymer Journal 160 (2021) 110778

4.1.3. Enzyme assisted polymerization because of the plasticization effect, and a subsequent rise in the melting
Apart from chemical catalysis, enzymatic assisted ROP has been re­ enthalpy (ΔHm) with increasing Mw [78]. The tensile properties of PET,
ported by various authors for the synthesis of FDCA-based polyesters such as Young’s modulus (E), strength at break (σb), and elongation at
[81,84,88–90]. The catalytic activity of these enzymes depends on the break (εb) are respectively 2.2 GPa, 57 MPa, and 200%, while the cor­
solvent used being more efficient in a hydrophilic environment. Lipases responding values for PEF are 3.7 GPa, 72 MPa, and 1%. Even though
have been successfully used during ROP. For instance, cyclic (butylene PEF has some superior properties to replace PET, the main disadvan­
2,5-furandicarboxylate) oligomers were synthesized and copolymerized tages are brittleness, low εb, and slow crystallization rate-limiting its
with ε-caprolactone by ROP in presence of Candida antarctica lipase B processability during stretch-blow molding. In particular, a study
(CALB). Since the process avoids the use of both organic solvents and demonstrated that the incorporation of CHDM units in PEF led to
organometallic catalysts, it is, therefore, a greener option as compared comparable E (2.2 GPa) and σb (56 MPa), but superior εb (180%) to that
to other polymerization techniques [84,88]. However, milder conditions of PEF, proving its suitability for beverage packaging applications [98].
used during ROP prevent the decarboxylation reaction of FDCA leading
to low Mw products in comparison to melt polycondensation. The 4.3. Gas barrier properties
copolymer poly(butylene succinate-co-butylene furanoate) PBSF syn­
thesized by ROP using CALB as catalyst had lower Mw (21–47 kg/mol) as The crystallinity level has a major influence on gas permeation rates
compared to chemical assisted ROP (53–63 kg/mol) [81,84]. as the gas molecules cannot permeate, diffuse, or transport through the
crystalline regions of a polymer. Burgess extensively studied the O2,
4.1.4. Solution and interfacial polycondensation CO2, and water vapor transport phenomena in PEF and compared it to
Polymerization in solution has the merit of being carried out under PET, the latter is not suitable for O2 sensitive beverage packaging.
mild reaction conditions but has associated shortcomings such as Despite its higher free volume (1.6 times), the O2, CO2, and water vapor
coloration in the product, low Mw, prolonged reaction time, and the use permeability of PEF were significantly lower (11, 19, and 2.1 times,
of large volumes of solvent. PEF synthesized by solution polymerization respectively) than PET. The enhanced barrier properties for PEF were
has significantly lower Mw (2500 g/mol) than that produced by trans­ ascribed to the suppressed ring flipping motion due to the nonlinear axis
esterification reaction (44500 g/mol) [91]. In terms of Mw and yield of and ring polarity of the furan ring as compared to PET’s symmetrical and
isolated polymers, the transesterification of monomers gave superior non-polar phenyl ring [11,99,100]. Besides the hindered furan ring-
results than those prepared by solution polycondensation for PEF (as flipping, the gas barrier properties of PEF is also affected by the furan
indicated in Table 3). ring’s polarity [101]. In addition, the inclusion of a cyclic monomer
(CHDM) with a furan ring enhanced the O2 barrier properties of PEF, as
4.2. Thermomechanical properties the PECF bottles showed better barrier properties for beverage pack­
aging than commercially available PET [87,102]. The rise in CHDM
Amongst all thermal properties, glass transition temperature, Tg is an units (10 to 98%) improved the O2 and CO2 BIFP of PECF polymers,
important polymeric property to decide the applicability of the respec­ resulting in an increase in CO2 and O2 BIFp from 2.4 to 6.8 and 2.3 to 6.7,
tive polymers in different fields. Compared to PET, PEF is more amor­ respectively [103].
phous due to its higher specific and free fractional volume. However, Furthermore, the syn or anti-conformation of FDCA polymeric chains
PEF displays a higher Tg than PET due to its rigid chain structure which have a significant effect on the gas and water vapor permeation rate. The
restrains ring flipping. As a result of the suppressed nonlinear axis permeability through the polymer is restricted by the formation of
rotation, PEF experiences delay in the crystallization process, thus C–H⋯O interactions which occurs when FDCA exists in a syn confor­
lacking the presence of cold crystallization temperature and subsequent mation. The syn or anti conformations are dependent on the number of
melting peak during differential scanning calorimetry analysis [11]. odd methylene groups present [104]. Vannini demonstrated that the
The use of cyclic diols, such as isosorbides, 2,2,4,4-tetramethyl-1,3- permeability of packaging films depends on the presence of an odd
cyclobutanediol (CBDO), and 1,4-cyclohexanedimethanol (CHDM), number of methylene groups and the length of the aliphatic chain of
when used as monomers during polymerization, further restrict the diols. The O2 and water vapor transmission rate of poly(propylene fur­
chain mobility and increases the polymer’s Tg. For instance, the poly­ andicarboxylate) (PPF) is lower than PEF and is further reduced with
merization of rigid CBDO and flexible CHDM with FDCA enhanced the aging time due to increasing crystallinity and density. The permeability
Tg of copolyesters ranging from 80.3 to 105.7 ◦ C. The Tg varied linearly loss in aged PPF films is attributed to the highly packed and ordered
with the amount of CBDO units present, while larger amount of CHDM polymer crystals with low free volume hindering molecular diffusion
units enhanced the Mw of the resulting copolyesters [91–93]. Apart from [105].
the concentration of cyclic diol (CHDM), the ratio of cis to trans CHDM The gas transmission rate in a film varies with temperature as
affects the thermal properties of the polyesters formed due to the steric described by the Arrhenius equation, i.e. exponential increase with
configuration of the CHDM units [94,95]. temperature. At elevated temperatures, the macromolecules’ diffusive
Also, increasing the number of C-atoms in linear aliphatic diols is jumps increase the free volume, which subsequently increases the gas
responsible for Tg lowering as observed for poly(pentylene fur­ permeation rate in the polymer matrix. Another important factor to
andicarboxylate) (PPeF), poly(hexamethylene furandicarboxylate) include in permeation phenomena is the activation energy (Ea), which
(PHeF), poly(heptylene furandicarboxylate) (PHepF), and poly(octylene represents the energy required to initiate the permeation process. A
furandicarboxylate) (POF). This phenomenon is related to the creation study has reported that the Ea for CO2 transmission in PPF was 32 kJ/
of more amorphous regions in the polymer [91,96,97]. mol, which is higher than for PEF (23.7 kJ/mol reported by Burgess
For copolyesters, the Tg lies between the Tg values of its corre­ et al., 2014) [11,104]. Furthermore, a novel film, poly(neopentyl glycol
sponding homopolyesters and can be described by several mathematical furandicarboxylate) (PNF) was tested in contact with food simulants to
models namely, Flory-Fox, Gordon-Taylor, Couchman-Karasz, and Kwei evaluate the influence of temperature on gas permeability and food
equations. The Flory-Fox equation is relatively simpler but sometimes respiration rate. The presence of branched methylene groups in PNF
fails to represent the experimental values, while other equations take reduced the mobility of the carbonyl groups, thus lowering the gas
additional parameters into account, such as heat capacity, hence fitting transmission ratio. The food simulants used to evaluate the gas trans­
the data more accurately than the former equation [97]. The Mw also has mission ratio were ethanol (10, 20, and 50% (v/v)) and acetic acid (3%,
a prominent influence on the thermomechanical properties of the FDCA v/v), representing food with a hydrophilic nature, alcoholic beverages,
polyesters. For instance, a study reported that low Mw species exhibited lipophilic foods, and foods under a pH of 4.5, respectively. The highest
reduced Tg, storage modulus (E′ ), and crystallinity for PBF samples gas transmission ratio was reported for foods with a pH under 4.5.

6
Table 3

S. Pandey et al.
FDCA based homopolyesters and their properties.
Polymer Staring materials Polymerization conditions Catalyst used Polymer properties Ref

PEF Ethylene glycol Esterification: 280 ◦ C, 4 h Sn/Ti Mn = 23000 g/mol; Tm = 170 ◦ C; Tg = 85 ◦ C; Tc = 156 ◦ C; Tdec = [70]
Polycondensation: 280 ◦ C, 6.5 h 332 ◦ C
Solid state condensation: 180 ◦ C
Ethylene glycol Transesterification Sb2O3 Mn = 22400 g/mol; Mw = 44500 g/mol; PDI = 1.99; Tm = 215 ◦ C; [91]
Tg = 80 ◦ C; Tc = 165 ◦ C; Tdec = 398 ◦ C; Y = 79%
DMF + ethylene glycol Transesterification TTIP Y = 93.24%; Mn = 13050 g/mol; Mw = 34000 g/mol; PDI = 2.6 [113]
Ethylene glycol Polytransesterification: 200 ◦ C for about 2 h TBT (0.1 mmol) Mw = 25973 g/mol, ƞ = 0.47 dL/g; Y = 82%; Tm = 201 ◦ C ; Tg = [116]
89 ◦ C ; Tdec = 326 ◦ C ; T50 = 376 ◦ C
2,5-furandicarbonyl chloride + ethylene Solution – Mn = 2000 g/mol; Mw = 2500 g/mol; PDI = 1.25; Tm = 215 ◦ C; Tg [91]
glycol = 80 ◦ C; Tc = 165 ◦ C; Tdec = 398 ◦ C; Y = 23%
Ethylene glycol – – O2 permeability: 2.8 ± 0.7 cm3 STP/(cm3 poly atm) [99]
CO2 permeability: 0.094 cm3 STP/(cm3 poly atm)
Water vapor permeability: 787 ± 110 cm3 STP/(cm3 poly atm)
Ethylene glycol – – O2 permeability: <0.0702 cm3⋅cm/m2⋅day.atm [105]
Ethylene glycol (EG), 1,4-cyclohexanedi­ Esterification: 180 ◦ C at 120 rpm, 150 min TPT (350 ppm) Mn = 27600–32000 g/mol; Mw = 53600–63880 g/mol; PDI = [108]
methanol (CHDM), and dimethyl furan Polycondensation: 230 ◦ C, 170 min 1.91–2, ƞ = 0.6 dL/g; Tm = 211 ◦ C; Tc = 115 ◦ C; Tg = 87 ◦ C; Tmm
dicarboxylate (DM-FDCA) = 170 ◦ C; Tdec = 395 ◦ C; ΔHm = 1 J/g,
PPF 1,3-propanediol Transesterification Sb2O3 Mn = 21600 g/mol; Mw = 27600 g/mol; PDI = 1.28; Tm = 174 ◦ C; [91]
Tg = 50 ◦ C; Tc = 127 ◦ C; Tdec = 390 ◦ C; Y = 76%
1,3-propanediol Esterification: 230 C, 4 h

Sn/Ti Mn = 15000 g/mol; Tm = 150 ◦ C; Tg = 39 ◦ C; Tc = 102 ◦ C; Tdec = [70]
Polycondensation: 230 ◦ C, 6.5 h 335 ◦ C
Solid state condensation: 140 ◦ C
1,3-propanediol – – Mn = 18200 g/mol; Mw = 54600 g/mol; PDI = 3; [105]
O2 permeability: 0.0472 cm3⋅cm/m2⋅day.atm
Water vapor permeability: 0.762 cm3⋅cm/m2⋅day.atm
1,3-propanediol – – O2 permeability: <0.005 cm3⋅cm/m2⋅day.atm [105]
Ageing time (30 min) Water vapor permeability: 0.243 cm3⋅cm/m2⋅day.atm
7

1,3-propanediol Esterification: 180–200 ◦ C, 2–3 h Antimony trioxide Mw = 30000 g/mol; PDI = 2.3; Tm = 169 ◦ C; Tg = 53 ◦ C; Tc = [90]
Polycondensation: 220 ◦ C at 0.1 mbar for 2 h (10.6 mg per gram of 138 ◦ C; Tdec = 335 ◦ C ; ΔHm = 7 J/g; E = 1363 MPa; σb = 31 MPa;
final theoretical εb = 3%; O2 GTR: 0.0224 ± 3x10− 6; CO2 GTR: 0.0288 ± 3 × 10− 6;
polymer) Water vapor GTR: 0.0157 ± 6 × 10− 5
PBF 1,4-butanediol Esterification: 170 ◦ C, 4 h Sn/Ti Mn = 60000 g/mol; Tm = 170 ◦ C; Tg = 31 ◦ C; Tc = 90 ◦ C; Tdec = [65]
Polycondensation: 180 ◦ C, 6.5 h 338 ◦ C
Solid state condensation: 150 ◦ C
DMF + 1,4-butanediol Transesterification TTIP Y = 103.9%; Mn = 13050 g/mol; Mw = 34000 g/mol; PDI = 2.6 [99]
DMF + 1,4-butanediol Transesterification DBTO Y = 103.9%; Mn = 13050 g/mol; Mw = 34000 g/mol; PDI = 2.6 [99]
2,5-furandicarboxylic acid (FDCA), and Esterification:190 to 210 ◦ C for 4 h with N2 TBT (0.5–1 mmol/mol E 875 ± 18 MPa; σmax = 35 ± 2.6 MPa, εmax = 55 ± 10% [80]
1,4-butanediol (BDO)-(diol/diacid molar Polycondensation: 220 to 250 ◦ C for 4 h under a reduced pressure of diacid)
ratio 1.5:1); of 10 Pa
1,4-butylene glycol Polytransesterification: 200 ◦ C for about 2 h TBT (0.1 mmol) Mw = 24364 g/mol, ƞ = 0.45 dL/g; Y = 93%; Tm = 169 ◦ C; Tg = [116]
36 ◦ C; Tdec = 329 ◦ C; T50 = 367 ◦ C
FDCA and 1,4-butanediol Esterification: FA was purified before use, TBT = 1.4 mmol/mol TBT, (La(acac)3⋅H2O Tm = 170 ◦ C; Tc = 115 ◦ C; Tg = 44.8 ◦ C; Tmm = 170 ◦ C; Tdec = [80]

European Polymer Journal 160 (2021) 110778


diacid, diol/diacid = 2.5, 170 ◦ C/1h + 180 ◦ C/2h + 190 ◦ C/1h; 395 ◦ C
Polycondensation: La(acac)3, La/Ti = 1/1 M ratio, 230 ◦ C/4h +
250 ◦ C/1h
2,5-furandicarboxylic acid (FDCA) and 1,4- Esterification: 190 to 210 ◦ C for 4 h with N2 TBT (0.1 mol% diacid); Tm = 167.8 ◦ C; Tc = 95.4 ◦ C; Tg = 35.6 ◦ C; Td max = 416; oC E = [132]
butanediol (BDO)-(diol/diacid molar ratio Polycondensation: 220 to 250 ◦ C for 4 h under a reduced pressure La(acac)3 (0.1 mol% 1860 ± 160 MPa; σmax = 55.6 ± 1.6 MPa; εmax = 256 ± 19%
2:1) of 10 Pa diacid)
PDASF 2,5-furandicarbonyl chloride + D- Solution – Mn = 13750 g/mol; Mw = 23670 g/mol; PDI = 1.72; Tg = 180 ◦ C; [91]
isosorbide Tdec = 450 ◦ C; Y = 91%
PDAIF 2,5-furandicarbonyl chloride + isoidide Solution – Mn = 5670 g/mol; Mw = 7270 g/mol; PDI = 1.28; Tg = 140 ◦ C ; [91]
Tdec = 396 ◦ C; Y = 80%
PBHMF Bis-(2,5-hydroxymethyl)furan Interfacial – Mn = 3880 g/mol; Mw = 5410 g/mol; PDI = 1.39; Tdec = 345 ◦ C, [91]
Y = 60%
PHQF Hydroquinone Interfacial – Tdec = 490 ◦ C; Y = 77% [91]
(continued on next page)
S. Pandey et al. European Polymer Journal 160 (2021) 110778

However, the stability of PNF was not disturbed during its contact with

[106]
[91] the food simulants. The gas transmission ratio was positively affected by
[75]

[75]

[98]
Ref

relative humidity due to high gas-water interactions [106].


