Вы находитесь на странице: 1из 23

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Polymer
Chemistry
View Article Online

REVIEW

View Journal

The quest for sustainable polyesters insights into


the future
Cite this: DOI: 10.1039/c3py01213a

nio C. Serra,b
Carla Vilela,*a Andreia F. Sousa,ab Ana C. Fonseca,b Arme
b
a
Jorge F. J. Coelho, Carmen S. R. Freire and Armando J. D. Silvestrea
Polyesters from renewable resources are an expanding area with a burgeoning scientic activity,
nevertheless little has been reviewed about this particular class of polymers. The present appraisal
intends to contribute to ll this literature gap by reviewing recent aspects related to the most promising
Received 2nd September 2013
Accepted 5th October 2013

renewable-based polyesters. Emphasis will be placed on bio-based polyesters that, given their
comparable properties, may replace polymers derived from fossil fuel feedstock, and on bio-based
polyesters with completely innovative properties for novel applications. Furthermore, the sources of

DOI: 10.1039/c3py01213a

renewable monomers will also be reviewed, together with the most relevant eco-friendly synthetic

www.rsc.org/polymers

approaches used in polycondensation reactions leading to polyesters.

1.

Introduction

The current technological, social and economic situation,


which involves issues related to the environment, waste
disposal, depletion of non-renewable resources and also the
a

CICECO and Chemistry Department, University of Aveiro, Campus de Santiago,


3810-193 Aveiro, Portugal. E-mail: cvilela@ua.pt; Fax: +351 234370084; Tel: +351
234370711
Department of Chemical Engineering, University of Coimbra, Rua Slvio Lima-Polo II,
3030-290, Coimbra, Portugal

This paper is part of a Polymer Chemistry themed issue on Sustainable


Polymers: replacing polymers derived from fossil fuels. Guest Editor: Stephen
A. Miller.

Carla Vilela was born in 1981


(Aveiro, Portugal), she graduated in Chemical Engineering in
2004 and obtained a Master
degree in Materials Science and
Engineering (EMMS) in 2008 at
the University of Aveiro (UA,
Portugal). In 2012, she obtained
her PhD at the same university,
working on monomeric and
polymeric materials from vegetable oils. Currently, she is a
postdoctoral researcher at the
Associate Laboratory CICECO (UA, Portugal) and her research
interests are focused on the sustainable use of renewable resources
(vegetable oils, cork, cellulose, proteins) for the development of
novel monomers, polymers and functional (nano)composite
materials.

This journal is The Royal Society of Chemistry 2013

unpredictable crude oil price uctuations or its ineluctable


depletion, stimulates investigation aimed at developing
macromolecular materials from renewable resources as alternatives to the current fossil-based polymers.14 The use of
renewable resources, generally understood as any animal or
vegetable species which is exploited without endangering its
survival and which is renewed by biological (short term) instead of
geochemical (very long term) activities,1 can considerably
contribute to the sustainable development of humankind.
Therefore, the search for the concept of acting responsibly to
meet the needs of the present without compromising the ability of
future generations to meet their own needs5 based on nonpetrochemical feedstock has become one of the social

Andreia F. Sousa studied Industrial Chemistry at the University


of Coimbra where she obtained
both her BSc and MSc degrees.
Aerwards, she received her PhD
in Chemistry at the University of
Aveiro in 2011. Currently, she is
a post-doc fellow at the University of Aveiro. She has been
acutely involved in research in
polymer chemistry both in the
elds of physical chemistry and
organic chemistry. Her main
research interests are currently focused on the synthesis and
characterisation of polyesters from renewable resources, namely
furan- and suberin-based polyesters.

Polym. Chem.

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Polymer Chemistry

Review

paradigms of the 21st century, since renewable raw materials are


potentially capable of providing a wide variety of monomers and
polymers as comprehensive as those presently produced by the
petrochemical industry.6,7
Although this renewed interest in polymers from renewable
resources has gained wings recently, this whole idea is not new.
Indeed, the advent of bio-based polymeric materials is entangled with the birth of macromolecular science and technology
in the 19th century with the synthesis of some very well-known
materials such as celluloid,8 vulcanised natural rubber1,9 or even
nylon-6,6.1,9 However, the relative importance of macromolecular materials based on renewable resources suered thereaer
a gradual setback essentially due to the petrochemical revolution of the last century. In fact, polymers prepared from fossil
resources, with huge quantities of polyethylene (HDPE, LDPE),
polypropylene (PP), poly(vinyl chloride) (PVC), poly(ethylene
terephthalate) (PET), and polystyrene (PS), together with smaller

quantities of other more sophisticated materials as poly(tetrauoroethylene) (PTFE), aramides, just to cite some of the more
common, are present in all domains of everyday life.10 In the
beginning of the twenty rst century, the scenario just
described is changing as the need for petrochemical plastic
alternatives becomes more urgent, and science drives attention
once more to an old alternative polymers based on renewable
resources. Actually, the production of bio-plastics (partially or
totally bio-based, biodegradable or both) is already a growing
niche market that has increased from less than 300 000 metric
tons in 2009 to surpass 1.0 million metric tons in 2011 and with
a forecast of 5.8 million metric tons for 2016.11
Polyesters are denitely among the most promising families
of polymers based on renewable resources, thanks to their
unique and vast array of properties, e.g. bre forming ability,
potential biodegradability and, in some cases, biocompatibility,
among many others,12,13 but also because the bio-based

Ana C. Fonseca was born in


Viseu, in 1984. In 2007, she
graduated in Chemical Engineering, from the Faculty of
Sciences and Technology of the
University of Coimbra, Portugal.
In 2013, she obtained her PhD
from the same University
working on the development of
poly(ester amide)s based on aamino acids and a-hydroxy
acids. Currently, she is a postdoctorate researcher in the
Chemical Engineering Department of the University of Coimbra
and her research interests are focused on the development of
biodegradable polymers from renewable resources. She co-authored 14 papers and 3 book chapters.

Carmen S. R. Freire was born in


1976 (Caracas, Venezuela), she
graduated in Industrial Chemistry in 1998 and obtained her
PhD in Chemistry in 2003 at the
University of Aveiro (UA, Portugal). She is currently Principal
Researcher at the Associate
Laboratory CICECO (University
of Aveiro-UA). Her research
interests are centered on the
chemistry of polymers from
renewable resources (characterization and application), development of new sustainable
composite materials, nanostructured biomaterials and generally
on the chemistry of natural compounds. She is author/co-author of
98 articles in SCI journals, two national and two international
patents.