Mn = 22100 g/mol; Mw = 47500 g/mol; PDI = 2.15; Tg = 87 ◦ C ;

0.57–0.66 dL/g; Tm = 239–268 ◦ C; Tc = 115 ◦ C; Tg = 80–84 ◦ C;

MPa; εb = 4%; O2 GTR = 0.0323 cm3⋅cm/m2⋅day.atm; CO2 GTR


4.4. Optical properties

138 ◦ C; Tdec = 335 ◦ C; ΔHm = 30 J/g; E = 1648 MPa; σb = 45


Mw = 50400 g/mol; Mn = 16000 g/mol; ƞ = 0.31 dL/g; Tg =

Mw = 34000 g/mol; PDI = 4; Tm = 197 ◦ C; Tg = 73 ◦ C; Tc =


Tmm = 170 ◦ C; Tdec = 395 ◦ C; ΔHm = 30.6–44.7 J/g; Tcc =
Mn = 33700 g/mol; Mw = 63100 g/mol; PDI = 1.87, ƞ = The commercialization of a polymeric film can also depend on its
transparency. This property is mainly governed by the crystallinity of
150–169 ◦ C; O2 permeability = 0.0036 cc/pkg.day the packaging film since light scattering mainly occurs in the crystalline
region. The commercialization of FDCA based polyesters is also impeded
due to the coloration of the polymeric films, mainly attributed to the
different catalysts or additives used, parallel side reactions occurring at
47.14 ◦ C; T50 = 377 ◦ C; PDI = 3.15

high temperature, presence of sugar-based impurities, and low mono­


= 0.0223 cm3⋅cm/m2⋅day.atm

mer purity [107]. Some purification steps have been adopted during
which the light brown PEF samples turned into white polyester by re-
Tdec = 390 ◦ C; Y = 75%

dissolving in solvents like DMSO or TFA/CHCl3 (1/4 v/v), followed by


Polymer properties

precipitation in methanol [108,109]. It is well-known that the starting


materials used for polymerization have a major influence on the optical
ƞ = 0.34 dL/g

properties of the final polymer. For instance, it was shown that using
DMFD as the starting material for PEF and PBF synthesis generated
fewer colored products than FDCA due to decarboxylation of the latter
under harsh reaction conditions [110,111].
The catalysts generally used for PET synthesis, such as magnesium,
Zn(Ac)2 (0.1%) and TBT

Zn(Ac)2 (0.1%) and TBT

germanium, and cobalt were found to increase the color of FDCA-based


Antimony trioxide

polyesters. The high temperature polycondensation reactions resulted in


TPT (350 ppm)

(0.00075 mol)

darker PEF as compared to the transparent PET. The formation of


Catalyst used

carbonized sugar impurities present in monomers or possible polymer


(0.1%)

(0.1%)

oxidation are two possible reasons for PEF darkening. Since FDCA is a
strong chelating agent, it was proposed that a longer polymerization

time might result in the formation of colored furan-metal complexes


[112]. Results have shown that adding a heat stabilizer (tris(non­
ylphenyl)phosphite) reduced the PEF coloration [113]. Another study
used L*a*b color space system to quantify the change in coloration
during PEF synthesis by bulk condensation. The coloration was found to
Esterification: 180–230 ◦ C at 120 rpm, 60–120 min

be dependent on the type of catalysts and the polycondensation time.


Polycondensation: 220 ◦ C at 0.1 mbar for 3 h

For example, titanium-based catalysts imparted higher yellowness


indices for FDCA-based polyesters as compared to zirconium-based
Polycondensation: 280 ◦ C, 80–120 min

catalysts used for polycondensation reactions [108,114]. In addition,


Transesterification: 220 ◦ C for 4 h

Transesterification: 220 ◦ C for 4 h

ROP and solution polymerization yield fewer colored products as


Polycondensation: 200 ◦ C for 5 h

Polycondensation: 200 ◦ C for 5 h

compared to conventional melt polymerization [115].


Note: tetra-isopropyl titanate (TPT), GTR = gas transmission ratio, Y = polymer yield.

Interestingly, the carbon source used for FDCA synthesis also affects
Polymerization conditions

Esterification: 180 ◦ C, 4 h

the optical properties of the PEF film. The PEF films synthesized from
furan-based FDCA via Friedel-Craft reaction and subsequent iodoform
reaction had better transparency than for HMF-based FDCA. The
coloration of the HMF-based PEF films was attributed to the soluble
Solution

humins which coexist with HMF and are difficult to remove even after
their oxidation to FDCA, later affecting the FDCA purity [110].

4.5. Impact of types of diol and their reactivity on polymer characteristics


Ethylene glycol (EG), 1,4-cyclohexanedi­
methanol (CHDM), and dimethyl furan
carboxylic acid) and 1,3-propanediol

The optical properties of furanic homopolyesters can be improved


5,5-(propane-2,2-diyl)-bis(furan-2-
Bis-(1,4-hydroxymethyl)benzene

using copolymerization by adjusting the ratio of the diol units. The main
issue with this process is the variation of diols’ reactivity with FDCA
FDCA and 1,3-propanediol

dicarboxylate (DM-FDCA)

which affects the polymerization rates and ultimately influences the


properties of the homopolymers, or copolymers formed. A kinetic study
showed that the reactivity of diols depends upon the electronic effect of
Staring materials

Neopentyl glycol

–OH groups and increases with the length of the carbon chain. This
explains why 1,4-butylene glycol is more reactive towards FDCA than
ethylene glycol [116].
The presence of nonvolatile cyclic diols (such as isosorbide, isoidide,
Table 3 (continued )

and hydroquinone) have positive effects on the stiffness of polymeric


chains as compared to aliphatic diols (such as ethylene glycol) [91].
PECF(25–100)

Various authors have reported novel polyesters developed using long-


chain aliphatic diols, mostly consisting of odd methylene groups
Polymer

PHMBF

PPF1

PPF2

[96,106,117]. As expected, longer chain diols (1,5-pentanediol, 1,7-hep­


PNF

tanediol, and 1,9-nonanediol) should have exhibited higher Tm for their

8
S. Pandey et al. European Polymer Journal 160 (2021) 110778

corresponding polyesters. However, the results were surprising as PPeF adipate-co-terephthalate) (PBA-co-PBT), exhibit superior thermo­
displayed higher Tm than PHeF and PNF. Unfortunately, the study did mechanical properties over PLA; and have been shown to biodegrade
not investigate the diol’s reactivity towards polymerization with FDCA under 50% aromatic content [127–129]. The latter has been commer­
[96]. Later, Xie et al. applied the Mayo-Lewis equation for free radical cialized under the trade name Ecoflex (BASF) and Easter-bio (Eastman
copolymerization and deduced the apparent reactivity ratio for ethylene Chemical) [130]. However, the non-renewable nature of terephthalic
glycol (0.75), 1,5-pentanediol (3.78), and 1,6-hexanediol (4.6), which acid has opened opportunities for the search for new biobased mono­
supported the long-chain theory suggested by Ma et al. (2012) [125]. mers for its replacement. Avantium estimated that PET bottles take
The lower reactivity of ethylene glycol was attributed to its higher approximately 300–500 years to decompose, while full PEF biodegra­
volatility [97,117]. dation can be achieved in 250–400 days under an industrial composting
Enzyme assisted polymerization reaction was carried out for a series environment (58 ◦ C in soil) [29]. The incorporation of biobased FDCA
of linear aliphatic diols (having 4–12 C-atoms) and their corresponding instead of nonrenewable terephthalic acid helped the biodegradability
diacid ethyl esters with DMFD to test the specificity of these substrates to of the polymer which was comparable to well-known copolyesters like
CALB. Based on the results, diols seem to be better substrates than the PBS-co-PBT and poly(butylene adipate-co-terephthalate) (PBA-co-PBT)
corresponding diacid ester as they produced a higher polymerization [131]. Thus, the forthcoming sections discuss the thermomechanical
degree. Besides, monomers having higher steric hindrance have an properties and biodegradability of FDCA-based copolyesters which are
electronic advantage towards their reactivity during polymerization summarized in Tables 3 and 4. Table 3 enlists the polymerization
[90]. The novel monomer formed by this new rigid difuranic C11 diol techniques and the homopolyesters properties while Table 4 compares
monomers showed excellent reactivity towards FDCA, due to the pres­ the copolymerization conditions and the properties of the copolyesters
ence of pendant group allowing easier condensation reactions leading to produced.Table 5 summarizes the conditons for the hydrolytic or
higher Mw polyesters [118]. enzymatic degrdation and compostability of FDCA-based copolyesters.

5. Biodegradability induced FDCA based copolyesters 5.2. Copolymerization with aliphatic monomers

5.1. Biodegradability 5.2.1. FDCA and adipic acid copolyesters


Biobased copolyesters of adipic acid and FDCA with different diols
Biodegradability is the property by which a polymer completely (ethylene glycol, 1,3 propanediol, and 1,4-butanediol) have been syn­
decomposes into CO2, CH4, H2O, inorganic compounds, or biomass thesized using melt polycondensation and SSP in presence of different
under the action of microorganisms, light, or chemicals. However, if the catalysts [68,71,75,132]. The addition of biobased aromatic moiety
decomposition of the polymers takes place at a compost site with known (FDCA) in the aliphatic backbone improved the chain rigidity, thermal
microorganisms under controlled conditions, it is called compostable. As stability, and biodegradability of commercially available adipic acid
per the European Bioplastics Association, all food and beverage pack­ polymers, such as PBA and poly(butylene adipate-co-terephthalate)
aging materials should be either biodegradable or compostable to fulfill (PBA-co-PBT) [71]. For instance, poly(butylene adipate-co-
the criteria of being carbon neutral and environment friendly [119,120]. furandicarboxylate) (PBA-co-PBF) can be an alternative to PBA-co-PBT
Although most of the oil-based plastics, such as polyethylene (PE), (also known as Ecoflex). However, the poor optical properties of the
polypropylene (PP), and poly(ethylene terephthalate) (PET), are cate­ former restrict its use in beverage packaging. The color of PBA-co-PBF
gorized as non-biodegradable; some of them, like poly(e-caprolactone) varies with the FDCA content displaying yellowish, claret red, and
(PCL) and poly(butylene adipate-co-terephthalate) (PBA-co-PBT), are chestnut brown polymers at 0–10, 20–60, and 70–100 mol% FDCA
biodegradable in nature. Similarly, it is a misconception that all bio­ content, respectively. Several purification steps have been carried out to
based plastics are biodegradable in nature. For instance, cellulose ace­ obtain FDCA copolyesters of high purity and to reduce the side reactions
tate, which is a natural derivative of polysaccharide cellulose, does not which are the origin of unwanted coloration [132].
degrade into nature and loses its biodegradability due to chemical The thermomechanical properties, crystallinity, and biodegradation
modifications [121]. Thus, the biodegradability of a polymer is rate of PBA-co-PBF range between the adipic acid and FDCA homo­
controlled by the material’s properties rather than its origin. The poly­ polyesters and can be controlled by tuning the amount of FDCA in the
mer degradation rate depends on intrinsic factors, such as types of feed. According to Zhou et al., the addition of FDCA from 0 to 50 mol%
functional groups, wettability, crystallinity, Mw distribution, and reac­ decreased the crystallinity and Tm of PBA-co-PBF due to increased
tivity with catalysts; as well as extrinsic factors including temperature, randomization in the polymers. However, further addition of FDCA (75
relative humidity, and type of microbes present [122,123]. The biode­ to 100 mol%) enhanced those properties due to the crystallization of the
gradability and compostability of bioplastics can be determined using butylene furanoate units. Nearly amorphous copolyesters, owing to low
various standardized tests such as a) anaerobic biodegradability (ASTM Tg (− 44 to − 19 ◦ C) and high elongation (1000%), were obtained for low
D5511 and ASTM D5526), b) aerobic biodegradability (ASTM D5338 FDCA content (20–50 mol%). Conversely, the copolyesters obtained
and ASTM D5209), and c) compostability (ASTM D6400 and EN 13432) were semi-crystalline with high Tg (7–36 ◦ C), good tensile modulus, and
[124]. better tensile strength at higher FDCA content (75–100 mol%) [73].
Currently, PLA (trade name: Bio-flex, Biophan, Ecoloju, Lacea), PHA Moreover, PBA-co-PBF has excellent tensile properties (E = 18 to 160
(trade name: Nodax, Biopol), PBS, and starch-based blends (trade name: MPa, σb = 9 to 17 MPa, and εb = 370 to 910%) making them suitable for
Ecostar, Bioplast, Novon) are commercially available biodegradable a wide range of applications from biodegradable elastomers to ther­
bioplastics for food and beverage packaging. Other aliphatic polyesters, moplastics [132].
such as poly(butylene adipate) (PBA), poly(propylene succinate), poly The main factors affecting the biodegradability of a polymer include
(butylene succinate-co-adipate), and poly(propylene succinate-co- chain mobility, Tm, Mw, and crystallinity. In general, poly(ethylene
adipate), are gaining importance due to their renewable nature and adipate-co-furandicarboxylate) (PEA-co-PEF) and PBA-co-PBF, having
biodegradability. Unfortunately, these aliphatic polyesters are charac­ 10% aromatic content, show better lipase-assisted enzymatic hydrolysis
terized by poor thermomechanical and barrier properties restricting (100% mass loss) than their respective aliphatic homopolyesters. This is
their application as food and beverage packaging [125,126]. As a so­ due to a reduction in Tm and crystallinity for PEA-co-PEF and PBA-co-
lution, aromatic segments (terephthalic acid) have been introduced in PBF. Further increase in FDCA content (>10 mol%) can decrease the
the aliphatic backbone to obtain copolymers with desirable properties biodegradation rate as the aromatic ester bonds are more resistant to
and high biodegradability. Aliphatic-aromatic copolyesters, such as poly enzymatic hydrolysis than the aliphatic ones [68,73]. The hydrolytic
(butylene succinate-co-terephthalate) (PBS-co-PBT) and poly(butylene degradation depends on the chemical structure, crystallinity, and

9
Table 4

S. Pandey et al.
Mechanical, thermal, and molecular characteristics of FDCA based aliphatic-aromatic copolyesters.
Copolyesters (FDCA Copolyester synthesis Catalysts Mw Mn PDI Ƞ Tg Tm Td max Tc ΔHm E σb εb [Ref]
mol%) (g/mol) (g/mol) (-) (dL/g) (oC) (oC) (oC) (oC) (J/g) (MPa) (MPA) (%)

FDCA and adipic acid copolyesters


PBAF (10–75) Direct esterification: TBT 54100–76800 26400–41300 1.72–2.23 – – – – – – 0.14–110 2.6–30 425–1850 [73]
190–210 ◦ C for 4 h under N2 (0.5–1
Polycondensation: mmol/mol
220–250 ◦ C at 10 Pa for 4 h diacid)

PBAF (10–90) Direct esterification: TBT (0.1 65000–103000 30000–57000 1.79–2.25 – − 53.2–23.3 27.4–155.7 417–430 – 6.4–38.6 17.7–122 9.8–42 307–748 [132]
200–230 ◦ C for 4–6 h mol%
Polycondensation: 240 ◦ C at diacid); La
200–400 Pa for 2.5–6 h (acac)3
Drying: Vacuum (60 ◦ C for (0.1 mol%
12 h) diacid)
PPAF Esterification: 200 ◦ C for 5 h Zn(Ac)2 39,300 13,400 2.93 0.21 − 14.2 79.26 400 – 5.92 – – – [75]
under N2 (0.1 mol
Polycondensation: 220 ◦ C for %) and
4 h under vacuum TBT (0.1
mol%)
PEAF (5–95) Transesterification:150 to TBT (400 – – Ng 0.36–0.47 − 32 to 75 44.8–190.3 393.7–406.7 – – – – – [68]
170 ◦ C for 5 h under Ar ppm)
Polycondensation: 170 ◦ C at
5 Pa for 30 min; 220 to
240 ◦ C for 5 h at 720 rpm (for
oligomers)
Melt polycondensation:
220–240 ◦ C for 1–4 h at 5 Pa,
10

400 rpm (for copolyesters)


P(BDA)AF (10–30) Solution precipitation: 1 h – 4470–13560 – 3.9–5.3 0.19–0.51 54–73 262–293 358–365 – 57–73 – – – [71]
stirring and vacuum drying
at 50 ◦ C for 48 h
Polycondensation: 170 ◦ C for
1 h, 220 ◦ C for 6 h under 0.3
MPa
SSP: 220 ◦ C for 12 h, gas flow
rate of 2000 mL/min
PBAF (40–60) Direct esterification: TBT (0.1 99000–167200 – 1.87–2.18 1.12–1.8 – 53.6–112 – – 13–34.2 – – – [69]
200–220 ◦ C for 3–4 h under mol% of
N2 diacid)
Polycondensation: 230 ◦ C at
50–300 Pa for 1 h and 240 ◦ C
for 2–5 h
Drying: Vacuum (40 ◦ C for