Jorge F. J. Coelho was born in


Figueira da Foz, Portugal, in
1978. He graduated from the
Faculty of Science and Technology of the University of
Coimbra, Portugal, in Chemical
Engineering. In 2006, he
obtained his PhD from the
University of Coimbra working
on new living radical polymerization methods for vinyl chloride. In 2009, he was appointed
as assistant professor at the
University of Coimbra. He has coauthored 60 peer-reviewed
research papers and 7 book chapters. His research interests include
living radical polymerization, bio-based and biodegradable polymers, novel pharmaceutical products, supramolecular chemistry
and scale-up approaches for living radical polymerization.

Armando J. D. Silvestre was born


in 1968 (Aveiro, Portugal), he
graduated in Chemistry in 1990,
got his PhD in 1994 and the
Habilitation degree in 2008 at
the University of Aveiro (UA,
Portugal). He is currently Associate Professor at the Chemistry
Department of UA, Researcher of
the same department and of the
Associate Laboratory CICECO
(UA). He is author/co-author of
170 papers in international
journals, 8 book chapters, two national and two international
patents. His research interests comprise the extraction, characterization and valorisation of biomass components in areas such
as low molecular weight bioactive components, biobased polyesters, and (nano)cellulose bre-based nanocomposite materials.

Polym. Chem.

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Review
monomers for polyesters are relatively easy accessible
(compared with e.g. isocyanates, diamines, etc.). Polyesters
include numerous materials with dierent monomer structures
and properties, and nd a wide range of applications including
bottles, bres, photographic lms, recording tapes, among
others, thus reecting the high versatility and importance of
such a class of materials.14 There is already a wide number of
studies concerning the use of renewable resources in polymer
synthesis, and the arsenal of polyesters partially or totally
obtained from natural origin that could mimic plastics of
petrochemical origin is steadily growing.15
Notwithstanding the increasing scientic activity, there is
still one major problem usually associated with them, or with
renewable-based polymers in that matter: their relatively high
cost when compared with their petrochemical homologues.8,16
This is essentially due to the current high cost of bio-based
chemicals used as building blocks. However, the eective
implementation of the biorenery concept17 is pointed out by
many as the solution to bring these chemicals available at
competitive prices.6,18,19 In general terms, this concept describes
a platform in which multiple biomass feedstocks are transformed into a portfolio of products, including fuels, chemicals
and materials in addition to power,6,18 being the base of the socalled bioeconomy. Among these bioreneries of the future,20
the lignocellulosic feedstock (LCF) bioreneries19 will probably
be the most successful, on the one hand, because of the availability of feedstock (e.g. agro-food wastes, or even paper wastes)
at competitive prices and, on the other hand, because they do
not compete with the food supply. Furthermore, it has been
estimated that of the 170 trillion tonnes of biomass produced by
nature, only 3.5% are utilised by mankind and only 5% of those
are consumed for non-food purposes.15 Therefore, bioreneries
will be able to supply all chemical and material needs based on
lignocellulosic feedstocks without competing with the food
supply chain for cultivation areas and without endangering
biodiversity and protected ecosystems. So, polysaccharides (e.g.
cellulose and hemicelluloses), lignin, non edible vegetable oils,
and also suberin will certainly be the key feedstocks for the
production of monomeric building blocks and of the ensuing
sustainable polyesters, as will be seen below.
Numerous biomass-based chemicals are potentially available today in large amounts and ready to be used as building
blocks for polyester production (some are already in use),
namely sugars and their derivatives, vegetable oils, organic
acids, glycerol, suberin, cutin, among many others.1,10,21 Despite
the cost issues, there are already a few successful commercial
examples of biomass derived polyesters such as poly(lactic acid)
(PLA), poly(butylene succinate) (PBS), poly(butylene succinate
adipate) (PBSA), and poly(ethylene-2,5-furandicarboxylate)
(PEF). NatureWorks LLC is the main supplier of PLA under the
brand name Ingeo,22 and 100% natural PLA tableware is
already being commercialized by CornFlower.23 Showa Denko is
one of the global producers of PBS and PBSA under the trade
name Bionolle,24 whereas Avantium is manufacturing PEF for
packaging of so drinks, water, alcoholic beverages, among
others, based on the Avantium's YXY technology.25 Furthermore, there are other examples of commercial polyesters that

This journal is The Royal Society of Chemistry 2013

Polymer Chemistry
are only partially renewable, namely Sorona EP (poly(trimethylene terephthalate), PTT) and Hytrel RS (block copolyester of poly(butylene terephthalate) and poly(tetramethylene
glycol), PBT-b-PTMG) produced by DuPont,26 and the rstgeneration PET PlantBottle.27 Many more examples will come to
light in the near future as chemicals from forest and agriculture
feedstock become cheaper and available on a large scale, and as
new and more advanced synthetic approaches to polymerization of these renewable based chemicals emerge. This will
hopefully guarantee a shi of paradigm with the industrialisation and ensuing widespread commercialisation of these polyesters from renewable resources.
Numerous reviews on polyesters2831 have been published
recently, including e.g. the appraisals on cyclic,29 star,30 and
dendritic31 polyesters. Nevertheless, and to the best of our
knowledge, there have been no comprehensive reviews devoted
to polyesters from renewable resources. In this vein, the
purpose of the present appraisal is to provide a succinct
assessment of the recent advances in the domain of polyesters
based on plant-derived renewable resources. Although renewable molecules and derivatives present a panoply of possibilities
for polyester science, the examples in this appraisal will focus
on bio-based polyesters with aliphatic and/or aromatic domains
that, given their comparable properties, may replace polymers
derived from fossil fuel feedstock, and on bio-based polyesters
with completely innovative properties for novel applications.
Furthermore, polymers from natural monomers, such as
PLA, and polymers from microbial fermentation, such as
poly(hydroxyalkanoates) (PHAs), will be merely mentioned here
because these have been broadly reviewed quite recently.32,33
Additionally, technical aspects of polyesterication processes
and catalysts will be reviewed succinctly.

2.

Renewable monomers

The replacement of petroleum-based monomers by the equivalents coming from renewable sources turned into one of the
paramount challenges in current polymer research. The biomass
components, namely carbohydrates, lignin, oils and proteins,
can be subjected to a set of transformations to yield renewable
chemical building blocks for the preparation of sustainable
polymers.3436 The list of monomers that can be obtained from
biomass is vast and has been recently reviewed.3,35,37 In general,
biomass-derived monomers can be divided according to their
natural molecular biomass origins as: (i) oxygen-rich monomers
namely carboxylic acids, polyols, dianhydroalditols, and furans,
(ii) hydrocarbon-rich monomers including vegetable oils,
fatty acids, terpenes, terpenoids and resin acids, (iii) hydrocarbon monomers such as bio-ethene, bio-propene, bio-isoprene
and bio-butene, and (iv) non-hydrocarbon monomers namely
carbon dioxide and carbon monoxide, as critically discussed by
Yao and Tang in their recent perspective.35
Among the myriad sources of building blocks for polyester
synthesis, carbohydrates, lignin, vegetable oils and, to a lesser
extent suberin, are the most interesting ones. From the stand
point of this review, the following paragraphs are devoted only
to a short excursion into the eld of bio-based aliphatic and
Polym. Chem.

View Article Online

Polymer Chemistry

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

aromatic monomers that are used to synthesize polyesters that


dier in properties and end-uses.
2.1

Aliphatic monomers

2.1.1 Diacids. Polysaccharides, and particularly cellulose


(the most abundant fraction of plant biomass), are considered as
the most important raw materials for the production of
sustainable platform chemicals,38 suitable to prepare dierent
types of polymers (e.g. polyesters, polyamides, polycarbonates,
polyurethanes, polyethers)39 with key applications particularly in
the biomedical eld, due to their biocompatibility and biodegradability. Alditols,39 aldonic and aldaric acids,40 and furans,37
produced from monosaccharides via biological or chemical
conversion are the ones deserving particular consideration as
aliphatic or aromatic monomers for polyester synthesis. A few
important examples of aliphatic dicarboxylic acids obtained
from carbohydrates by the LCF biorenery19 include succinic
acid (SA), fumaric acid (FA) and itaconic acid (IA) (Fig. 1).
SA, or butanedioic acid, was obtained for the rst time from
amber distillation, in 1546, by Georgius Agricola.41,42 This
dicarboxylic acid serves as a precursor to other bulk materials,
namely 1,4-butanediol, g-butyrolactone, tetrahydrofuran, and
adipic acid.42 Nowadays, SA is mainly obtained by chemical
processes that involve the catalytic hydrogenation of maleic
anhydride, followed by hydration, or by direct catalytic hydrogenation of maleic acid.43 In recent years, more attention has
been given to the production of SA through fermentation
processes, and several reviews have been published on this
subject.4145 In the fermentation process, substrates such as
corn starch, whey, cane molasses, glycerol, and wood hydrolysates are transformed into SA through the action of specic
bacterial strains. Actinobacillus succinogenes, Anaerobiospirillum
succiniciproducens,
Mannheimia
succiniciproducens
and
recombinant Escherichia coli are those that allow the preparation of SA in high yields, with lower amounts of by-products (e.g.
acetate and formate derivatives).41,42,46 The fermentative
production of SA is particularly advantageous from an environmental standpoint, since it uses CO2 to convert the
substrates into the product.34,41 Although the fermentation
process brings important benets to the environment, the
optimization of the process is still required, namely in what
concerns the isolation and purication of the product from the
fermentation broth.43 Currently, several companies (e.g. BASF
and Purac, Bioamber, Reverdia, Myriant) are producing biobased SA. Its global production is ca. 50 kton per year, but it is
expected to increase to 270 kton per year in the coming years.47
FA, or (E)-2-butendioic acid, is a naturally occurring dicarboxylic acid that was rst isolated from Fumaria ocinalis.48 FA

Fig. 1

Chemical structures of succinic, fumaric and itaconic acids.

Polym. Chem.

Review
is employed in the preparation of unsaturated polyester resins
and other synthetic polymers. Additionally, it can also be used
as an intermediate for the production of other chemical products (e.g. L-malic acid and L-aspartic acid).4951 FA is currently
obtained from the isomerization of maleic anhydride, but in the
last few years renewed interest has been directed towards its
synthesis from fermentation processes. Curiously, these
processes were used at a commercial scale during the 1940s,48,51
but their low productivity and poor economic competitiveness
dictated their discontinuation in the 1970s.49,51 Presently,
attempts are being made to redeploy the production of FA by
fermentation processes at the industrial level.51 Fungi from
Rhizopus, Mucor, Cunninghamella and Circinella genera are able
to produce FA during fermentation of renewable raw materials,
but only a few species do it in reasonable yields. Rhizopus
species (R. nigricans, R. arrhizus, R. oryzae and R. formosa) have
shown to be the most promising ones, due to their ability to
produce FA both under aerobic and anaerobic conditions.50,51
Although fungi are the most widely used microorganisms in the
production of FA, some bacteria (e.g. Lactobacillus ssp.) and
metabolic engineered yeasts (e.g. Saccharomyces cerevisiae)52,53
have also been used. Regarding the fermentation substrates,
starch and lignocellulosic materials are the most used ones.
IA, also known as methylene SA, is an unsaturated dicarboxylic acid that was identied for the rst time in 1837 during
the thermal decomposition of citric acid.54,55 This diacid can be
used in the synthesis of unsaturated polyesters,5658 which can
nd applications in the coating industry and also in the
biomedical eld. In 1932, Kinoshita observed that an osmophilic strain of Aspergillus species had the ability to produce IA
from carbohydrates.54,59 Since 1960s, IA has been produced from
sugar fermentation using A. terreus, which has shown to be the
most eective microorganism for such purposes. The yield of IA
production is highly dependent on the fermentation substrate,
the highest value being obtained with pure glucose. However,
due to its high price, other carbohydrate sources such as starch,
molasses, corn syrup hydrolysates, wood, and even mixtures of
such components have been tested. Nowadays, the main
producers of IA are located in China, and the annual production
of this diacid is estimated in 30 000 tons.54
Another important carboxylic acid is citric acid (CA), a
tricarboxylic acid with a sterically hindered tertiary hydroxyl
group, which is a common metabolite of plants and animals. CA
is prepared industrially by a fermentation process with A. niger
strains and has an annual production estimated in 9 million
tons.60 Many substrates can be used in the production of CA
such as rapeseed oil, corncobs, and brewery wastes.61 CA is
mainly used in the manufacture of medicinal citrates, confectionary, so drinks and eervescent salts.62
2.1.2 Diols. Some examples of carbohydrate-based
aliphatic diols comprise 1,4:3,6-dianhydrohexitols (DAHs),
namely isosorbide, isomannide and isoidide, 1,3-propanediol
(PDO) and 1,4-butanediol (BDO). DAHs are cyclic dihydroxyethers (Fig. 2) obtained by cyclization through acid dehydration
of hexitols, which are prepared by reduction of hexoses (with
glucose and mannose followed by idose as the most commonly
used).63 DAHs present two hydroxyl groups located at the C2 and

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Review

Fig. 2

Chemical structures of 1,4:3,6-dianhydrohexitols.