European Polymer Journal 160 (2021) 110778


24 h)

FDCA and succinic acid copolyesters


PBSF (10–50) Direct esterification: 190 ◦ C TBT (1.4 35000–65000 18500–29000 1.7–2.7 – − 25-(− 3.5) 75.5–105 396–400 25.2–53.1 19.9–79.5 – – – [81]
for 2 h, 200 ◦ C for 2 h mmol/mol
Polycondensation: 230 ◦ C for diacid)
1 h, 250 ◦ C for 4 h under
vacuum
PBSF (60–90) Direct esterification: 170 ◦ C TBT (1.4 38600–89100 15700–40800 2.38–2.45 – 6.5–30.5 111–159 394–398 76.3–88.2 14.7–37.9 – – – [81]
for 1 h, 180 ◦ C for 2 h, 190 ◦ C mmol/mol
for 1 h diacid), La
Polycondensation: 230 ◦ C for (acac)3
(continued on next page)
Table 4 (continued )

S. Pandey et al.
Copolyesters (FDCA Copolyester synthesis Catalysts Mw Mn PDI Ƞ Tg Tm Td max Tc ΔHm E σb εb [Ref]
mol%) (g/mol) (g/mol) (-) (dL/g) (oC) (oC) (oC) (oC) (J/g) (MPa) (MPA) (%)

4 h, 250 ◦ C for 1 h under (Ti: La =


vacuum 1: 1)
PBSF (5–60) Esterification: 60–80 ◦ C Zr(OBu)4 – 1250–61700 1.8–6.5 0.022–0.191 c31 –( 54–111 – 50–63 31–57 75–84 18–23 952–1400 [135]
under N2 (Zr/diacid 17)
Polycondensation: = 0.0018)
230–270 ◦ C at 0.7 mbar
PESF (5–95) Esterification: 190 ◦ C under TBT: – – – 0.34–0.41 – – – – – – – – [137]
Ar, 500 rpm diacid-
Polycondensation: 230 ◦ C for 0.0003
60 min under vacuum (5 Pa),
720 rpm
PBSF (40–60) Direct esterification: TBT (0.1 162100–167200 – 1.95–2.07 1.12–1.53 – 55.5–112 – – 18.4–34.2 – – – [69]
200–220 ◦ C for 3–4 h under mol% of
N2 diacid)
Polycondensation: 230 ◦ C at
50–300 Pa for 1 h and 240 ◦ C
for 2–5 h
Drying: Vacuum (40 ◦ C for
24 h)
PBSF (10–90) Cyclic oligomers of FDCA CALB 21000–47000 – 1.3–1.7 0.25–0.55 − 25–20 72–140 396–400 – 4–55 – – – [84]
and SA vacuum dried for 24 (40% w/
h w)
Ring opening
polymerization: 120–150 ◦ C
for 24 h
PBSF (10–90) Cyclic oligomers of FDCA Sn(Oct)2 53000–63000 – 1.75–1.95 0.54–0.74 –23–28 60–166 395–401 – 6–71 – – – [84]
and SA vacuum dried for 24 (0.5 mol
11

h %)
Ring opening
polymerization: 180–230 ◦ C
for 12 h
PPSF (10–90) Transesterification: 170 ◦ C TBT (0.15 – 48900–81600 1.7–2 1.03–1.6 − 24.6–47.7 – 410–416 – ng 1–2374 0.2–87.3 6–1730 [72]
(before annealing) for 1 h, 180 ◦ C for 3–5 h mol%
Polycondensation: diacid)
230–240 ◦ C at 10–30 Pa
PPSF (10–90) Transesterification: 170 ◦ C TBT (0.15 – 48900–81600 1.7–2 1.03–1.6 − 24.6–47.7 80.3–151.2 410–416 – 12.5–33.2 45–3077 9.2–95.8 5–725 [72]
(after annealing) for 1 h, 180 ◦ C for 3–5 h mol%
Polycondensation: diacid)
230–240 ◦ C at 10–30 Pa
14
CO2 permeability coefficient: 0.06–0.6 Barrer, O2 permeability coefficient: 0.0072–0.04 Barrer, and water vapor transmission rate: 3.52–9.37 × 10− g⋅cm/ cm2⋅s⋅Pa
PNSF (30–90) Transesterification: 180 ◦ C Zn(Ac)2 – 118000–718000 1.6–2.0 0.9–1.67 – 165–186.3 – – 0.4–15.6 3.1–2527 1.9–88 5.8–2109 [136]
(before annealing) for 3–5 h and Sb2O3

European Polymer Journal 160 (2021) 110778


Polycondensation: 240 ◦ C for
4 h at < 50 Pa
14
CO2 permeability coefficient: 0.03–0.053 Barrer, O2 permeability coefficient: 0.022–0.042 Barrer, and water vapor transmission rate: 3.47–4.83 × 10− g⋅cm/ cm2⋅s⋅Pa
Transesterification: 180 ◦ C Zn(Ac)2 – 118000–718000 1.6–2.0 0.9–1.67 13.5–61.9 66.7–186 441–460 – 9.9–29.2 52–3560 8.9–102 3.5–454 [72]
PNSF (30–90) for 3–5 h and Sb2O3
(after annealing) Polycondensation: 240 ◦ C for
4 h, at < 50 Pa
14
CO2 permeability coefficient: 0.027–0.051 Barrer, O2 permeability coefficient: 0.022–0.04 Barrer, and water vapor transmission rate: 2.98–4.79 × 10− g⋅cm/ cm2⋅s⋅Pa

FDCA and lactic acid copolyesters


PELF (7–73) Solubilization: 80–210 ◦ C for Sb2O3 (1 6900–9000 – – – 25–76.4 119.6 289.3–570.5 84.3 – – – – [140]
5 h, kept for 2 h under wt%) or
vacuum (0.001 mbar) and SnCl2 . p-
constant stirring TSA
(continued on next page)
Table 4 (continued )

S. Pandey et al.
Copolyesters (FDCA Copolyester synthesis Catalysts Mw Mn PDI Ƞ Tg Tm Td max Tc ΔHm E σb εb [Ref]
mol%) (g/mol) (g/mol) (-) (dL/g) (oC) (oC) (oC) (oC) (J/g) (MPa) (MPA) (%)

Precipitation: Dissolved in
TFA, and precipitated by
adding excess of ethanol
PBLF (5–20) (after Transesterification: 180 ◦ C TBT (0.5 – – – – 60 168 360–368 120.5 30.2–36.4 2661–3200 64.1–74.8 183.5–223 [139]
annealing) for 4 h, under N2 mmol)
Polycondensation:
230–245 ◦ C for 2 h, kept at
60 Pa for 0.5 h, Vacuum
drying at 80 ◦ C for 6 h,
Extrusion at 185 ◦ C and 30
rpm, Injection molding:
injection at 185 ◦ C, injection
pressure 60 MPa, back
pressure 100 MPa, mold at
35 ◦ C
PBLF (10–40) Solubilization and TBT (0.15 73500–96500 – 2–2.3 1.1–1.3 37–38.7 90–162.2 402–404 100.9 19.5–25.7 1044–1220 39.8–52.7 6.2–7.8 [86]
precipitation: 150 ◦ C for > 5 mmol),
h, dissolved in CHCl3, Sb2O3
precipitated using CH3OH (0.15 wt
Transesterification: 180 ◦ C %), SnOct2
for 4–6 h, under N2 (0.1 wt%)
Polycondensation: 210 ◦ C for
6 h, under 300 Pa
14
CO2 permeability coefficient: 0.057–0.1 Barrer, O2 permeability coefficient: 0.015–0.033 Barrer, and water vapor transmission rate: 2.4–3.7 × 10− g⋅cm/ cm2⋅s⋅Pa

FDCA and caprolactam copolyesters


High dilution condensation: CALB (10 22000–50000 – 1.25–2 – − 37–19 52–148 382–392 27–136 12–35 – – – [88]
12

PB(CL)F (10–80) 130–150 ◦ C under N2 wt%) or


Transesterification by ROP: TBT (1
200 ◦ C for 24 h mol%)
PB(CL)F (10–90) Transesterification: TBT (0.15 10500–97900 – 1.5–1.8 1.1–1.65 − 34.5–28.4 41–159.5 – − 15.4–100.4 4.7–34.6 11–902 3–68 720–1270 [74]
170–200 ◦ C for 4–6 h, mol% of
Polycondensation: 220 ◦ C at the
15 Pa for 5 h DMFD),
Purification: dissolved in Sb2O3
CHCl3 and stirred at room (0.15 wt
temperature, then %)
precipitated using CH3OH,
dried at 50 ◦ C for 72 h
PPe(CL)F (50–90) Transesterification: TTIP, – – – 0.34–0.42 − 16.1–12.1 40.9–41.3 403.2–408.9 – – – – – [78]
160–190 ◦ C for 4.5 h, under SnOct2
N2, (0.1 mol%
Polycondensation: CL)

European Polymer Journal 160 (2021) 110778


220–230 ◦ C at 5 Pa for 2 h,
ROP: 190 ◦ C for 1.5 h under
N2, 20 min under 5 Pa
PHe(CL)F (50–90) Transesterification: TTIP, – – – 0.24–0.36 − 31.7- 76.3–134.1 400.8–405.9 28.1–89.9 18.1–52.1 – – – [86]
160–190 ◦ C for 4.5 h, under SnOct2 (− 0.1)
N2, (0.1 mol%
Polycondensation: CL)
220–230 ◦ C at 5 Pa for 2 h,
ROP: 190 ◦ C for 1.5 h under
N2, 20 min under 5 Pa
PI(CL)F (10–50) Esterification: 210 ◦ C for 3 h TBT (0.2 33300–1260001 13800–54300 2.32–2.59 0.91–2.14 − 9.9–132.1 – 303.3–369.4 – – 858–1649 44–73 23–480 [152]
at 150 rpm, under N2, mol% of

(continued on next page)


Table 4 (continued )

S. Pandey et al.
Copolyesters (FDCA Copolyester synthesis Catalysts Mw Mn PDI Ƞ Tg Tm Td max Tc ΔHm E σb εb [Ref]
mol%) (g/mol) (g/mol) (-) (dL/g) (oC) (oC) (oC) (oC) (J/g) (MPa) (MPA) (%)

ROP: 220 ◦ C for 3 h at 150 FDCA),


rpm, under N2, (0.2 mol%
Polycondensation: 250 ◦ C for of CL)
3 h at 200 rpm, under 100 Pa
FDCA and isosorbide copolyesters
PDIsF Transesterification: TTIP (400 55300–84500 11500–25400 – – − 1.2–20.6 64.3–110.9 439.2–443.8 39–57.4 30.1–37.9 88.5–284.5 0.75–20.53 191.7–274.5 [77]
160–190 ◦ C for 4.5 h, under ppm)
N2,
Polyesterification:
160–190 ◦ C for 4.5 h, under
N2,
Polycondensation:
220–250 ◦ C for 2 h, under
vacuum
PBImF High dilution condensation: DABCO 39000–63000 – 1.5–2.0 – 45–97 160–169 388–403 146 12–30 – – – [87]
In FDCA-Cl2 and Im (5 (12.5
mmol), THF was added, mmol), Sn
filtered ad then washed with (Oct)2
HCl (0.1 M) (0.5 mol
ROP: 220 ◦ C for 6 h, under %)
N2
PICBFS (10–80) Transesterification: – 57500–67800 – 1.6–1.81 0.77–0.82 74.8–102.3 – 402–403 – – 1242–1330 69–71.5 7–23 [155]
100–180 ◦ C at 1 atm,
160–180 ◦ C at 1 atm
Melt polymerization:
200–240 ◦ C, at 10 Pa
CO2 permeability coefficient: 0.19–1.55 Barrer and O2 permeability coefficient: 0.2–1.67 Barrer
13

PBIF Transesterification: 200 ◦ C DBTO 18400–37400 9300–19100 1.94–2.04 0.45–0.88 55.3–150.6 – 405–416.9 – – 1464–1900 53–140 15–435 [76]
for 2 h at 150 rpm, 210 ◦ C for (0.15 mol
1 h Polycondensation: % of
250 ◦ C for 6 h at 200 rpm DMFD)
under 100 Pa

FDCA and sebacic acid copolyesters


PBScF (5–20) Direct esterification: 200 ◦ C TBT (1 71000–103000 35600–46200 2–2.25 – 6.39–28.26 144.73–166.17 387–393 – 24.57–36.61 – 33.6–50.8 100–780 [144]
for 3 h, at 170 rpm mmol/mol
Polycondensation: 230 ◦ C for diacid)
5 h at 90 rpm under vacuum

FDCA and polyglycolic acid copolyesters


PBF-DGA Direct esterification: 180 ◦ C TBT, TTIP – – – 0.521–0.572 6–26 106–151 380–388 82 14–36 130–688 13–29 241–419 [145]
for 4 h, (150 ppm)
Polycondensation: 220 ◦ C at

European Polymer Journal 160 (2021) 110778


0.1 mbar, 235 ◦ C for 5 min
CO2 GTR: 198–456 (cm3/m2.day.bar), O2 GTR: 113–193 (cm3/m2.day.bar), N2 GTR: 96–189 (cm3/m2.day.bar), and C2H4 GTR: 112–185 (cm3/m2.day.bar)
PBFGA (10–60)- Esterification (OGA): 160 ◦ C Zn(Ac)2 68500–89500 – 1.8–1.9 1–1.15 36.3–38.2 85–156 395–402 – 2.7–33.2 524–1262 15.5–44 223–332 [146]
annealed for 4–6 h under N2, (0.5 wt%),
Transesterification (OBF): TBT,
170 ◦ C for 4–6 h under N2, Sb2O3
Polycondensation:
190–195 ◦ C for 4–5 h at 30
Pa
CO2 permeability coefficient: 0.05–0.11Barrer, O2 permeability coefficient: 0.02–0.05 Barrer, and water vapor transmission rate: 2.65–5.61 × 10− 14 g⋅cm/ cm2⋅s⋅Pa

FDCA and PEG copolyesters


PEF-PEG (20–80) 80500–126000 34000–49000 2.34–2.61 – 78.5–84.9 26–58.4 351.5–360.1 – 10.03–29.85 – 8–27 4–61 [148]
(continued on next page)
S. Pandey et al. European Polymer Journal 160 (2021) 110778

hydrophilicity of the polymer, which is favored by the nucleophilic

[150]

[147]

[149]
substitution of the ester groups by water molecules [75]. This also de­
[Ref]

pends on the pH of the medium, which subsequently controls the mass


loss of the polymeric chains. Under neutral conditions, the carboxylic

565–1087

403–1091
groups of the oligomers and copolyesters autocatalyze the hydrolytic

52.4
degradation process. However, this process is accelerated under alkaline
(%)
εb

conditions. For instance, mass loss of 52% was observed for PBA-co-PBF

22.6–56.2
(having 40–60 mol% butylene furanoate units) compared to 1% in
16.1–66
(MPA)

neutral and acidic conditions within 22 weeks. The PBA-co-PBF with


σb

higher FDCA units (60 mol%) is 100% compostable (ISO 14855-1:2005


and GB/T 19277.2-2013) and degrades faster than its terephthalic acid
29–476

34–655
counterparts (45 mol%) after 180 days. The oxygen-containing furan
(MPa)

ring of FDCA is more hydrophilic than the benzene ring, which explains
E

100.93–164.9 24.42–32.36 –
why PBA-co-PBF is more prone to being attacked by the enzymes
secreted by microorganisms [69]. Furthermore, the copolymerization of
3.1–28.2

8.4–16.9

dodecanedioic acid (DDCA), FDCA and ethylene glycol resulted in series


(J/g)
ΔHm

of fully bio-based poly(ethylene dodecanedioate-2,5-


furandicarboxylate) (PEDF) copolyesters. The weight loss in the PEDF
26.7–107

copolyesters increased linearly with the molar content of the aliphatic


41–115

units [133].
(oC)
Tc

5.2.2. FDCA and succinic acid copolyesters


406–430

430–438

392–401

Another attractive biobased monomer to yield semi-crystalline


Td max

aliphatic polyester is succinic acid, which has been integrated with


(oC)

FDCA during polymerization to synthesize a plethora of novel bio­


polymers [73,134]. In comparison to PBS (E = 680 MPa, σb = 30 MPa,
− 34.3–35.5 121.9–167.6

108.5–168.7

and εb = 130%), PBF has higher E (1.8 GPa) and σb (35 MPa), but lower
20.93–62.19 185–213

εb (2.5%) with reduced biodegradability. Thus, to enhance these prop­


(oC)
Tm

erties, Wu et al. developed poly(butylene succinate-co-


furandicarboxylate) (PBS-co-PBF) having tunable properties ranging
from amorphous polymers with low Tg and very high εb (~600%), to
crystalline polymers with high E (360–1800 MPa) and σb (20–35 MPa).
(oC)
Tg

The higher furanic content (>60%) induced darker pigmentation in the


polymer which might be due to impurities in the FDCA monomer or a
0.62–0.99
1.39–1.9
1–1.42

coloration effect of TBT. The purification using acetic acid significantly


(dL/g)

lightened the color of the PBS-co-PBF [81].