C5 positions that dier in their endo or exo stereochemistry,


leading to three possible stereoisomers: isosorbide, isomannide
and isoidide (Fig. 2). In isosorbide, one of the hydroxyl groups is
in the endo position and the other in the exo position, in isomannide, both hydroxyl groups are located in the endo position,
whereas in isoidide, both groups are in the exo position.39,63 The
dierent positions of the hydroxyl groups render the DAHs with
distinct reactivities, isomannide and isoidide being the least
and most reactive, respectively, while isosorbide has an intermediate reactivity. Taking this into account, one could expect
isoidide to be the most used DAH, which is not the case due to
the scarcity of its precursor (L-idose) in nature. In fact, nowadays
isosorbide is the only DAH produced at an industrial level and
inevitably the most exploited in the development of new
applications. Moreover, it is an interesting candidate for the
preparation of polymers, namely polyesters58,6466 with high
glass transition temperatures (Tg).
PDO is a commodity chemical that has attracted a huge
amount of attention due to its vast range of applications (e.g.
polymer synthesis, cosmetics, lubricants, solvents). PDO can be
synthesized by chemical pathways from acrolein or from
ethylene oxide. The chemical methods are dependent on nonrenewable resources, use harsh reaction conditions, and release
toxic by-products.67,68 In order to overcome the main drawbacks
of the chemical synthesis, renewed attention has been paid to
the production of PDO by biological methods from renewable
resources. This route was started when PDO was isolated for the
rst time in 1881, by August Freund, from glycerol fermentation
using a mixed culture of microorganisms in which Clostridium
pasteurianum was the active one.69,70 The use of glycerol as a raw
material for the production of PDO became a major issue given
the market surplus of this chemical as a by-product of biodiesel
production.71 Several bacteria are used in the bioconversion of
glycerol into PDO, among which Klebsiella pneumoniae, Citrobacter freundii and Clostridia butyricum are the most productive. Fungi from the Aspergillus strain (A. niger and A. oryzae)
were also used in the production of PDO, but low yields and
productivities were observed.34,67,68 Usually, the bioconversion of
glycerol into PDO occurs anaerobically, but in the recent years
the aerobic process has been the focus of great attention
because it is more attractive from an economical point of view.68
Another interesting substrate for the production of PDO is
glucose. The main challenge of using this substrate is related to
the fact that no natural microorganisms are available to directly
convert glucose into PDO. To overpass such challenges,
dierent strategies have been attempted, namely the use of
mixed microorganism cultures or the utilization of genetically
engineered microorganisms that are able to transform glucose
into glycerol, followed by the conversion of the latter into
PDO.72,73 Currently, PDO is being produced at an industrial level

This journal is The Royal Society of Chemistry 2013

Polymer Chemistry
by the aerobic fermentation of glucose (from corn starch), by the
joint venture DuPont and Tate&Lyle LLC, with an annual
production of ca. 45 000 tons.34
BDO is mainly produced from maleic anhydride in a process
owned by the company Davy Process Technology. This diol can
be used in the preparation of polymers, but also serves as a
precursor for other chemicals, namely tetrahydrofuran.46,47
Several companies and consortiums, namely BioAmber, Myriant/Davy Process Technology, and Genomatica/Tate & Lyle/Beta
Renewables, are working on the synthesis of BDO from renewable raw materials. The leading methodology involves the
fermentation of sugars into SA, followed by reduction to BDO.47
Nevertheless, recently the preparation of BDO directly from the
fermentation of sugars by using engineered E. coli has been
reported.74
2.1.3 Miscellaneous monomers. Vegetable oils, which
consist mainly of triglycerides, are also perceived as a good
source of building blocks for the preparation of aliphatic polyesters. Most common oils present fatty acids whose chain
lengths vary from 12 to 22 carbon atoms, and with zero to three
double bonds per chain.75 Moreover, some oils can also present
fatty acids with other types of functionalities, namely hydroxyl
or epoxy groups at various positions along the carbon chains.10
The chemical modication of fatty acids further broadens
their structures into a,u-dicarboxylic acids, a,u-dialcohols or
u-hydroxyacids. Currently, vegetable oils are mainly used for
biodiesel production, but given their wide availability, biodegradable nature, and low price, they are also seen as promising
raw materials for monomer and polymer synthesis.76 Soybean,
corn, tung, linseed and castor oils are the most used vegetable
oils for the preparation of monomers and polymers.77 It is worth
pointing out that specialty chemicals from vegetable oils are
already being commercialized. As an illustrative example, Elevance Renewable Sciences recently announced the commercial availability of Inherent C18 Diacid, also known as
octadecanedioic diacid, the rst based on Elevance's proprietary metathesis technology.78 Further details related to oleochemicals are present in the relevant literature.79,80
Suberin and cutin are two other interesting sources of
monomers suitable to prepare hydrophobic and biodegradable
aliphatic polyesters. The rst is a naturally occurring aromaticaliphatic cross-linked polyester almost ubiquitous in the vegetable realm where it plays the fundamental role of a protective
barrier,81 whereas the second is an extracellular aliphatic polyester covering most of the aerial surfaces of plants whose
structure is close to that of aliphatic suberin.82 The hydrolysis of
these natural polymers generates mixtures of long-chain
aliphatic (mainly C18 and C22) a,u-dicarboxylic acids or
u-hydroxyacids, which are suitable monomers for polymer
synthesis.81 Supplementary details regarding suberin-based
monomers are available elsewhere.81

2.2

Aromatic monomers

Regarding aromatic monomers, carbohydrates and lignin (the


second most abundant plant component, also recognized as a
key raw material for sustainable chemicals83) are the major

Polym. Chem.

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Polymer Chemistry

Fig. 3

Conversion of C6-sugars into FDCA.89

sources, with 2,5-furandicarboxylic acid (FDCA) and vanillic


acid (VA) as the most important examples to be highlighted.
FDCA can be used as a substitute of terephthalic acid and
isophthalic acid in the synthesis of polyesters.47,8486 This
dicarboxylic acid results from the catalytic oxidation of 5hydroxymethylfurfural (HMF), which is obtained from the acid
catalysed-dehydration of C6-sugars (mainly fructose).47 The
conversion of hexoses into HMF is still the most challenging
step of this process, mainly due to the low conversion yields and
the intrinsic instability of HMF. Nevertheless, the recognition
that HMF could be one of the key platform chemicals of the
future, which leads to its designation as the sleeping giant,87
has triggered huge amounts of research in recent years in order
to maximise conversion yields and minimize subsequent
degradation. The conversion of HMF into FDCA comprises two
steps: conversion of HMF to a dialdehyde derivative, followed by
the oxidation of the two aldehyde groups to carboxylic acid
groups (Fig. 3).88
Some decades ago, the oxidation of HMF to FDCA was
carried out using nitric acid, potassium permanganate or silver
oxide, in organic solvents (e.g. DMSO).90 In the recent years, the
oxidation reactions are essentially carried out with more ecofriendly oxidants and solvents, using heterogeneous metal
catalysts.89,90 Metal catalysts such as platinum supported on
alumina (Pt/Al) or carbon (Pt/C) have demonstrated to be very
ecient in the conversion of HMF to FDCA, in water under
atmospheric oxygen pressure. Nevertheless, in such reactions,
high catalyst/substrate ratios are still needed to obtain high
yields.86 Gold nanoparticles supported on cerium oxide or titanium oxide have shown to be excellent catalysts in these
oxidation reactions, under mild conditions, using water or
slightly basic aqueous solutions as solvents.86,89 Another recent
strategy proposes the use of enzymes for the bioconversion of
HMF into FDCA.91 Currently, FDCA is being industrially
produced by Avantium and the company is expecting a potential
market for this biobased dicarboxylic acid of ca. 100 000 ton per
year, with a price below 1000 V per ton.47
Lignin is a polymer that plays the role of matrix surrounding
the hemicelluloses and cellulose in all vascular plants.92 Lignin
is biosynthesised from three basic monolignol structures (coumaryl alcohol, coniferyl alcohol and sinapyl alcohol),92 which
become chemically bonded in a panoply of ways, leading to an
aromatic complex, amorphous, cross-linked structure. This
natural polymer is an interesting source of macromolecular
materials as well as of phenolic monomeric units.36
Vanillin is an aromatic aldehyde that was isolated for the rst
time from vanilla extract derived mainly from orchids from the
genus Vanilla planifolia.93 Currently, vanillin is mostly obtained
through the air oxidation of lignin in the presence of metal
catalysts.9496 The biotechnological production of this building

Polym. Chem.

Review
block by fungi, bacteria, and genetically modied microorganisms is also possible and has been reviewed.93,97 The interest on
vanillin in the present context relies on its chemical or biotechnological98 conversion into vanillic acid, another monomer of
interest in the preparation of polyesters.99,100 This hydroxyacid
can also be obtained by the bioconversion of ferulic acid.101,102

3.

Renewable aliphatic polyesters

Aliphatic polyesters represent one of the most relevant groups


of polyesters, which are preferentially obtained via polycondensation reactions of alkanediols with a,u-alkanedioic
acids (or alternatively derived from hydroxyalkanoic acids)
readily available from inexpensive renewable resources. The
biodegradability of aliphatic polyesters has spotlighted them,
since in practice only this family of polyesters degrade in a
reasonable time scale.
Among the aliphatic polyester family, PLA is one of the most
promising industrial commodities for conventional plastics with
the potential to replace polymers such as PET, PS and PC.32,103
PLA can be produced with high molecular weights by ringopening polymerization (ROP) of the corresponding lactide
(Scheme 1), presenting typically glass transition temperatures
ranging from 50 to 80  C, whereas its melting temperature ranges
from 130 to 180  C depending on the degree of crystallinity and
molecular weight.104 PLA main applications are in short-term
packaging, due to its biodegradability, and also in biomedical
applications such as implants, sutures, or drug encapsulations
due to its biocompatibility in contact with living tissues.104
Currently, NatureWorks LLC (Blair, Nebraska, USA) is the
leading supplier of PLA biopolymer under the brand name
Ingeo with a production capacity of 140 000 metric tons per
year.22 Several interesting reviews32,33,105 have already exhaustively covered this topic; hence and, as mentioned above, the
present appraisal will not focus on it any further. Nevertheless, it
is important to emphasize that the scientic and industrial
research devoted to this eco-friendly thermoplastic, whose use
remains limited due to the high production costs, is currently
oriented toward the improvement of specic properties, namely
its poor toughness, hydrophobicity, and lack of reactive sidechain groups, in order to widen the areas of its applications.104,106
PHAs, a class of polymers naturally produced by a variety of
microorganisms as storage materials, are another important
class of polyesters that have been identied as promising
alternatives to conventional hydrocarbon-based polymers.107
Bio-On (Bologna, Italy) is currently the main supplier of various
types of PHAs under the brand name Minerv-PHA with a
production capacity of 10 000 metric tons per year.108 The
simplest PHA is poly(3-hydroxybutyrate) (PHB) and is the one
receiving the most attention as a target molecule for production

Scheme 1

Ring-opening polymerization of lactide.

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Review

Polymer Chemistry

from plant cells.107 Applications of PHAs span from plastics,


printing and photographic materials, drugs and ne chemicals,
to medical implants, drug delivery carriers and nutritional
supplements.109 PHAs do not require any extensive coverage
here given the recently published reviews on all aspects related
to these materials.107,109
Other remarkable examples of aliphatic polyesters from
renewable resources include the ones obtained e.g., from
vegetable oils, suberin, cutin, sugars, polycarboxylic acids and
their corresponding derivatives, as discussed in detail below.
Other less-common polyesters worth pointing out include e.g.,
the ones obtained from aminoacids110,111 which may serve as
potential biocompatible materials for medical applications, or
terpenoids112,113 which display shape memory properties.
3.1

Vegetable oil-based polyesters

Among the most interesting vegetable oil-based monomers,


ricinoleic acid (12-hydroxy-9-cis-octadecenoic acid) obtained
from castor oil occupies undoubtedly a privileged position due
to its bifunctionality,10,16 oppositely to most naturally occurring
aliphatic acids having only one functional group, and that
therefore cannot be used in polyester synthesis before any
intermediate chemical functionalization/transformation step is
carried out.
The use of ricinoleic acid to produce homopolyesters is quite
rare. Indeed, only a reference to poly(ricinoleic acid) prepared
by lipase catalysed condensation of methyl ricinoleate (Scheme
2) was made by Ebata et al.,114 which was subsequently crosslinked to produce thermosetting elastomers. This work has
the merit of having used both a renewable monomer and an
eco-friendly synthesis approach, i.e. bulk enzymatic polytransesterication using several lipases. The ensuing polyesters
presented weight-average molecular weights between 2 and
100 kDa with the best results being obtained with a Pseudomonas cepacia lipase. For the highest molecular weight
polymers, glass transition temperatures around 75  C were
observed. Aer radical cross-linking, the ensuing product had a
gel fraction of approximately 98% and a glass transition
temperature of about 65  C.114 Furthermore, it has been also
demonstrated that poly(ricinoleic acid) can be easily recycled by
enzymatic hydrolysis followed by repolymerization.115
Still in the eld of bifunctional monomers but now using
9-hydroxynonanoic acid, obtained from castor oil aer
ozonolysis, followed by reduction and transesterication, Petrovi
c et al.116 synthesized high molecular weight linear polyesters
with higher melting point and glass transition temperature,
better thermal stability and lower solubility in chlorinated
solvents than the analogous poly(3-caprolactone) (PCL).
Furthermore, a considerable number of studies have dealt
with copolyesters of ricinoleic acid with, notably and among