Ƞ

The presence of furanic units (butylene furanoate) showed an overall


increase in Tg, while a negative effect on crystallinity and Tm was
PDI

observed in butylene succinate rich copolyesters. The poor crystalliza­


(-)

tion rate of furanoate segments arises due to asymmetric furan ring and
30600–46900

45600–66800

odd C-atom in the diol groups. Yet, the low melt strength and poor
crystallization rate of PBS-co-PBF (≥60 mol% FDCA) limited the film-
(g/mol)

forming ability via extrusion-blowing [135]. It has been observed that


Mn

the spherulite morphology contributing to the crystalline structure in


homopolyesters is destroyed during copolymerization. However, it can


be restored by annealing (post-treatment). For instance, thermal
annealing of poly(propylene succinate-co-furandicarboxylate) (PPS-co-
(g/mol)

PPF) and poly(neopentyl glycol succinate-co-furandicarboxylate) (PNS-


Mw

co-PNF) enhanced their crystallinity. Moreover, the E and σb values of


TBT (300
Catalysts

Sb2(EG)3
TBT (0.2

TBT (0.2

these copolyesters exceeded the commercial biodegradable packaging


mol% of

mol% of

(0.15%/
DMFD),
Zn(Ac)2
µL/mol

DMFD)

DMFD)

DMFD)
(0.2%/
diacid)

materials, while εb decreased after the treatment. The CO2, O2, and H2O
mol

mol

barrier properties for these novel copolymers were superior to


Polycondensation: 230 ◦ C for

Polycondensation: 220 ◦ C for

Polycondensation: 220 ◦ C for

Polycondensation: 235 ◦ C for


170–200 ◦ C for 3–5 h under

170–200 ◦ C for 3–5 h under


Direct esterification: 190 ◦ C

commercially available bioplastics like PLA and PBA-co-PBT. The


Transesterification: 190 ◦ C

cm3⋅cm/cm2⋅s⋅cmHg.

excellent barrier properties are related to higher steric hindrance from


for 1 h, then heated to
Copolyesters (FDCA Copolyester synthesis

for 3–4 h at 480 rpm

the additional methyl groups decreasing the gas and water vapor
Transesterification:

Transesterification:
5 h under vacuum

adsorption, as well as diffusion and desorption through the polymeric


3–5 h at 5 Pa

3–5 h at 5 Pa

film [72,136].
Since the catalysts used during polymerization reactions have a
210 ◦ C,

direct influence on the optical properties of the polymers, the replace­


4h
N2

N2

10

ment of Ti with Zr-based catalyst by Jacquel et al. caused much less


Note: 1 Barrer = 10−
Table 4 (continued )

coloration in the product during the melt polycondensation of PBS-co-


PBF-PEG (10–60)

PEF-PEG (10–60)

PBF than reported by Wu et al. [81,135]. Another strategy applied for


PBS-co-PBF synthesis was ROP under a milder condition (120–150 ◦ C)
PBF-PEG

than melt polycondensation (180–230 ◦ C). On the downside, the lipase-


mol%)

catalyzed ROP exhibited lower yield (88–93%) and Mw (15–45 kg/mol)


for PBS-co-PBF than what was obtained by a chemocatalytic process

14
S. Pandey et al. European Polymer Journal 160 (2021) 110778

(50–65 kg/mol, 65–91%) [84]. obtained were in a rubbery state at low FDCA(40–60%) which then
Furthermore, several studies have shown that increasing the succinic transformed into fully amorphous structure at higher FDCA content
acid content improved the enzymatic and hydrolytic biodegradability of (60–90%) [143].
copolyesters because of the dissimilarities generated in the crystal lat­ It should be noted that the introduction of sebacic units in PBF
tices [72,136,137]. During lipase action, poly(ethylene succinate-co- reduced the crystallization rate and caused instability in the crystal
furandicarboxylate) (PES-co-PEF) (with 17% ethylene furanoate units) lattice of poly(butylene sebacate-co-furandicarboxylate) (PBSe-co-PBF)
was reported to have 15% weight loss within a month. The higher fur­ because of additional flexibility due to long chain. Increasing the sebacic
anic content negatively affected the weight loss of the polyester due to acid content (5 to 20%) reduced the σb from 51 to 33 MPa, while the εb
high rigidity, hydrophobicity, Tg, Tm, and limited chain mobility hin­ increased from 8.1 to 780%. The decomposition temperature shifted to a
dering the enzymatic attack on the ester groups [137]. The enzymatic higher value with sebacic units addition, as the former is more resistant
biodegradability of PPS-co-PPF was predicted using number average to heat than furan rings [144]. Until now, no study has been conducted
sequence length of aromatic units as the indicator, where they reported to investigate the biodegradability of sebacic acid series polymers with
that a value less than 3 indicated weight loss of around 15% [72]. FDCA units.
Moreover, it has been suggested that lipase-assisted hydrolytic degra­
dation results in better biodegradation of copolymeric chains. For PBS- 5.2.5. Polyglycolic acid
co-PBF, the enzyme-assisted hydrolytic degradation rate in neutral pH Polyglycolic acid (PGA) has a similar structure to lactic acid but
(7.4) was triggered in the presence of lipase and the weight loss was without the methyl group which can successfully induce biodegrad­
reported as 20% in 22 weeks [84]. For PNS-co-PNF, the enzymatic and ability when copolymerized with PBF [145,146]. This flexible como­
hydrolytic degradation was limited in the range of 3 to 13% owing to the nomer can impart structural mobility due to its amorphous nature by
high steric hindrance, crystallinity, and Mw that restricted the water conversion of a rigid plastic into a soft elastomer, thus decreasing the Tg
diffusion in the polymeric chain [136]. Compared to adipic acid, suc­ and crystallinity of the copolymer formed. Compared to PCL and PLA,
cinic acid as a comonomer imparts higher crystallinity in the copo­ PGA-based polyesters have superior barrier properties and are more
lyesters. This is why only 14% weight loss was observed for PPS-co-PPF susceptible to hydrolytic degradation. For instance, enzymatic hydro­
compared to 52% for PBA-co-PBF in alkaline medium. On the other lysis was responsible for a 60% loss in weight for poly(ethylene glycol-
hand, PBS-co-PBF samples (15 mol%) have been found to be compost­ co-butylene furandicarboxylate) (PGA-co-PBF). The same copolyester
able, while according to EN 13432 [135], the time required to reach can also replace PBA-co-PBT owing to its excellent rigidity and barrier
90% biodegradation increased from 71 to 106 days with the addition of properties. The copolymer has lower CO2 (53–118 times), O2 (15 times),
butylene furanoate units (>45 mol%) [69]. and water vapor (6 times) permeability than Ecoflex [146]. The use of
diglycolic acid instead of glycolic acid for copolymerization with PBF
5.2.3. FDCA and lactic acid copolyesters increased the hydrophilic and compostable nature of the polyester
The majority of compostable plastics are composed of PLA having formed. Moreover, the copolyester poly(diglycolate-co-butylene fur­
lactic acid as the monomer being produced by bacterial fermentation of andicarboxylate) (PDG-co-PBF) has a high CO2 gas transmission rate
carbohydrates. The industrial production of PLA is carried through relative to O2, which makes it an interesting candidate for food pack­
polycondensation of lactic acid. However, the ROP of lactic acid dimers aging applications under modified atmosphere packaging (MAP) [145].
(lactides) is gaining attention due to its solvent-free synthesis process
[138]. Attempts have been made to copolymerize PLA with PEF and PBF 5.2.6. Polyethylene glycol
homopolyesters to improve brittleness, crystallization, Tg and, barrier Polyethylene glycol (PEG) is the base for the most common class of
properties. The incorporation of PBF units in PLA backbone improved its linear polymer with a wide variation in Mw ranging from 600 to 20000
tensile and gas barrier properties. The εb, CO2, O2, and water vapor g/mol. The low Mw PEG (1000–6000 g/mol), when copolymerized with
transmission rates of poly(lactic acid-co-butylene furandicarboxylate) PBF, decreases the E and σb of the poly(ethylene glycol-co-butylene
(PLA-co-PBF) increased by 25, 7.4, 6.3, and 2 times respectively, furandicarboxylate) (PEG-co-PBF) copolyester with a significant in­
compared to that of PLA homopolyesters [139]. However, the Tm and crease in εb. On the other hand, larger moieties (10 and 20 kg/mol),
crystallization rate decreased with respect to the homopolyesters, after reduce the εb and increase the E because of their high crystallinity [147].
the addition of lactide units. The copolymerization of PLA and PEF led to As in the case of random copolymers, the presence of the rigid furan ring
the disruption of the original poly-(L-lactic acid) structure yielding severely inhibits or disrupts the crystallization of the PEG block. The
random polymer structures with low Mw ranging from 6900 to 9000 g/ study conducted by Wang et al. (2017a) showed that increasing the PEG
mol. Inevitably, the loss of crystallinity and hydrophilicity of lactide chain length could facilitate the copolyesters to more easily crystallize,
units during copolymerization improved the water absorption and hy­ while the enthalpy of melting of the PEG block and the crystallization of
drolytic degradation rate of poly(lactic acid-co-ethylene fur­ the copolymers also significantly increased [148]. For isothermal crys­
andicarboxylate) PLA-co-PEF (60% in 12 weeks) [140]. Compared to tallization, the introduction of PEG significantly enhanced the crystal­
ethylene furanoate analogues, weight loss of PLA-co-PBF was limited to lization rates as expected, which can be determined from the
35%, while Tg increased due to intermolecular forces provided by the crystallization half-times. Depending on their Mw and molar content,
furan rings [86]. Moreover, the optical properties of PBF-co-PLA formed PEG moiety can induce biodegradability in PEF [149] and PBF
by hot pressing, injection molding, or 3D printing were superior to that [147,150]. The hydrophilicity of the furanic homopolyesters increases
of neat PLA or PBF [141]. with the addition of PEG moiety, thus making the polymer more sus­
ceptible to hydrolytic and soil degradation.
5.2.4. Sebacic acid
Sebacic acid is a naturally occurring long chain dicarboxylic acid that 5.3. Copolymerization of FDCA and cyclic monomers
has been incorporated with other polyesters to improve their crystal
quality, Tg, and structural morphology [142]. The tensile property of 5.3.1. FDCA and caprolactam copolyesters
poly(ethylene sebacate) (PESe) was comparable to that of PE, having Poly(ε-caprolactone) (PCL) is a semi-crystalline polymer produced
values of E = 110 MPa, σb = 20 MPa, and εb = 510%. Meanwhile, poly by ROP of ε-caprolactone. It has exceptional biodegradability but is not
(ethylene sebacate-co-furandicarboxylate) (PESe-co-PEF) was shown to used as a beverage packaging material due to its low Tm (~65 ◦ C), Tg
have similar properties to well-known copolyesters like PBS-co-PBT. (~− 60 ◦ C), and poor elasticity [151]. To improve these thermo­
Increasing the amount of FDCA produced imperfect crystals which mechanical properties, functional copolyesters like poly(ε-caprolactone-
later caused more difficult PESe-co-PEF crystallization. The copolyesters co-butylene furandicarboxylate) (PCL-co-PBF) were synthesized using

15
S. Pandey et al. European Polymer Journal 160 (2021) 110778

Table 5
Summary of FDCA-based biodegradable or compostable copolyesters.
Copolymer Degradation medium pH Temperature Specimen Time Weight loss (%) Reference
(◦ C)

FDCA and adipic acid copolyesters


PEAF Rhizopus oryzae lipase (0.09 mg/mL) and 7.2 37 Films (5 cm2 × 0.4 mm) 22 100 [68]
Pseudomonas cepacia lipase (0.01 mg/mL) days
PBAF Lipase from porcine pancreas (0.1 mg/mL) 7.2 37 Films (10 cm2 × 0.3 28 9–100 [73]
mm) days
PBAF NaH2PO4 (0.2 M)/acetic acid 4 25 Films (15 × 6) 22 1 [73]
weeks
PBAF Sodium acetate/NaOH (0.2 M/0.2 M) 12 25 Films (15 × 6) 22 52 [73]
weeks
PBAF Neutral 7 25 Films (15 × 6) 22 1.4–1.8 [73]
weeks
PBAF Compost (ISO 14855–1:2005 and GB/T ng 58 Films (20 mm × 20 110 >90 [73]
19277.2–2013) mm), 0.0015 g days
PPAF Acidic aqueous solution 4.35 37 – 28 0.69 [75]
days

FDCA and succinic acid copolyesters


PBSF Compost (ISO 14855–1:2005 and EN 13432) – 58 – – – [135]
PESF Rhizopus delemar lipase (0.1 mg/mL) and 7.2 50 Film (5 cm2 × 2 mm) 30 2.5–15 [137]
Pseudomonas cepacia lipase (0.01 mg/mL) days
PBSF NaH2PO4 (0.2 M)/ acetic acid 4 25 Films (15 × 6) 22 1 [69]
weeks
PBSF Sodium acetate/NaOH (0.2 M/0.2 M) 12 25 Films (15 × 6) 22 14 [69]
weeks
PBSF Neutral 7 25 Films (15 × 6) 22 1.4–1.8 [69]
weeks
PBSF Compost (ISO 14855-1:2005 and GB/T 19277.2- – 58 Films (20 mm × 20 110 90 [69]
2013) mm), 0.0015 g days
PBSF Citric acid buffer 2 37 Disks (10 mm dia), 30 10 [84]
20–30 mg days
PBSF Sodium phosphate buffer 7.4 37 Disks (10 mm dia), 30 5 [84]
20–30 mg days
PBSF Sodium phosphate buffer/Lipase porcine pancreas 7.4 37 Disks (10 mm dia), 30 15–20 [84]
(10 mg) 20–30 mg days
PPSF Phosphate buffer solution/Lipase porcine pancreas 7.4 37 Film (2 cm2 × 0.3 mm) 28 5.7–14.8 [72]
(0.1 mg/mL) days
PNSF PBS/CALB (0.1 mg/mL) 7.4 37 Film (0.3 mm thick) 70 3–13 [136]
days
PNSF PBS 7.4 37 Film (0.5 mm thick) 70 <10 [136]
days

FDCA and lactic acid copolyesters


PELF Phosphate buffer solution 6.9 37 Square samples (12–50 12 22–60 [140]
mg) weeks
PBFL Phosphate buffer solution 7 50 – 80 3–35 [86]
days

FDCA and caprolactum copolyesters


PB(CL)F Sodium phosphate buffer 7.4 25 Disk (10 mm dia, 200 40 2 [88]
µm thick, 20–30 mg) days
PB(CL)F Sodium phosphate buffer/Lipase porcine pancreas 7.4 25 Disk (10 mm dia, 200 40 40 [88]
(10 mg) µm thick, 20–30 mg) days
PB(CL)F Phosphate buffer solution 7.4 37 Film (2 × 2 × 0.03 cm3) 8 ~10 [74]
weeks
3
PB(CL)F Phosphate buffer solution/Lipase porcine pancreas 7.4 37 Film (2 × 2 × 0.03 cm ) 8 ~24 [74]
(0.1 mol/L) weeks
3
PB(CL)F Phosphate buffer solution/ CLAB (0.1 mol/L) 7.4 37 Film (2 × 2 × 0.03 cm ) 8 ~34 [74]
weeks
PPe(CL)F Phosphate buffer solution/Pseudomonas cepacia 7.4 37 Square shaped film (5 25 15 [85]
lipase (0.01 mg/mL) & Rhizopus oryzae lipase (0.09 cm2 × 0.4 mm) days
mg/mL)
PHe(CL)F Phosphate buffer solution/Pseudomonas cepacia 7.4 37 Square shaped film (5 25 32 [85]
lipase (0.01 mg/mL) & Rhizopus oryzae lipase (0.09 cm2 × 0.4 mm) days
mg/mL)

FDCA and isosorbide copolyesters


PDIsF Garden soil 6.5 25 Film (20 mm × 10 mm 23 15–16 [77]
× 0.1 mm) weeks
PBImF Sodium phosphate buffer 7.4 37 Disks (10 mm dia, 200 30 20 [87]
µm thick, 20–30 mg) days
PBImF Sodium phosphate buffer/Lipase porcine pancreas 7.4 37 Disks (10 mm dia, 200 30 50 [87]
(10 mg) µm thick, 20–30 mg) days
PICBF(S) Phosphate buffer solution/Lipase porcine pancreas 7.4 – Film (-) 29 13.2 [155]
(0.1 mol/L) days

FDCA and PGA copolyesters


(continued on next page)

16
S. Pandey et al. European Polymer Journal 160 (2021) 110778

Table 5 (continued )
Copolymer Degradation medium pH Temperature Specimen Time Weight loss (%) Reference
(◦ C)

PBF-DGA Compost – 60 Film (20 × 40 mm, 0.2 62 40 [145]


mm thick) days
PBFGA Phosphate buffered saline 7.4 37 Film (0.3 mm thick) 35 <1.5 [146]
days
PBFGA Phosphate buffered saline/Lipase porcine pancreas 7.4 37 Film (0.3 mm thick) 70 35–6- [146]
(0.1 mg/mL) days

FDCA and PEF copolyesters


PBF-PEG PBS solution 7.4 37 – 5 44 [150]
weeks
PBF-PEG NaOH solution (0.01 mol/L) – 37 – 5 64.6 [150]
weeks
PBF-PEG Pure water 7 37 – 49 < 7 for low Mw, [147]
days 11–22 for low Mw
PBF-PEG Phosphate buffer saline 7.2–7.4 37 – 49 18 for low Mw [147]
days
PBF-PEG NaOH solution (0.0001 mol/L) 10 37 – 49 40 for high Mw [147]
days
PBF-PEG NaOH solution (0.01 mol/L) 12 37 – 24 h 100 [147]
PEF-PEG Phosphate buffered solution 7.2 37 Film: 1 cm × 3 cm × 100 15.3 [149]
(0.1–0.3) mm days

cyclic oligomers of FDCA and ε-caprolactone by enzyme and chemical high Mn copolymers poly(isosorbide-co-butylene furandicarboxylate)
assisted ROP. The Tg of PCL-co-PBF increased with the addition of rigid (PIsBF) ranging from 18400 to 37499 g/mol [76]. Due to these prop­
butylene furanoate units due to a reduction in chain mobility. Compared erties, isohexides have been proposed as the most suitable candidate to
to PBF homopolyesters, PCL undergoes faster crystallization during replace toxic bisphenol-A as the hardening agent in food and beverage
thermal processing. Therefore, copolymerization resulted in a slow packaging. Despite these advantages, the insertion of isohexides in the
crystallization rate for PCL-co-PBF, which further deteriorated with polymeric chain often results in chain irregularity and disrupts the
increasing butylene furanoate units. Regardless of the insensitivity of crystalline structure, thus restricting their application at the industrial
these copolyesters to hydrolytic degradation under neutral conditions, scale [153]. A relevant example is presented by Gomes et al. where they
they degrade well in presence of enzymes and higher caprolactam units synthesized fully amorphous homopolyesters by solution polymeriza­
[88]. The slow degradation for PCL-co-PBF occurred due to the hydro­ tion of FDCA with Is and Id [91].
phobicity of the copolyesters limiting hydrolysis under ambient condi­ The presence of linear aliphatic diols with numerous methylene
tions. Moreover, the protein structure of the enzyme has a significant groups in the furanic chain has also been shown to control the crystal­
impact when it comes to breaking ester bonds. A study revealed that lization ability with a Tm suitable for thermoplastic processing. FDCA-
Candida antarctica lipase B (CALB) is more suitable for enzymatic hy­ based homopolyesters comprising of diols with 8, 9, 10, and 12 meth­
drolysis than lipase from porcine pancreas, as it has the appropriate ylene groups were reported to display a significant degree of crystal­
protein structure needed for the ester bond breakage [74]. linity [154]. In order to address the issue of poor crystallinity, Is and
Furthermore, two novel thermally improved copolyesters, poly 1,10-decanediol were used as comonomers to produce a novel copo­
(ε-caprolactone-co-pentylene furandicarboxylate) (PCL-co-PPeF) and lyester: poly(decamethylene-co-isosorbide furandicarboxylate) (PDIsF).
poly(ε-caprolactone-co-hexamethylene furandicarboxylate) (PCL-co- The presence of Is units effectively increased the Tg and delayed the
PHeF) consisting of higher boiling point diols (1,5-pentanediol and 1,6- onset of decomposition. The asymmetric secondary hydroxyl groups in Is
hexanediol) as comonomers, were synthesized by ROP using stannous imposed large steric hindrance for soil microorganisms to act on the
octoate as the catalyst. Both the copolymers exhibited low Mw because ester linkages resulting in delayed biodegradation. As a consequence,
of the competitive side reaction and thermal degradation during harsh the Is units negatively affected the weight loss in PDIsF chains [77].
conditions of melt polycondensation. The increase in caprolactam units Moreover, the stability of the secondary hydroxyl groups present in
negatively affected the thermal stability of the polymeric chain, thus isohexides leads to low reactivity during polymerization reactions and
yielding low Tm, Tg, and ΔHm for these copolyesters. The enzymatic hinders industrial-scale production of such FDCA-based copolyesters.
weight loss of PCL-co-PHeF was faster than PCL-co-PPeF, achieving a Ouyang et al. described a semi-continuous method for large-scale pro­
maximum of 32% vs. 15% respectively, within 25 days. This enzyme duction of poly(ester carbonates) having FDCA and Is as the co­
facilitated weight loss was associated with the lower Tm, Tg, and crys­ monomers, where the oligomers of Is and dimethyl carbonate were
tallinity of the copolyester. However, the lower Mw played a key role in reacted with DMFD and 1,4-butanediol to obtain poly(isosorbide
disintegration during hydrolysis [85]. Furthermore, the brittleness of carbonate-co-butylene furandicarboxylate) (PIsCBF). These PIsCBF
the poly(isosorbide 2,5-furandicarboxylate) (PIsF) homopolyesters was copolyesters were characterized by high Mw and Tg, but the process did
modified by copolymerization with caprolactam leading to increased not improve the crystallinity of the polyesters. This was coupled with
mobility of the chain resulting in amorphous poly(isosorbide 2,5-fur­ poor O2 and CO2 barrier properties of the PIsCBF copolyesters due to
andicarboxylate-co-ε-caprolactone) (PIs(CL)F) copolyesters. Increasing their amorphous structure [155]. Another study compared the behavior
the number of caprolactam units improved the εb of PIs(CL)F copo­ of three isohexides (Is, Im, and Id) produced through cyclic oligomeri­
lyesters from 23 to 480% [152]. zation. It was reported that cyclic oligomers of Im were more stable than
those of Is and Id. The higher reactivity of Is and Id was attributed to the
5.3.2. FDCA and isohexides copolyesters instability in spatial orientation of the secondary hydroxyl groups at the
Isohexides constitute a group of renewable bicyclic rigid diols pro­ endo and exo positions. The addition of Im units had a positive effect on
duced by dehydration of sugar alcohols, thus yielding three isomers Tg, as well as the hydrolytic degradation of PBF (20% in 30 days), which
namely Isosorbide (Is), Isomannide (Im), and Isoidide (Id). They impart further accelerated up to 50% in presence of lipase [87].
excellent structural rigidity and thermal stability to the polymeric chain.
Incorporation of Is in PBF improved the Tg of the homopolyester yielding

17
S. Pandey et al. European Polymer Journal 160 (2021) 110778

6. Conclusion [8] K. Dietrich, M.-J. Dumont, L.F. Del Rio, V. Orsat, Producing PHAs in the
bioeconomy — Towards a sustainable bioplastic, Sustain. Prod. Consum. 9 (2017)
58–70, https://doi.org/10.1016/j.spc.2016.09.001.
Although significant progress has been made on the preparation of [9] S. Chatti, S.M. Weidner, A. Fildier, H.R. Kricheldorf, Copolyesters of isosorbide,
HMF and biomass-derived FDCA, there are still several challenges that succinic acid, and isophthalic acid: Biodegradable, high Tg engineering plastics,
must be addressed to reach its sustainable production at a reasonable J. Polym. Sci., Part A: Polym. Chem. 51 (11) (2013) 2464–2471, https://doi.org/
10.1002/pola.v51.1110.1002/pola.26635.
cost. Firstly, the catalyst selectivity and recovery during HMF oxidation [10] T.R. Boussie, E.L. Dias, Z.M. Fresco, V.J. Murphy, Production of adipic acid and
is a critical factor that should be considered. While noble metal catalysts derivatives from carbohydrate-containing materials, Google Patents (2013).
have exhibited excellent catalytic performance in FDCA production, the [11] S.K. Burgess, J.E. Leisen, B.E. Kraftschik, C.R. Mubarak, R.M. Kriegel, W.J. Koros,
Chain mobility, thermal, and mechanical properties of poly (ethylene furanoate)
development of efficient, low-cost, and multi-functional green catalysts compared to poly (ethylene terephthalate), Macromolecules 47 (4) (2014)
in a base-free environment still needs further investigation. In addition, 1383–1391, https://doi.org/10.1021/ma5000199.
the feedstocks used for FDCA synthesis have a major impact on the cost [12] S.K. Burgess, O. Karvan, J. Johnson, R.M. Kriegel, W.J. Koros, Oxygen sorption
and transport in amorphous poly (ethylene furanoate), Polymer 55 (18) (2014)
of FDCA production. Therefore, the use of food and lignocellulosic 4748–4756, https://doi.org/10.1016/j.polymer.2014.07.041.
wastes as potential low-cost substrates should be further exploited for [13] A. Eerhart, A. Faaij, M.K. Patel, Replacing fossil based PET with biobased PEF;
the development of such processes. To achieve large-scale FDCA pro­ process analysis, energy and GHG balance, Energy Environ. Sci. 5 (4) (2012)
6407–6422, https://doi.org/10.1039/C2EE02480B.
duction, the design should be based on all the technical, economical, and [14] I.K.M. Yu, D.C.W. Tsang, A.C.K. Yip, S.S. Chen, Y.S. Ok, C.S. Poon, Valorization of
sustainable challenges. Moreover, copolymerization could be used as a starchy, cellulosic, and sugary food waste into hydroxymethylfurfural by one-pot
way to create FDCA-based polymers for food and beverage packaging catalysis, Chemosphere 184 (2017) 1099–1107, https://doi.org/10.1016/j.
chemosphere.2017.06.095.
with the required physicochemical properties. Simultaneously, prob­
[15] I.K.M. Yu, D.C.W. Tsang, A.C.K. Yip, S.S. Chen, L. Wang, Y.S. Ok, C.S. Poon,
lems related to coloration, high cost, and poor biodegradability can be Catalytic valorization of starch-rich food waste into hydroxymethylfurfural
solved using copolymerization. It is important not to overlook the (HMF): Controlling relative kinetics for high productivity, Bioresour. Technol.
method of bioplastics disposal, their impact on microplastic formation in 237 (2017) 222–230, https://doi.org/10.1016/j.biortech.2017.01.017.
[16] O. Yemiş, G. Mazza, Acid-catalyzed conversion of xylose, xylan and straw into
the environment and marine life. Therefore, the life cycle assessment furfural by microwave-assisted reaction, Bioresour. Technol. 102 (15) (2011)
studies of these polymeric materials are important before bringing them 7371–7378, https://doi.org/10.1016/j.biortech.2011.04.050.
into the industrial chain. [17] S. Dutta, S. De, M.I. Alam, M.M. Abu-Omar, B. Saha, Direct conversion of
cellulose and lignocellulosic biomass into chemicals and biofuel with metal
chloride catalysts, J. Catal. 288 (2012) 8–15, https://doi.org/10.1016/j.
7. Funding Sources jcat.2011.12.017.
[18] Y.H. Seo, J.-I. Han, Direct conversion from Jerusalem artichoke to
hydroxymethylfurfural (HMF) using the Fenton reaction, Food Chem. 151 (2014)
This study was financially supported by the Ministère de l’Economie 207–211, https://doi.org/10.1016/j.foodchem.2013.11.067.
et de l’Innovation (MEI), through “Transform Action, programme [19] I.K.M. Yu, D.C.W. Tsang, A.C.K. Yip, S.S. Chen, Y.S. Ok, C.S. Poon, Valorization of
d’appui au développement des secteurs stratégiques et des créneaux food waste into hydroxymethylfurfural: Dual role of metal ions in successive
conversion steps, Bioresour. Technol. 219 (2016) 338–347, https://doi.org/
d’excellence en Montérégie”, Agriculture and Agri-Food Canada, 10.1016/j.biortech.2016.08.002.
through “Growing Forward 2′′ , the Ministère de l’Agriculture, des [20] Guillermo Alberto Portillo Perez, Marie-Josée Dumont, Polyvinyl sulfonated
Pêcheries, et de l’Alimentation du Québec (MAPAQ) through the Con­ catalyst and the effect of sulfonic sites on the dehydration of carbohydrates,
Chem. Eng. J. 419 (2021) 129573, https://doi.org/10.1016/j.cej.2021.129573.
sortium de Recherche Innovation Transformation Alimentaire (RITA)
[21] Guillermo Portillo Perez, Marie-Josée Dumont, Production of HMF in high yield
and also by Québec food industries through the Conseil de la Trans­ using a low cost and recyclable carbonaceous catalyst, Chem. Eng. J. 382 (2020)
formation Alimentaire du Québec (CTAQ). 122766, https://doi.org/10.1016/j.cej.2019.122766.
[22] M. Sajid, X. Zhao, D. Liu, Production of 2,5-furandicarboxylic acid (FDCA) from
5-hydroxymethylfurfural (HMF): recent progress focusing on the chemical-
Declaration of Competing Interest catalytic routes, Green Chem. 20 (24) (2018) 5427–5453, https://doi.org/
10.1039/c8gc02680g.
The authors declare that they have no known competing financial [23] Guangshun Yi, Siew Ping Teong, Yugen Zhang, The direct conversion of sugars
into 2,5-furandicarboxylic acid in a triphasic system, ChemSusChem 8 (7) (2015)
interests or personal relationships that could have appeared to influence 1151–1155, https://doi.org/10.1002/cssc.v8.710.1002/cssc.201500118.
the work reported in this paper. [24] Luana Dessbesell, Sadra Souzanchi, Kasanneni Tirumala Venkateswara Rao,
Alejandro Alarcon Carrillo, Devon Bekker, Kristen Amanda Hall, Katherine
Mary Lawrence, Carolyn Lindsay Judge Tait, Chunbao (Charles) Xu, Production
References of 2,5-furandicarboxylic acid (FDCA) from starch, glucose, or high-fructose corn
syrup: techno-economic analysis, Biofuels, Bioprod. Biorefin. 13 (5) (2019)
[1] B. Imre, L. García, D. Puglia, F. Vilaplana, Reactive compatibilization of plant 1234–1245, https://doi.org/10.1002/bbb.v13.510.1002/bbb.2014.
polysaccharides and biobased polymers: Review on current strategies, [25] G. Yi, S.P. Teong, X. Li, Y. Zhang, Purification of biomass-derived 5-hydroxyme­
expectations and reality, Carbohydr. Polym. 209 (2019) 20–37, https://doi.org/ thylfurfural and its catalytic conversion to 2, 5-furandicarboxylic acid,
10.1016/j.carbpol.2018.12.082. ChemSusChem 7 (8) (2014) 2131–2135, https://doi.org/10.1002/
[2] F. Zheng, L. Chen, P. Zhang, J. Zhou, X. Lu, W. Tian, Carbohydrate polymers cssc.201402105.
exhibit great potential as effective elicitors in organic agriculture: A review, [26] M.L. Ribeiro, U. Schuchardt, Cooperative effect of cobalt acetylacetonate and
Carbohydr. Polym. 230 (2020) 115637, https://doi.org/10.1016/j. silica in the catalytic cyclization and oxidation of fructose to 2, 5-furandicar­
carbpol.2019.115637. boxylic acid, Catal. Commun. 4 (2) (2003) 83–86, https://doi.org/10.1016/
[3] A. Gandini, Polymers from renewable resources: a challenge for the future of S1566-7367(02)00261-3.
macromolecular materials, Macromolecules 41 (24) (2008) 9491–9504, https:// [27] M. Kröger, U. Prüße, K.-D. Vorlop, A new approach for the production of 2, 5-
doi.org/10.1021/ma801735u. furandicarboxylic acid by in situ oxidation of 5-hydroxymethylfurfural starting
[4] C. Vilela, A.F. Sousa, A.C. Fonseca, A.C. Serra, J.F. Coelho, C.S. Freire, A. from fructose, Top. Catal. 13 (3) (2000) 237–242, https://doi.org/10.1023/A:
J. Silvestre, The quest for sustainable polyesters–insights into the future, Polym. 1009017929727.
Chem. 5 (9) (2014) 3119–3141, https://doi.org/10.1039/C3PY01213A. [28] R. Wojcieszak, I. Itabaiana, Engineering the future: Perspectives in the 2,5-fur­
[5] K.M. Nampoothiri, N.R. Nair, R.P. John, An overview of the recent developments andicarboxylic acid synthesis, Catal. Today 354 (2020) 211–217, https://doi.org/
in polylactide (PLA) research, Bioresour. Technol. 101 (22) (2010) 8493–8501, 10.1016/j.cattod.2019.05.071.
https://doi.org/10.1016/j.biortech.2010.05.092. [29] Avantium. Avantium annual report, 2019. https://www.avantium.com (assesed
[6] H.-M. Müller, D. Seebach, Poly (hydroxyalkanoates): a fifth class of on 7 July, 2020).
physiologically important organic biopolymers? Angew. Chem. Int. Ed. Engl. 32 [30] A.H. Motagamwala, W. Won, C. Sener, D.M. Alonso, C.T. Maravelias, J.
(4) (1993) 477–502, https://doi.org/10.1002/(ISSN)1521-377310.1002/anie. A. Dumesic, Toward biomass-derived renewable plastics: Production of 2,5-fur­
v32:410.1002/anie.199304771. andicarboxylic acid from fructose, Sci. Adv. 4 (1) (2018), https://doi.org/
[7] N. Jacquel, F. Freyermouth, F. Fenouillot, A. Rousseau, J.P. Pascault, P. Fuertes, 10.1126/sciadv.aap9722 eaap9722.
R. Saint-Loup, Synthesis and properties of poly (butylene succinate): Efficiency of [31] A.H. Motagamwala, K.F. Huang, C.T. Maravelias, J.A. Dumesic, Solvent system
different transesterification catalysts, J. Polym. Sci., Part A: Polym. Chem. 49 (24) for effective near-term production of hydroxymethylfurfural (HMF) with
(2011) 5301–5312, https://doi.org/10.1002/pola.v49.2410.1002/pola.25009. potential for long-term process improvement, Energy Environ. Sci. 12 (7) (2019)
2212–2222, https://doi.org/10.1039/c9ee00447e.