Scheme 2

Lipase-catalysed synthesis of poly(ricinoleate).114

This journal is The Royal Society of Chemistry 2013

others, PLA117119 and sebacic acid,120,121 which, have been used


as biodegradable delivery systems in various biomedical applications. For example, Slivniak and Domb117 focused their work
on the synthesis of a ricinoleic acidlactic acid copolymer using
two dierent approaches viz. direct polycondensation and polytransesterication. The ensuing copolyesters were obtained
with variable molecular weights depending on the synthesis
approach used (number average molecular weight around 2
11 kDa), and may be used as sealants and as injectable carriers
of drugs.
Other original contributions describe the preparation of longchain aliphatic semicrystalline polyesters from fatty acids with
systems in which a fatty acid or ester is converted into a saturated
a,u-diacid or diester, which in turn is reduced to its diol counterpart in order to obtain the complementary monomer with
which the corresponding polyester is synthesized by esterication or transesterication. Mecking et al.122126 have extensively
studied the preparation of linear saturated semicrystalline
polyesters with long-chain hydrocarbon segments. For example,
they synthesized aliphatic long-chain C19, C20, C23 and C26
monomers generated from oleic acid, 10-undecenoic acid or
erucic acid, by conversion into a,u-diacids via carbonylation122,124 and olen metathesis123,125 followed by reduction to
a,u-diols, which were then polymerized to yield saturated polyesters with high melting points. Vilela et al.127 prepared renewable long-chain aliphatic polyesters from erucic acid using olen
self-metathesis for monomer synthesis, as depicted in Scheme 3.
Polycondensation of the obtained 1,26-hexacosanedioic acid
with a,u-diols yielded the desired long-chain aliphatic polyesters
with high crystallinity, melting and degradation temperatures.
Following this idea, Mecking's team also prepared C26 monomers and polymerized them to obtain a polyester featuring a
melting point higher than the model ultralong-chain polyesters
that they investigated.125 All the above studies122127 yielded
aliphatic polyesters with chemical and physical properties which
can mimic polyethylene or other polyolens.
Yang et al.128 obtained unsaturated and epoxidized fatty acidbased polyesters from a,u-dicarboxylic acids, prepared by
biochemical oxidation of terminal methyl groups of unsaturated fatty acids into carboxylic groups, followed by enzymatically catalyzed polycondensation (Scheme 4). The obtained
unsaturated polyesters can be further modied, derivatized, or
cross-linked, using the mid-chain unsaturation in order to
adjust the material properties. Zhang et al.129 developed unique
cross-linkable unsaturated polyesters, by a one-pot lipase-catalyzed oxidation followed by copolymerization of crude linoleic

Scheme 3 Synthetic pathways for the conversion of erucic acid into 1,26-diacid
and corresponding 1,26-diol and subsequent polycondensation between
them.127

Polym. Chem.

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Polymer Chemistry

Review

Scheme 5 Hyperbranched polyester from a glycerol-based AB2 type monomer;


R H or polymer unit.143

the bulk. The ensuing hyperbranched polyester was used as a


cross-linking agent for acrylic ester latex.

3.2

Scheme 4 Biochemical conversion of unsaturated fatty acids into a,u-diacids


and subsequent polycondensation with a,u-diols.128

acid, glycerol and 1,18-cis-9-octanedecenedioic acid. Following a


similar approach, Roumanet et al.130 prepared aliphatic unsaturated polyesters derived from the 1,18-(Z)-octadec-9-enedioic
acid, a sunower oil fatty acid derivative, and aliphatic diols of
dierent molecular weights. The double bonds of the ensuing
polyester backbones were then epoxidized and photo-crosslinked, yielding transparent and homogeneous cross-linked
lms with high hydrophobicity.
In a dierent vein, several authors have reported the
successful preparation of aliphatic polyesters by the application
of alternative methods e.g. thiolene131 and metathesis132 polymerizations. As an illustrative example, Akintayo et al.133
showed that Plukenetia conophora oil can be used for the
synthesis of hyperbranched polyesters, with interesting rheological properties, via olen metathesis (acyclic triene metathesis, ATMET) polymerization. T
ur
un and Meier134 synthesized
a number of aliphatic polyesters, in the form of hyperbranched,
dendritic, and linear chains, from a castor oil derived renewable
platform chemical using thiolene&yne addition reactions
featuring click chemistry conditions.
Also in the eld of oleochemistry, glycerol is currently
available in high amounts mostly because of biodiesel
production from vegetable oils.71 As a consequence, the surplus
glycerol is being used as an intermediate in the synthesis of a
large number of industrial commodities135 and a considerable
amount of research has been focused on its use as a platform
chemical to replace some mainstream petroleum-derived
chemicals.136 Its utilization was broadly reviewed recently137,138
and includes, for example, the application as a green solvent,139
and in the production of hydrogels140,141 and syngas.142
Furthermore, recent research also reported the synthesis and
properties of a series of glycerol-based polyesters. For example,
Parzuchowski et al.143 reported the synthesis of a glycerol-based
AB2 type monomer (ethyl{3-[2-hydroxy-1-(hydroxymethyl)ethoxy]propyl}thioacetate) for the preparation of hyperbranched polyesters that easily undergo hydrolysis or
alcoholysis, making them interesting as recyclable materials
(Scheme 5). Zhao et al.144 synthesized a water-soluble hyperbranched polyester with a considerable number of hydroxy
terminal groups by reacting maleic anhydride and glycerol in

Polym. Chem.

Suberin and cutin-based polyesters

Other attractive sources of monomers to produce aliphatic


polyesters are suberin145147 and cutin.148,149 There are already
some reports on the use of long chain hydroxyalkanoic acids
isolated from suberin or cutin or, instead, using a mixture of
monomers to produce polyesters. The pioneering work of Olsson et al.150 in the eld reported the use of 18-hydroxy-9,10epoxyoctadecanoic acid, isolated from birch outer bark suberin,
to prepare polyesters under mild conditions using the Candida
antarctica lipase B (CALB) as a catalyst (Scheme 6). The ensuing
polyester presented interesting properties as the persistence of
its epoxy function and high molecular weights (up to 20 kDa).150
However, this approach requires very laborious, costly and not a
very green processing of suberin depolymerization mixtures
and purication steps aimed at obtaining pure compounds.
A similar study by Heredia-Guerrero et al.148 focusing on the
polycondensation reaction of a single cutin monomer was
recently published. The most abundant cutin component,
9(10),16-dihydroxyhexadecanoic
acid,
was
successfully
self-polymerized on a mica surface, despite the intrinsic stoichiometric imbalance of reactive groups (1CO2H/2OH) of
9(10),16-dihydroxyhexadecanoic acid. The same authors also
reported a study on the emulsion polycondensation of a cutin
mimetic polyester using a very similar monomer, 9,10,16-trihydroxyoctadecanoic acid (Scheme 7).149 The detailed characterisation of this polyester showed that, similar to its cutin
counterpart, it was mostly an amorphous polymer as conrmed
by X-ray diraction spectroscopy. The dierential scanning
calorimetry (DSC) thermogram of the resulting polyesters
showed a glass transition around 2  C and two small endothermic events around 46 and 73  C, that could be related to the
presence of larger short-range ordered domains. This diered
from cutin which presented a lower glass transition temperature (47  C) and no endothermic peaks.