18
S. Pandey et al. European Polymer Journal 160 (2021) 110778

[32] I.K.M. Yu, D.C.W. Tsang, A.C.K. Yip, A.J. Hunt, J. Sherwood, J. Shang, H. Song, Y. [54] R. Wojcieszak, C.P. Ferraz, J. Sha, S. Houda, L.M. Rossi, S. Paul, Advances in base-
S. Ok, C.S. Poon, Propylene carbonate and γ-valerolactone as green solvents free oxidation of bio-based compounds on supported gold catalysts, Catalysts 7
enhance Sn(iv)-catalysed hydroxymethylfurfural (HMF) production from bread (11) (2017) 352, https://doi.org/10.3390/CATAL7110352.
waste, Green Chem. 20 (9) (2018) 2064–2074, https://doi.org/10.1039/ [55] N.K. Gupta, S. Nishimura, A. Takagaki, K. Ebitani, Hydrotalcite-supported gold-
c8gc00358k. nanoparticle-catalyzed highly efficient base-free aqueous oxidation of 5-hydrox­
[33] Corbian, 2020. http://www.corbion.com (assesed on 9 July). ymethylfurfural into 2,5-furandicarboxylic acid under atmospheric oxygen
[34] F.R. Duke, T.W. Haas, The homogeneous base-catalyzed decomposition of pressure, Green Chem. 13 (4) (2011) 824–827, https://doi.org/10.1039/
hydrogen peroxide1, J. Phys. Chem. 65 (2) (1961) 304–306, https://doi.org/ c0gc00911c.
10.1021/j100820a028. [56] D.K. Mishra, H.J. Lee, J. Kim, H.S. Lee, J.K. Cho, Y.W. Suh, Y. Yi, Y.J. Kim,
[35] W. Partenheimer, V.V. Grushin, Synthesis of 2, 5-Diformylfuran and Furan-2, 5- MnCo2O4 spinel supported ruthenium catalyst for air-oxidation of HMF to FDCA
Dicarboxylic Acid by Catalytic Air-Oxidation of 5-Hydroxymethylfurfural. under aqueous phase and base-free conditions, Green Chem. 19 (7) (2017)
Unexpectedly Selective Aerobic Oxidation of Benzyl Alcohol to Benzaldehyde 1619–1623, https://doi.org/10.1039/c7gc00027h.
with Metal= Bromide Catalysts, Adv. Synth. Catal. 343 (1) (2001) 102–111, [57] B.N. Zope, S.E. Davis, R.J. Davis, Influence of reaction conditions on diacid
https://doi.org/10.1002/1615-4169(20010129)343:1<102::AID-ADSC102>3.0. formation during Au-catalyzed oxidation of glycerol and hydroxymethylfurfural,
CO;2-Q. Top. Catal. 55 (1–2) (2012) 24–32, https://doi.org/10.1007/s11244-012-9777-3.
[36] Zehui Zhang, Kejian Deng, Recent advances in the catalytic synthesis of 2, 5- [58] X. Wan, C. Zhou, J. Chen, W. Deng, Q. Zhang, Y. Yang, Y. Wang, Base-free aerobic
furandicarboxylic acid and its derivatives, ACS Catal. 5 (11) (2015) 6529–6544, oxidation of 5-hydroxymethyl-furfural to 2, 5-furandicarboxylic acid in water
https://doi.org/10.1021/acscatal.5b01491. catalyzed by functionalized carbon nanotube-supported Au–Pd alloy
[37] R.O. Rajesh, T.K. Godan, R. Sindhu, A. Pandey, P. Binod, Bioengineering nanoparticles, ACS Catal. 4 (7) (2014) 2175–2185, https://doi.org/10.1021/
advancements, innovations and challenges on green synthesis of 2, 5-furan cs5003096.
dicarboxylic acid, Bioengineered 11 (1) (2020) 19–38, https://doi.org/10.1080/ [59] D. Yan, J. Xin, Q. Zhao, K. Gao, X. Lu, G. Wang, S. Zhang, Fe–Zr–O catalyzed base-
21655979.2019.1700093. free aerobic oxidation of 5-HMF to 2,5-FDCA as a bio-based polyester monomer,
[38] P. Verdeguer, N. Merat, A. Gaset, Oxydation catalytique du HMF en acide 2, 5- Catal. Sci. Technol. 8 (1) (2018) 164–175, https://doi.org/10.1039/c7cy01704a.
furane dicarboxylique, J. Mol. Catal. 85 (3) (1993) 327–344, https://doi.org/ [60] D. Yan, J. Xin, C. Shi, X. Lu, L. Ni, G. Wang, S. Zhang, Base-free conversion of 5-
10.1016/0304-5102(93)80059-4. hydroxymethylfurfural to 2,5-furandicarboxylic acid in ionic liquids, Chem. Eng.
[39] H.A. Rass, N. Essayem, M. Besson, Selective aqueous phase oxidation of 5- J. 323 (2017) 473–482, https://doi.org/10.1016/j.cej.2017.04.021.
hydroxymethylfurfural to 2, 5-furandicarboxylic acid over Pt/C catalysts: [61] Ruru Chen, Jiayu Xin, Dongxia Yan, Huixian Dong, Xingmei Lu, Suojiang Zhang,
influence of the base and effect of bismuth promotion, Green Chem. 15 (8) (2013) Highly efficient oxidation of 5-hydroxymethylfurfural to 2, 5-furandicarboxylic
2240–2251, https://doi.org/10.1039/C3GC40727F. acid with heteropoly acids and ionic liquids, ChemSusChem 12 (12) (2019)
[40] S.E. Davis, L.R. Houk, E.C. Tamargo, A.K. Datye, R.J. Davis, Oxidation of 5- 2715–2724, https://doi.org/10.1002/cssc.v12.1210.1002/cssc.201900651.
hydroxymethylfurfural over supported Pt, Pd and Au catalysts, Catal. Today 160 [62] A.B. Gawade, A.V. Nakhate, G.D. Yadav, Selective synthesis of 2, 5-furandicar­
(1) (2011) 55–60, https://doi.org/10.1016/j.cattod.2010.06.004. boxylic acid by oxidation of 5-hydroxymethylfurfural over MnFe2O4 catalyst,
[41] Z. Miao, T. Wu, J. Li, T. Yi, Y. Zhang, X. Yang, Aerobic oxidation of 5-hydroxy­ Catal. Today 309 (2018) 119–125, https://doi.org/10.1016/j.
methylfurfural (HMF) effectively catalyzed by a Ce 0.8 Bi 0.2 O 2− δ supported Pt cattod.2017.08.061.
catalyst at room temperature, RSC Adv. 5 (26) (2015) 19823–19829, https://doi. [63] S. Wang, Z. Zhang, B. Liu, Catalytic conversion of fructose and 5-hydroxyme­
org/10.1039/C4RA16968A. thylfurfural into 2,5-furandicarboxylic acid over a recyclable Fe3O4–CoOx
[42] S. Siankevich, G. Savoglidis, Z. Fei, G. Laurenczy, D.T. Alexander, N. Yan, P. magnetite nanocatalyst, ACS Sustain. Chem. Eng. 3 (3) (2015) 406–412, https://
J. Dyson, A novel platinum nanocatalyst for the oxidation of 5-Hydroxymethyl­ doi.org/10.1021/sc500702q.
furfural into 2, 5-Furandicarboxylic acid under mild conditions, J. Catal. 315 [64] W. Partenheimer, V.V. Grushin, Synthesis of 2, 5-diformylfuran and furan-2,5-
(2014) 67–74, https://doi.org/10.1016/j.jcat.2014.04.011. dicarboxylic acid by catalytic air-oxidation of 5-hydroxymethylfurfural.
[43] B. Siyo, M. Schneider, M.-M. Pohl, P. Langer, N. Steinfeldt, Synthesis, Unexpectedly selective aerobic oxidation of benzyl alcohol to benzaldehyde with
characterization, and application of PVP-Pd NP in the aerobic oxidation of 5- metal bromide catalysts, Adv. Synth. Catal. 343 (1) (2001) 102–111, https://doi.
hydroxymethylfurfural (HMF), Catal. Lett. 144 (3) (2014) 498–506, https://doi. org/10.1002/1615-4169(20010129)343:13.0.co;2-q.
org/10.1007/s10562-013-1186-0. [65] M. Kröger, U. Prüße, K.D. Vorlop, A new approach for the production of 2, 5-
[44] S. Albonetti, A. Lolli, V. Morandi, A. Migliori, C. Lucarelli, F. Cavani, Conversion furandicarboxylic acid by in situ oxidation of 5-hydroxymethylfurfural starting
of 5-hydroxymethylfurfural to 2, 5-furandicarboxylic acid over Au-based from fructose, Top. Catal. 13 (2000) 237–242, https://doi.org/10.1023/A:
catalysts: Optimization of active phase and metal–support interaction, Appl. 1009017929727.
Catal. B 163 (2015) 520–530, https://doi.org/10.1016/j.apcatb.2014.08.026. [66] M.L. Ribeiro, U. Schuchardt, Cooperative effect of cobalt acetylacetonate and
[45] Yury Y. Gorbanev, Søren Kegnæs, Anders Riisager, Effect of support in silica in the catalytic cyclization and oxidation of fructose to 2,5-furandicarbox­
heterogeneous ruthenium catalysts used for the selective aerobic oxidation of ylic acid, Catal. Commun. 4 (2) (2003) 83–86, https://doi.org/10.1016/S1566-
HMF in water, Top. Catal. 54 (16-18) (2011) 1318–1324, https://doi.org/ 7367(02)00261-3.
10.1007/s11244-011-9754-2. [67] Zehui Zhang, Judun Zhen, Bing Liu, Kangle Lv, Kejian Deng, Selective aerobic
[46] Y.Y. Gorbanev, S. Kegnæs, A. Riisager, Selective aerobic oxidation of 5-hydrox­ oxidation of the biomass-derived precursor 5-hydroxymethylfurfural to 2, 5-fur­
ymethylfurfural in water over solid ruthenium hydroxide catalysts with andicarboxylic acid under mild conditions over a magnetic palladium
magnesium-based supports, Catal. Lett. 141 (12) (2011) 1752–1760, https://doi. nanocatalyst, Green Chem. 17 (2) (2015) 1308–1317, https://doi.org/10.1039/
org/10.1007/s10562-011-0707-y. C4GC01833H.
[47] C. Moreau, R. Durand, S. Razigade, J. Duhamet, P. Faugeras, P. Rivalier, P. Ros, [68] L. Papadopoulos, A. Magaziotis, M. Nerantzaki, Z. Terzopoulou, G.
G. Avignon, Dehydration of fructose to 5-hydroxymethylfurfural over H- Z. Papageorgiou, D.N. Bikiaris, Synthesis and characterization of novel poly
mordenites, Appl. Catal. A 145 (1) (1996) 211–224, https://doi.org/10.1016/ (ethylene furanoate-co-adipate) random copolyesters with enhanced
0926-860X(96)00136-6. biodegradability, Polym. Degrad. Stab. 156 (2018) 32–42, https://doi.org/
[48] Michael Jerry Antal, William S.L. Mok, Geoffrey N. Richards, Mechanism of 10.1016/j.polymdegradstab.2018.08.002.
formation of 5-(hydroxymethyl)-2-furaldehyde from d-fructose and sucrose, [69] S.B. Peng, L.B. Wu, B.G. Li, P. Dubois, Hydrolytic and compost degradation of
Carbohydr. Res. 199 (1) (1990) 91–109, https://doi.org/10.1016/0008-6215(90) biobased PBSF and PBAF copolyesters with 40–60 mol% BF unit, Polym. Degrad.
84096-D. Stab. 146 (2017) 223–228, https://doi.org/10.1016/j.
[49] Z. Zhang, J. Zhen, B. Liu, K. Lv, K. Deng, Selective aerobic oxidation of the polymdegradstab.2017.07.016.
biomass-derived precursor 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid [70] Y.N. Huang, W.D. Su, C.B. Lee, On the Weissenberg effect of turbulence, Theor.
under mild conditions over a magnetic palladium nanocatalyst, Green Chem. 17 Appl. Mech. Lett. 9 (4) (2019) 236–245, https://doi.org/10.1016/j.
(2) (2015) 1308–1317, https://doi.org/10.1039/c4gc01833h. taml.2019.03.004.
[50] A. Karich, S. Kleeberg, R. Ullrich, M. Hofrichter, Enzymatic preparation of 2,5- [71] Y.K. Endah, S.H. Han, J.H. Kim, N.K. Kim, W.N. Kim, H.S. Lee, H. Lee, Solid-state
furandicarboxylic acid (FDCA)—A substitute of terephthalic acid—By the Joined polymerization and characterization of a copolyamide based on adipic acid, 1,4-
Action of Three Fungal Enzymes, Microorganisms 6 (1) (2018) 5, https://doi.org/ butanediamine, and 2,5-furandicarboxylic acid, J. Appl. Polym. Sci. 133 (18)
10.3390/microorganisms6010005. (2016). DOI: ARTN 43391.
[51] B. Saha, S. Dutta, M.M. Abu-Omar, Aerobic oxidation of 5-hydroxylmethylfurfu­ [72] H. Hua, R.Y. Zhang, J.G. Wang, W.B. Ying, J. Zhu, Fully bio-based poly(propylene
ral with homogeneous and nanoparticulate catalysts, Catal. Sci. Technol. 2 (1) succinate-co-propylene furandicarboxylate) copolyesters with proper mechanical,
(2012) 79–81, https://doi.org/10.1039/C1CY00321F. degradation and barrier properties for green packaging applications, Eur. Polym.
[52] Onofre Casanova, Sara Iborra, Avelino Corma, Biomass into chemicals: aerobic J. 102 (2018) 101–110, https://doi.org/10.1016/j.eurpolymj.2018.03.009.
oxidation of 5-hydroxymethyl-2-furfural into 2, 5-furandicarboxylic acid with [73] W.D. Zhou, X.W. Wang, B. Yang, Y. Xu, W. Zhang, Y.J. Zhang, J.H. Ji, Synthesis,
gold nanoparticle catalysts, ChemSusChem: Chem. Sustain. Energy Mater. 2 (12) physical properties and enzymatic degradation of bio-based poly(butylene
(2009) 1138–1144, https://doi.org/10.1002/cssc.v2:1210.1002/ adipate-co-butylene furandicarboxylate) copolyesters, Polym. Degrad. Stab. 98
cssc.200900137. (11) (2013) 2177–2183, https://doi.org/10.1016/j.
[53] S.E. Davis, B.N. Zope, R.J. Davis, On the mechanism of selective oxidation of 5- polymdegradstab.2013.08.025.
hydroxymethylfurfural to 2, 5-furandicarboxylic acid over supported Pt and Au [74] Han Hu, Ruoyu Zhang, Wu Bin Ying, Zhengyang Kong, Kai Wang,
catalysts, Green Chem. 14 (1) (2012) 143–147, https://doi.org/10.1039/ Jinggang Wang, Jin Zhu, Biodegradable elastomer from 2,5-furandicarboxylic
C1GC16074E. acid and ε-caprolactone: Effect of crystallization on elasticity, ACS Sustain. Chem.
Eng. 7 (21) (2019) 17778–17788, https://doi.org/10.1021/
acssuschemeng.9b0421010.1021/acssuschemeng.9b04210.s001.

19
S. Pandey et al. European Polymer Journal 160 (2021) 110778

[75] S. Hbaieb, W. Kammoun, C. Delaite, M. Abid, S. Abid, R. El Gharbi, New [96] V. Tsanaktsis, Z. Terzopoulou, M. Nerantzaki, G.Z. Papageorgiou, D.N. Bikiaris,
copolyesters containing aliphatic and bio-based furanic units by bulk New poly(pentylene furanoate) and poly(heptylene furanoate) sustainable
copolycondensation, J. Macromol. Sci. Part A-Pure Appl. Chem. 52 (5) (2015) polyesters from diols with odd methylene groups, Mater. Lett. 178 (2016) 64–67,
365–373, https://doi.org/10.1080/10601325.2015.1018807. https://doi.org/10.1016/j.matlet.2016.04.183.
[76] X.S. Wang, Q.Y. Wang, S.Y. Liu, G.Y. Wang, Synthesis and characterization of [97] Hongzhou Xie, Linbo Wu, Bo-Geng Li, Philippe Dubois, Biobased poly(ethylene-
poly(isosorbide-co-butylene 2,5-furandicarboxylate) copolyesters, Eur. Polym. J. co-hexamethylene 2,5-furandicarboxylate) (PEHF) copolyesters with superior
115 (2019) 70–75, https://doi.org/10.1016/j.eurpolymj.2019.03.025. tensile properties, Ind. Eng. Chem. Res. 57 (39) (2018) 13094–13102, https://
[77] Yosra Chebbi, Nejib Kasmi, Mustapha Majdoub, Pierfrancesco Cerruti, doi.org/10.1021/acs.iecr.8b03204.
Gennaro Scarinzi, Mario Malinconico, Giovanni Dal Poggetto, George [98] S. Hong, K.D. Min, B.U. Nam, O.O. Park, High molecular weight bio furan-based
Z. Papageorgiou, Dimitrios N. Bikiaris, Synthesis, characterization, and co-polyesters for food packaging applications: synthesis, characterization and
biodegradability of novel fully biobased poly(decamethylene-co-isosorbide 2,5- solid-state polymerization, Green Chem. 18 (19) (2016) 5142–5150, https://doi.
furandicarboxylate) copolyesters with enhanced mechanical properties, ACS org/10.1039/c6gc01060a.
Sustain. Chem. Eng. 7 (5) (2019) 5501–5514, https://doi.org/10.1021/ [99] S.K. Burgess, R.M. Kriegel, W.J. Koros, Carbon dioxide sorption and transport in
acssuschemeng.8b0679610.1021/acssuschemeng.8b06796.s001. amorphous poly(ethylene furanoate), Macromolecules 48 (7) (2015) 2184–2193,
[78] J.H. Zhu, J.L. Cai, W.C. Xie, P.H. Chen, M. Gazzano, M. Scandola, R.A. Gross, Poly https://doi.org/10.1021/acs.macromol.5b00333.
(butylene 2,5-furan dicarboxylate), a biobased alternative to PBT: Synthesis, [100] S.K. Burgess, D.S. Mikkilineni, D.B. Yu, D.J. Kim, C.R. Mubarak, R.M. Kriegel, W.
physical properties, and crystal structure, Macromolecules 46 (3) (2013) J. Koros, Water sorption in poly(ethylene furanoate) compared to poly(ethylene
796–804, https://doi.org/10.1021/ma3023298. terephthalate). Part 2: Kinetic sorption, Polymer 55 (26) (2014) 6870–6882,
[79] J.A. Moore, J.E. Kelly, Polyesters derived from furan and tetrahydrofuran nuclei, https://doi.org/10.1016/j.polymer.2014.10.065.
Macromolecules 11 (3) (1978) 568–573, https://doi.org/10.1021/ [101] L. Sun, J. Wang, S. Mahmud, Y. Jiang, J. Zhu, X. Liu, New insight into the
ma60063a028. mechanism for the excellent gas properties of poly(ethylene 2,5-furandicarboxy­
[80] S. Kato, A. Kasai, Method for producing polyester resin including furan structure, late) (PEF): Role of furan ring’s polarity, Eur. Polym. J. 118 (2019) 642–650,
Japanese patent application JP2008/291244, filed April, 24, 2007, 2007. https://doi.org/10.1016/j.eurpolymj.2019.06.033.
[81] L. Wu, R. Mincheva, Y. Xu, J.M. Raquez, P. Dubois, High molecular weight poly [102] J. Wang, X. Liu, Y. Zhang, F. Liu, J. Zhu, Modification of poly(ethylene 2,5-fur­
(butylene succinate-co-butylene furandicarboxylate) copolyesters: from catalyzed andicarboxylate) with 1,4-cyclohexanedimethylene: Influence of composition on
polycondensation reaction to thermomechanical properties, Biomacromolecules mechanical and barrier properties, Polymer 103 (2016) 1–8, https://doi.org/
13 (9) (2012) 2973–2981, https://doi.org/10.1021/bm301044f. 10.1016/j.polymer.2016.09.030.
[82] N. Kasmi, G. Papageorgiou, D. Achilias, D. Bikiaris, Solid-state polymerization of [103] J. Wang, X. Liu, Z. Jia, L. Sun, Y. Zhang, J. Zhu, Modification of poly(ethylene 2,5-
poly(ethylene furanoate) biobased polyester, II: An efficient and facile method to furandicarboxylate) (PEF) with 1, 4-cyclohexanedimethanol: Influence of
synthesize high molecular weight polyester appropriate for food packaging stereochemistry of 1,4-cyclohexylene units, Polymer 137 (2018) 173–185,
applications, Polymers 10 (5) (2018) 471, https://doi.org/10.3390/ https://doi.org/10.1016/j.polymer.2018.01.021.
polym10050471. [104] G. Guidotti, M. Soccio, N. Lotti, M. Gazzano, V. Siracusa, A. Munari, Poly
[83] Y. Chebbi, N. Kasmi, M. Majdoub, G.Z. Papageorgiou, D.S. Achilias, D.N. Bikiaris, (propylene 2,5-thiophenedicarboxylate) vs. poly(propylene 2,5-furandicarboxy­
Solid-State Polymerization of Poly(Ethylene Furanoate) Biobased Polyester, III: late): Two examples of high gas barrier bio-based polyesters, Polymers (Basel) 10
Extended Study on Effect of Catalyst Type on Molecular Weight Increase, (7) (2018) 785, https://doi.org/10.3390/polym10070785.
Polymers (Basel) 11 (3) (2019) 438, https://doi.org/10.3390/polym11030438. [105] M. Vannini, P. Marchese, A. Celli, C. Lorenzetti, Fully biobased poly(propylene
[84] J.C. Morales-Huerta, C.B. Ciulik, A.M. de Ilarduya, S. Munoz-Guerra, Fully bio- 2,5-furandicarboxylate) for packaging applications: excellent barrier properties as
based aromatic-aliphatic copolyesters: poly(butylene furandicarboxylate-co- a function of crystallinity, Green Chem. 17 (8) (2015) 4162–4166, https://doi.
succinate)s obtained by ring opening polymerization, Polym. Chem. 8 (4) (2017) org/10.1039/c5gc00991j.
748–760, https://doi.org/10.1039/c6py01879c. [106] L. Genovese, N. Lotti, V. Siracusa, A. Munari, Poly(neopentyl glycol furanoate): A
[85] N. Kasmi, M. Wahbi, L. Papadopoulos, Z. Terzopoulou, N. Guigo, N. Sbirrazzuoli, member of the furan-based polyester family with smart barrier performances for
G.Z. Papageogiou, D.N. Bikiaris, Synthesis and characterization of two new sustainable food packaging applications, Materials (Basel) 10 (9) (2017) 1028,
biobased poly(pentylene 2,5-furandicarboxylate-co-caprolactone) and poly https://doi.org/10.3390/ma10091028.
(hexamethylene 2,5-furandicarboxylate-co-caprolactone) copolyesters with [107] X. Fei, J. Wang, J. Zhu, X. Wang, X. Liu, Bio-based poly (ethylene 2, 5-furanoate):
enhanced enzymatic hydrolysis properties, Polym. Degrad. Stab. 160 (2019) no longer an alternative, but an irreplaceable one in polymer industry, ACS
242–263, https://doi.org/10.1016/j.polymdegradstab.2019.01.004. Sustain. Chem. Eng. 8 (23) (2020) 8471–8485, https://doi.org/10.1021/
[86] Y. Long, R.Y. Zhang, J.C. Huang, J.G. Wang, J.W. Zhang, N. Rayand, G. Hu, acssuschemeng.0c01862.
J. Yang, J. Zhu, Retroreflection in binary bio-based PLA/PBF blends, Polymer 125 [108] Z. Terzopoulou, E. Karakatsianopoulou, N. Kasmi, V. Tsanaktsis, N. Nikolaidis,
(2017) 138–143, https://doi.org/10.1016/j.polymer.2017.08.004. M. Kostoglou, G.Z. Papageorgiou, D.A. Lambropoulou, D.N. Bikiaris, Effect of
[87] Juan Carlos Morales-Huerta, Antxon Martínez de Ilarduya, Salvador León, catalyst type on molecular weight increase and coloration of poly(ethylene
Sebastián Muñoz-Guerra, Isomannide-containing poly(butylene 2,5-furandicar­ furanoate) biobased polyester during melt polycondensation, Polym. Chem. 8
boxylate) copolyesters via ring opening polymerization, Macromolecules 51 (9) (44) (2017) 6895–6908, https://doi.org/10.1039/c7py01171g.
(2018) 3340–3350, https://doi.org/10.1021/acs.macromol.8b00487. [109] T. Ghosh, K. Mahajan, S. Narayan-Sarathy, M.N. Balgacem, P. Gopalakrishnan,
[88] Juan Carlos Morales-Huerta, Antxon Martínez de Ilarduya, Sebastián Muñoz- dicarboxylic acid-based polyesters prepared from biomass, Google Patents. 2
Guerra, Blocky poly (ε-caprolactone-co-butylene 2, 5-furandicarboxylate) (2013) 5-Furan.
copolyesters via enzymatic ring opening polymerization, J. Polym. Sci., Part A: [110] J.G. Wang, X.Q. Liu, J. Zhu, From furan to high quality bio-based poly(ethylene
Polym. Chem. 56 (3) (2018) 290–299, https://doi.org/10.1002/pola. furandicarboxylate), Chin. J. Polym. Sci. 36 (6) (2018) 720–727, https://doi.org/
v56.310.1002/pola.28895. 10.1007/s10118-018-2092-0.
[89] A. Todea, I. Bitcan, D. Aparaschivei, I. Pausescu, V. Badea, F. Peter, V. [111] E. de Jong, M.A. Dam, L. Sipos, G.J.M. Gruter, Furandicarboxylic acid (FDCA), A
D. Gherman, G. Rusu, L. Nagy, S. Keki, Biodegradable Oligoesters of versatile building block for a very interesting class of polyesters, in: Biobased
ε-Caprolactone and 5-Hydroxymethyl-2-Furancarboxylic Acid Synthesized by Monomers, Polymers, and Materials, 2012, pp. 1–13, https://doi.org/10.1002/
Immobilized Lipases, Polymers (Basel) 11 (9) (2019) 1402, https://doi.org/ app.43391.
10.3390/polym11091402. [112] E. Gubbels, L. Jasinska-Walc, B.A.J. Noordover, C.E. Koning, Linear and branched
[90] D. Maniar, Y. Jiang, A.J.J. Woortman, J. van Dijken, K. Loos, Furan-based polyester resins based on dimethyl-2, 5-furandicarboxylate for coating
copolyesters from renewable resources: Enzymatic synthesis and properties, applications, Eur. Polym. J. 49 (10) (2013) 3188–3198, https://doi.org/10.1016/
ChemSusChem 12 (5) (2019) 990–999, https://doi.org/10.1002/ j.eurpolymj.2013.06.019.
cssc.201802867. [113] M. Gruter, G.-J.L. Sipos, M. Adrianus Dam, Accelerating research into bio-based
[91] M. Gomes, A. Gandini, A.J.D. Silvestre, B. Reis, Synthesis and characterization of FDCA-polyesters by using small scale parallel film reactors, Comb. Chem. High
poly(2,5-furan dicarboxylate)s based on a variety of diols, J. Polym. Sci. Part A Throughput Screening 15 (2) (2012) 180–188, https://doi.org/10.2174/
Polym. Chem. 49 (17) (2011) 3759–3768, https://doi.org/10.1002/pola.24812. 138620712798868374.
[92] J. Wang, X. Liu, J. Zhu, Y. Jiang, Copolyesters based on 2,5-furandicarboxylic [114] Nicolas Jacquel, Floriane Freyermouth, Françoise Fenouillot, Alain Rousseau,
acid (FDCA): Effect of 2,2,4,4-tetramethyl-1,3-cyclobutanediol units on their Jean Pierre Pascault, Patrick Fuertes, René Saint-Loup, Synthesis and properties
properties, Polymers 9 (9) (2017) 305, https://doi.org/10.3390/polym9090305. of poly(butylene succinate): Efficiency of different transesterification catalysts,
[93] Jinggang Wang, Sakil Mahmud, Xiaoqin Zhang, Jin Zhu, Zhisen Shen, J. Polym. Sci. Part A Polym. Chem. 49 (24) (2011) 5301–5312, https://doi.org/
Xiaoqing Liu, Biobased amorphous polyesters with high Tg: Trade-off between 10.1002/pola.v49.2410.1002/pola.25009.
rigid and flexible cyclic diols, ACS Sustain. Chem. Eng. 7 (6) (2019) 6401–6411, [115] J.G. Rosenboom, D.K. Hohl, P. Fleckenstein, G. Storti, M. Morbidelli, Bottle-grade
https://doi.org/10.1021/acssuschemeng.9b00285. polyethylene furanoate from ring-opening polymerisation of cyclic oligomers,
[94] J. Wang, X. Liu, Z. Jia, L. Sun, J. Zhu, Highly crystalline polyesters synthesized Nat. Commun. 9 (1) (2018) 2701, https://doi.org/10.1038/s41467-018-05147-y.
from furandicarboxylic acid (FDCA): Potential bio-based engineering plastic, Eur. [116] J.P. Ma, Y. Pang, M. Wang, J. Xu, H. Ma, X. Nie, The copolymerization reactivity
Polym. J. 109 (2018) 379–390, https://doi.org/10.1016/j. of diols with 2,5-furandicarboxylic acid for furan-based copolyester materials,
eurpolymj.2018.10.014. J. Mater. Chem. 22 (8) (2012) 3457–3461, https://doi.org/10.1039/c2jm15457a.
[95] Jinggang Wang, Liyuan Sun, Zhisen Shen, Jin Zhu, Xingliang Song, Xiaoqing Liu, [117] Hongzhou Xie, Linbo Wu, Bo-Geng Li, Philippe Dubois, Modification of poly
Effects of various 1,3-propanediols on the properties of poly(propylene (ethylene 2,5-furandicarboxylate) with biobased 1,5-pentanediol: significantly
furandicarboxylate), ACS Sustainable Chem. Eng. 7 (3) (2019) 3282–3291, toughened copolyesters retaining high tensile strength and O2 barrier property,
https://doi.org/10.1021/acssuschemeng.8b05288. Biomacromolecules 20 (1) (2019) 353–364, https://doi.org/10.1021/acs.
biomac.8b01495.