Scheme 6 Synthesis of poly(18-hydroxy-9,10-epoxyoctadecanoate) catalysed


by CALB.150

Scheme 7 Synthesis of poly(9,10,16-trihydroxyhexadecanoate). R and R0 stand


for either H or another cross-linked fragment.149

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Review

Polymer Chemistry

Well-characterized151,152 samples of depolymerized suberin


from cork, mainly composed of long chain u-hydroxyacids and
a,u-dicarboxylic acids, were also used to prepare new polyesters
by polycondensation under green conditions, namely emulsion
polymerization or using CALB.147 In another study, Sousa et al.146
reported the preparation and characterization of several novel
polyesters from suberin depolymerization mixtures obtained
from both cork and birch outer barks, using in some cases an
extra source of functional groups (diols or diacids) to balance
the OH/COOH stoichiometry. The ensuing linear or partly crosslinked polyesters, obtained in good yields under mild reaction
conditions, displayed properties that resemble those of petroleum-based aliphatic polyesters.146 Recently, the concomitant
use of an approach bearing eco-friendly features and fast reaction times through the use of microwave irradiation led to the
formation of hydrophobic aliphatic polyesters in reasonable
yields.145 These polyesters displayed well-dened melting peaks,
typically at temperatures around 80  C, in agreement with those
reported for other long-chain aliphatic polyesters.153
3.3

Sugar diol-based polyesters

1,4:3,6-Dianhydrohexitols (DAHs), namely isosorbide, isomannide and isoidide, have deserved particular consideration
as monomers for polyester synthesis, as a result of the attractive
features of the ensuing polyesters such as rigidity, chirality,
non-toxicity, and renewable nature;39 however, the low reactivity
of the secondary and relatively hindered hydroxyl groups leads
in general to polyesters with low molecular weights. Still, isosorbide and its isomers can be used directly to prepare polyesters with high glass transition temperature and/or with
special optical properties.39 DAHs have been widely (co)polymerized with several diacids including, e.g. suberic acid,154 1,4cyclohexane dicarboxylic acid,155 1,4-cyclohexane dicarboxylic
acid and succinic acid,156 itaconic acid and succinic acid,58
succinic acid and adipic acid,64 among many others.39
Important recent additions to the eld of aliphatic polyesters
from DAHs include the work of Wu et al.157 on the melt polymerization of a series of rigid bio-based polyesters solely based
on renewable isohexide building blocks for performance polymers (Scheme 8), in which the diacid, isoiodide dicarboxylic
acid, and the diol, isoidide dimethanol, were prepared by a
multistep process from isomannide. Goerz and Ritter58
prepared fully bio-based unsaturated oligoesters from

Scheme 9 Polymerization reactions leading to random copolyesters from 1,4butanediol, bicyclic 2,4:3,5-di-O-methylene-D-mannitol and dimethyl succinate.160

Scheme 10 Synthesis
polymerization.161

Synthesis of polyesters from isohexide-based monomers.157

This journal is The Royal Society of Chemistry 2013

D-glucono-d-lactone

and

its

ring-opening

isosorbide, IA and SA which showed a shape memory eect aer


cross-linking with dimethyl itaconate.
Other interesting studies make use of other sugar derived
diol monomers. For example, Lavilla et al.158,159 reported the
synthesis and characterization of several aliphatic polyesters
with high solubility and wettability made from dimethyl 2,3:4,5di-O-methylene-galactarate, a bicyclic monomer obtained by
internal acetalization of galactaric acid. In another study, Lavilla et al.160 prepared a series of random copolyesters from
1,4-butanediol, bicyclic 2,4:3,5-di-O-methylene-D-mannitol and
dimethyl succinate by a two-step melt polycondensation process
(Scheme 9) whose properties are comparable to the copolyesters
prepared from isosorbide.
In a dierent approach, Tang et al.161 prepared an original
functionalized cyclic aliphatic polyester (Scheme 10) with
number average molecular weights ranging from 1800 to 7300
Da through the ROP of a new carbohydrate lactone derived from
D-gluconolactone in the presence of a tin butoxide initiator.
Kakasi-Zsurka et al.162 reported the production of bio-based
oligomeric polyesters from 3-hydroxybutyric acid and D-glucono-d-lactone via lipase catalysed ROP without previous
derivatization of the sugar moiety, which led to the presence of a
substantial number of free OH groups. The ensuing oligomers
containing inserted 2,3,4,5,6-pentahydroxy-caproate units were
considered as possible candidates for industrial or medical
applications.
3.4

Scheme 8

of

Other aliphatic polyesters

Within the vast array of monomers from renewable resources,


carboxylic acids such as SA and IA have also received serious
attention and produced interesting materials. Other worth
mentioning, but less-common carboxylic acids include e.g. tartaric acid (TA) and citric acid (CA). Despite the numerous
contributions discussed below, the potential of these organic
acids in polyester chemistry is still awaiting to be thoroughly
exploited.

Polym. Chem.