20
S. Pandey et al. European Polymer Journal 160 (2021) 110778

[118] Jedediah F. Wilson, Eugene Y.-X. Chen, Difuranic diols for renewable polymers Eng. 5 (10) (2017) 9244–9253, https://doi.org/10.1021/
with pendent furan rings, ACS Sustainable Chem. Eng. 7 (7) (2019) 7035–7046, acssuschemeng.7b02196.
https://doi.org/10.1021/acssuschemeng.8b06819. [140] M. Matos, A.F. Sousa, A.C. Fonseca, C.S.R. Freire, J.F.J. Coelho, A.J.D. Silvestre,
[119] N. Peelman, P. Ragaert, B. De Meulenaer, D. Adons, R. Peeters, L. Cardon, F. A new generation of furanic copolyesters with enhanced degradability: Poly
V. Impe, F. Devlieghere, Application of bioplastics for food packaging, Trends (ethylene 2,5-furandicarboxylate)-co-poly(lactic acid) copolyesters, Macromol.
Food Sci. Technol. 32 (2) (2013) 128–141, https://doi.org/10.1016/j. Chem. Phys. 215 (22) (2014) 2175–2184, https://doi.org/10.1002/
tifs.2013.06.003. macp.201400175.
[120] K. Petersen, P.V. Nielsen, G. Bertelsen, M. Lawther, M.B. Olsen, N.H. Nilsson, [141] Han Hu, Ruoyu Zhang, Lei Shi, Wu Bin Ying, Jinggang Wang, Jin Zhu,
G. Mortensen, Potential of biobased materials for food packaging, Trends Food Modification of poly(butylene 2,5-furandicarboxylate) with lactic acid for
Sci. Technol. 10 (2) (1999) 52–68, https://doi.org/10.1016/S0924-2244(99) biodegradable copolyesters with good mechanical and barrier Properties, Ind.
00019-9. Eng. Chem. Res. 57 (32) (2018) 11020–11030, https://doi.org/10.1021/acs.
[121] T. Iwata, Biodegradable and bio-based polymers: Future prospects of eco-friendly iecr.8b02169.
plastics, Angew. Chem. Int. Ed. 54 (11) (2015) 3210–3215, https://doi.org/ [142] S. Qiu, Z. Su, Z. Qiu, Isothermal and nonisothermal crystallization kinetics of
10.1002/anie.201410770. novel biobased poly (ethylene succinate-co-ethylene sebacate) copolymers from
[122] Mitchell F. Shockley, Anastasia H. Muliana, Modeling temporal and spatial the amorphous state, J. Therm. Anal. Calorim. 129 (2) (2017) 801–808, https://
changes during hydrolytic degradation and erosion in biodegradable polymers, doi.org/10.1007/s10973-017-6234-3.
Polym. Degrad. Stab. 180 (2020) 109298, https://doi.org/10.1016/j. [143] G.Q. Wang, M. Jiang, Q. Zhang, R. Wang, G.Y. Zhou, Biobased copolyesters:
polymdegradstab.2020.109298. synthesis, crystallization behavior, thermal and mechanical properties of poly
[123] P. Falkenstein, D. Gräsing, P. Bielytskyi, W. Zimmermann, J. Matysik, R. Wei, (ethylene glycol sebacate-co-ethylene glycol 2,5-furan dicarboxylate), RSC Adv. 7
C. Song, UV pretreatment impairs the enzymatic degradation of polyethylene (23) (2017) 13798–13807, https://doi.org/10.1039/c6ra27795k.
terephthalate, Front. Microbiol. 11 (2020) 689, https://doi.org/10.3389/ [144] G.Q. Wang, M. Jiang, Q. Zhang, R. Wang, X.B. Tong, S.Y. Xue, G.Y. Zhou,
fmicb.2020.00689. Biobased copolyesters: Synthesis, sequence distribution, crystal structure, thermal
[124] T.A. Cooper, Developments in bioplastic materials for packaging food, beverages and mechanical properties of poly(butylene sebacate-co-butylene
and other fast-moving consumer goods, in: N. Farmer (Ed.), Trends in Packaging furandicarboxylate), Polym. Degrad. Stab. 143 (2017) 1–8, https://doi.org/
of Food, Beverages and Other Fast-Moving Consumer Goods (FMCG), Woodhead 10.10165/j.polymdegradsiab.2017.05.027.
Publishing, 2013, pp. 108–152. [145] M. Soccio, M. Costa, N. Lotti, M. Gazzano, V. Siracusa, E. Salatelli, P. Manaresi,
[125] V. Siracusa, P. Rocculi, S. Romani, M. Dalla Rosa, Biodegradable polymers for A. Munari, Novel fully biobased poly(butylene 2,5-furanoate/diglycolate)
food packaging: a review, Trends Food Sci. Technol. 19 (12) (2008) 634–643, copolymers containing ether linkages: Structure-property relationships, Eur.
https://doi.org/10.1016/j.tifs.2008.07.003. Polym. J. 81 (2016) 397–412, https://doi.org/10.1016/j.eurpolymj.2016.06.022.
[126] Ashutos Parhi, Juming Tang, Shyam S. Sablani, Functionality of ultra-high barrier [146] H. Hu, R.Y. Zhang, J.G. Wang, W.B. Ying, L. Shi, C. Yao, Z. Kong, K. Wang, J. Zhu,
metal oxide-coated polymer films for in-package, thermally sterilized food A mild method to prepare high molecular weight poly(butylene
products, Food Packag. Shelf Life 25 (2020) 100514, https://doi.org/10.1016/j. furandicarboxylate-co-glycolate) copolyesters: effects of the glycolate content on
fpsl.2020.100514. thermal, mechanical, and barrier properties and biodegradability, Green Chem.
[127] R.J. Müller, U. Witt, E. Rantze, W.D. Deckwer, Architecture of biodegradable 21 (11) (2019) 3013–3022, https://doi.org/10.1039/c9gc00668k.
copolyesters containing aromatic constituents, Polym. Degrad. Stab. 59 (1–3) [147] H. Hu, R. Zhang, Z. Kong, K. Wang, W.B. Ying, J. Wang, J. Zhu, Bio-based poly
(1998) 203–208, https://doi.org/10.1016/s0141-3910(97)00186-9. (butylene furandicarboxylate)-b-poly(ethylene glycol) copolymers: The effect of
[128] U. Witt, R.J. Müller, W.D. Deckwer, Studies on sequence distribution of aliphatic/ poly(ethylene glycol) molecular weight on thermal properties and hydrolysis
aromatic copolyesters by high-resolution 13C nuclear magnetic resonance degradation behavior, Adv. Ind. Eng. Polym. Res. 2 (4) (2019) 167–177, https://
spectroscopy for evaluation of biodegradability, Macromol. Chem. Phys. 197 (4) doi.org/10.1016/j.aiepr.2019.09.002.
(1996) 1525–1535, https://doi.org/10.1002/macp.1996.021970428. [148] G. Wang, M. Jiang, Q. Zhang, R. Wang, G. Zhou, Biobased multiblock copolymers:
[129] A.A. Shah, S. Kato, N. Shintani, N.R. Kamini, T. Nakajima-Kambe, Microbial Synthesis, properties and shape memory performance of poly(ethylene 2,5-fur­
degradation of aliphatic and aliphatic-aromatic co-polyesters, Appl. Microbiol. andicarboxylate)-b-poly(ethylene glycol), Polym. Degrad. Stab. 144 (2017)
Biotechnol. 98 (8) (2014) 3437–3447, https://doi.org/10.1007/s00253-014- 121–127, https://doi.org/10.1016/j.polymdegradstab.2017.07.032.
5558-1. [149] P. Ji, D. Lu, S. Zhang, W. Zhang, C. Wang, H. Wang, Modification of poly(ethylene
[130] E. Cranston, J. Kawada, S. Raymond, F.G. Morin, R.H. Marchessault, 2,5-furandicarboxylate) with poly(ethylene glycol) for biodegradable
Cocrystallization model for synthetic biodegradable poly(butylene adipate-co- copolyesters with good mechanical properties and spinnability, Polymers (Basel)
butylene terephthalate), Biomacromolecules 4 (4) (2003) 995–999, https://doi. 11 (12) (2019) 2105, https://doi.org/10.3390/polym11122105.
org/10.1021/bm034089n. [150] H. Hu, R. Zhang, A. Sousa, Y. Long, W.B. Ying, J. Wang, J. Zhu, Bio-based poly
[131] S.B. Peng, B.S. Wu, L.B. Wu, B.G. Li, P. Dubois, Hydrolytic degradation of (butylene 2,5-furandicarboxylate)-b-poly(ethylene glycol) copolymers with
biobased poly(butylene succinate-co-furandicarboxylate) and poly(butylene adjustable degradation rate and mechanical properties: Synthesis and
adipate-co-furandicarboxylate) copolyesters under mild conditions, J. Appl. characterization, Eur. Polym. J. 106 (2018) 42–52, https://doi.org/10.1016/j.
Polym. Sci. 134 (15) (2017). DOI: ARTN 4467410.1002/app.44674. eurpolymj.2018.07.007.
[132] B.S. Wu, Y.T. Xu, Z.Y. Bu, L.B. Wu, B.G. Li, P. Dubois, Biobased poly(butylene 2,5- [151] M.A. Woodruff, D.W. Hutmacher, The return of a forgotten polymer-
furandicarboxylate) and poly(butylene adipate-co-butylene 2,5-furandicarboxy­ Polycaprolactone in the 21st century, Prog. Polym. Sci. 35 (10) (2010)
late)s: From synthesis using highly purified 2,5-furandicarboxylic acid to thermo- 1217–1256, https://doi.org/10.1016/j.progpolymsci.2010.04.002.
mechanical properties, Polymer 55 (16) (2014) 3648–3655, https://doi.org/ [152] Xiansong Wang, Qingyin Wang, Shaoying Liu, Teng Sun, Gongying Wang,
10.1016/j.polymer.2014.06.052. Synthesis and properties of poly(isosorbide 2,5-furandicarboxylate-co-
[133] Z. Jia, J. Wang, L. Sun, J. Zhu, X. Liu, Fully bio-based polyesters derived from 2,5- ε-caprolactone) copolyesters, Polym. Test. 81 (2020) 106284, https://doi.org/
furandicarboxylic acid (2,5-FDCA) and dodecanedioic acid (DDCA): From 10.1016/j.polymertesting.2019.106284.
semicrystalline thermoplastic to amorphous elastomer, J. Appl. Polym. Sci. 135 [153] Liliana Gustini, Cristina Lavilla, Antxon Martínez de Ilarduya, Sebastián Muñoz-
(14) (2018) 46076, https://doi.org/10.1002/app.46076. Guerra, Cor E. Koning, Isohexide and sorbitol-derived, enzymatically synthesized
[134] Z.L. Yu, J.D. Zhou, F. Cao, B.B. Wen, X. Zhu, P. Wei, Chemosynthesis and renewable polyesters with enhanced Tg, Biomacromolecules 17 (10) (2016)
characterization of fully biomass-based copolymers of ethylene glycol, 2,5-fur­ 3404–3416, https://doi.org/10.1021/acs.biomac.6b01224.
andicarboxylic acid, and succinic acid, J. Appl. Polym. Sci. 130 (2) (2013) [154] V. Tsanaktsis, G.Z. Papageorgiou, D.N. Bikiaris, A facile method to synthesize
1415–1420, https://doi.org/10.1002/app.39344. high-molecular-weight biobased polyesters from 2,5-furandicarboxylic acid and
[135] N. Jacquel, R. Saint-Loup, J.P. Pascault, A. Rousseau, F. Fenouillot, Bio-based long-chain diols, J. Polym. Sci. Part A Polym. Chem. 53 (22) (2015) 2617–2632,
alternatives in the synthesis of aliphatic-aromatic polyesters dedicated to https://doi.org/10.1002/pola.27730.
biodegradable film applications, Polymer 59 (2015) 234–242, https://doi.org/ [155] Q. Ouyang, J.J. Liu, C.C. Li, L.C. Zheng, Y.N. Xiao, S.H. Wu, B. Zhang, A facile
10.1016/j.polymer.2014.12.021. method to synthesize bio-based and biodegradable copolymers from
[136] Han Hu, Ruoyu Zhang, Yanhua Jiang, Lei Shi, Jinggang Wang, Wu Bin Ying, furandicarboxylic acid and isosorbide with high molecular weights and excellent
Jin Zhu, Toward biobased, biodegradable, and smart barrier packaging material: thermal and mechanical properties, Polym. Chem. 10 (41) (2019) 5594–5601,
modification of poly(neopentyl glycol 2,5-furandicarboxylate) with succinic acid, https://doi.org/10.1039/c9py01314h.
ACS Sustainable Chem. Eng. 7 (4) (2019) 4255–4265, https://doi.org/10.1021/ [156] J. Zhang, J. Li, Y. Tang, L. Lin, M. Long, Advances in catalytic production of bio-
acssuschemeng.8b05990. based polyester monomer 2,5-furandicarboxylic acid derived from lignocellulosic
[137] Z. Terzopoulou, V. Tsanaktsis, D.N. Bikiaris, S. Exarhopoulos, D.G. Papageorgiou, biomass, Carbohydr. Polym. 130 (2015) 420–428, https://doi.org/10.1016/j.
G.Z. Papageorgiou, Biobased poly(ethylene furanoate-co-ethylene succinate) carbpol.2015.05.028.
copolyesters: solid state structure, melting point depression and biodegradability, [157] J.-W. Rhim, S.-I. Hong, C.-S. Ha, Tensile, water vapor barrier and antimicrobial
RSC Adv. 6 (87) (2016) 84003–84015, https://doi.org/10.1039/c6ra15994j. properties of PLA/nanoclay composite films, LWT-Food Sci. Technol. 42 (2)
[138] E. Castro-Aguirre, F. Iñiguez-Franco, H. Samsudin, X. Fang, R. Auras, Poly(lactic (2009) 612–617, https://doi.org/10.1016/j.lwt.2008.02.015.
acid)—Mass production, processing, industrial applications, and end of life, Adv. [158] L.B. Tavares, N.M. Ito, M.C. Salvadori, D.J. dos Santos, D.S. Rosa, PBAT/kraft
Drug Deliv. Rev. 107 (2016) 333–366, https://doi.org/10.1016/j. lignin blend in flexible laminated food packaging: Peeling resistance and thermal
addr.2016.03.010. degradability, Polym. Test. 67 (2018) 169–176, https://doi.org/10.1016/j.
[139] Y. Long, R.Y. Zhang, J.C. Huang, J.G. Wang, Y.H. Jiang, G.H. Hu, J. Yang, J. Zhu, polymertesting.2018.03.004.
Tensile Property Balanced and Gas Barrier Improved Poly(lactic acid) by Blending
with Biobased Poly(butylene 2,5-furan dicarboxylate), ACS Sustainable Chem.

21
S. Pandey et al. European Polymer Journal 160 (2021) 110778

[159] S.J. Modi, Assessing the feasibility of poly-(3-hydroxybutyrate-co-3- [161] N. Poulopoulou, G. Kantoutsis, D.N. Bikiaris, D.S. Achilias, M. Kapnisti, G.
hydroxyvalerate)(PHBV) and poly-(lactic acid) for potential food packaging Z. Papageorgiou, Biobased Engineering Thermoplastics: Poly(butylene 2,5-fur­
applications, The Ohio State University, 2010. andicarboxylate) Blends, Polymers (Basel) 11 (6) (2019) 937, https://doi.org/
[160] V.P. Cyras, M.S. Commisso, A.N. Mauri, A. Vázquez, Biodegradable double-layer 10.3390/polym11060937.
films based on biological resources: Polyhydroxybutyrate and cellulose, J. Appl.
Polym. Sci. 106 (2) (2007) 749–756, https://doi.org/10.1002/app.26663.

22

You might also like