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Polymer Chemistry
SA is one of the platform chemicals for bio-based polymers
from renewable resources.163 The most attractive recent studies
regarding SA deal with the synthesis of a large spectrum of
monomeric structures, and also polyamides, polyesters, etc.163
On the topic of polyesters, Showa Denko fabricates compostable
and biodegradable poly(butylene succinate) (PBS) and poly(butylene succinate adipate) (PBSA) under the trade name
Bionolle (1000 series (PBS) and 3000 series (PBSA)), which are
commercial thermoplastic polyesters with good manufacturing
properties, exibility, and tenacity, comparable to those from
conventional plastics.24
Some interesting publications in the eld of SA-based polyesters comprise the study of Noordover et al.66 on co- and terpolyesters based on SA and isosorbide in combination with
other renewable monomers such as 2,3-butanediol, PDO and
CA. The obtained linear and branched polyesters presented
suitable properties with respect to solvent resistance, impact
resistance, and hardness, making them very promising materials for coating applications. Furthermore, these polyesters
also proved to be appropriate materials for solvent-cast and
powder coatings.164 Mincheva et al.165 synthesized poly(butylene
succinate-co-butylene azelate) aliphatic copolyesters from SA,
BDO, and dimethylazelate via a two-step polycondensation
reaction, as exemplied in Scheme 11. These environmentally
friendly copolyesters might nd applications from elastomers
to high-impact thermoplastics.
IA is another important chemical for the development of
polyesters. As an illustrative example, Barrett et al.56 reported
several polyester thermosets based on photocurable prepolymers composed of IA and various polyols. The enzymatic
synthesis of these itaconate-based polyesters yielded materials
that might nd wide use in the biomedical and biotechnological elds. Guo et al.57 prepared bio-based polymers with shape
memory from IA, sebacic acid, and PDO with adjustable
switching temperature and recovery speed. The biocompatibility and biodegradability of these polyesters make them suitable for fabricating biomedical devices. Finally, Okuda et al.166
used itaconic anhydride and lactic acid as starting materials for
the synthesis of bio-based copolymers with possible applications for coatings and plastics.
TA, a widely available and relatively cheap diacid from a large
variety of fruits, is an attractive monomer for the synthesis of
functional polymers such as polyamides,167,168 polycarbonates,169,170 polyesters,171174 etc. TA-based polyesters nd
applications mainly as carriers for the controlled release of

Review

Scheme 12

Synthesis of a triblock copolymer based on TA and lactide.173

Scheme 13
polyester.183

Polycondensation of glycerol with CA to produce glycerol citrate

drugs.175 As an illustrative example, this prolic molecule and


its derivatives have been used by Jacob's group for the synthesis
of a series of aliphatic polyesters: (i) polyesters with pendant
hydroxyl groups from dimethyl 2,3-O-isopropylidene-L-tartarate,171 (ii) hydroxy-functionalized copolyesters of dimethyl 2,3O-isopropylidene tartarate,172 (iii) ABA type triblock copolymers
based on L-tartaric acid and L-lactide (Scheme 12),173 and (iv)
biodegradable chain-coupled polyesters synthesized from TA
derived hydroxyl terminated polyesters with hexamethylene
diisocyanate.174
CA has been employed as a low cost monomer in the
synthesis of oligoesters176 and biodegradable polyesters for
tissue engineering,176179 drug delivery,180 antimicrobial181 and
high-tech coating182 applications, among others. Some interesting recent studies include the work by Chandorkar et al.180 on
the melt-condensation copolymerization of citric, sebacic and
ricinoleic acids, and mannitol as precursors, yielding polyesters
that can be potential candidates for drug delivery and tissue
engineering applications, given their demonstrated biocompatibility and the work by Tisserat et al.183 on the rapid, inexpensive and easy synthesis of solid glycerol citrate polyesters,
through microwave irradiation without any catalyst, for various
polymer applications (i.e., moulding) (Scheme 13). Very
recently, Halpern et al.184 prepared a degradable thermoset
polyester capable of drug delivery from glycerol and CA by a
melt polymerization reaction in the absence of a catalyst.

4.

Scheme 11 Synthetic pathway to poly(butylene succinate-co-butylene azelate)


copolyesters.165

Polym. Chem.

Renewable aromatic polyesters

As mentioned above, the aliphatic polyesters and copolyesters


are currently among the most promising biodegradable polymers but lack important properties to completely replace
conventional plastics. Conversely, the aromatic polyesters are
the most studied and used, and denitely, have a higher market
share, essentially due to their high performance properties
(high molecular weights, melting and glass transitions, or
mechanical behaviour). Important commercial examples of

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Review
aromatic polyesters include the high performance thermoplastic PET and poly(butylene terephthalate) (PBT). However,
their precursors are fossil based and they are resistant to
microbial attack and not degradable under normal environmental conditions. Thus, a signicant eort in the aromatic
polyester eld is devoted: (i) to mimic their appealing properties
using monomers of renewable origin instead of the petrochemical counterparts, and (ii) to modify them to be either
chemically or biologically degradable. Furthermore, emphasis
is also being placed on aliphaticaromatic polyesters that are
expected to show both good biodegradability185 as well as good
material properties.
Several companies are already commercializing biodegradable or partially renewable-based aromatic polyesters. BASF has
been commercializing poly(butylene adipate-co-terephthalate)
(PBAT) with the brand name Ecoex for more than a decade.
This fossil-based compostable polyester has a melting
temperature of 110120  C, good mechanical properties,
excellent thermal stability, and is suitable for numerous
applications, such as mulch lms for agriculture, cling-wrap
lms or coatings for packaging and breathable lms in the
hygiene sector.186 Moreover, there have been some studies on
the biodegradation of PBAT using several plastic-degrading
enzymes that bear out its biodegradability through biochemical
monomer recycling.187189 The plastics giant DuPont made
available partially renewable terephthalate based polyesters,
namely Sorona EP (PTT) and Hytrel RS (PBT-b-PTMG),
which contain, respectively, PDO derived from corn sugar and
polyether glycols made from non-food biomass.26 The former is
mainly applied in electrical/electronic devices, automotive
components, appliance parts, furniture, food contact
compliant grade materials, whereas the latter nds key applications in hose and tubing, air bag doors, energy dampers,
sporting goods, among others.26
Another example worth mentioning comes from The CocaCola Company that is currently using the rst-generation
PlantBottle, a fully recyclable PET bottle composed of 30%
plant-based material (Fig. 4).27 Despite the use of bioethanol as
a bio-based precursor of ethylene glycol, the PET PlantBottle
possesses the same degradation behaviour as its fossil counterpart.190 It is noteworthy that, although the packaging PlantBottle pilot has already ended, the company still has a huge
interest in producing the 100% biorenewable PET PlantBottle
2.0,191 that will inevitably involve the production of terephthalic
acid based on renewable starting chemicals.3,34 Additionally, the
Coca-Cola Company together with Danone and ALPLA are
collaborating in the Avantium's Joint Development Platform for

Fig. 4 Synthetic route to the rst-generation PET PlantBottle27 composed of


30% plant-based material.

This journal is The Royal Society of Chemistry 2013

Polymer Chemistry
bottles based on PEF,25 the 100% bio-based and recyclable
polymer that will hopefully replace conventional PET bottles.
A variety of aromatic monomers have been identied as
product
opportunities
namely
polysaccharide-based
compounds such as furan derivatives and lignin-based
compounds.21,36 The emphasis of the publications surveyed in
the following paragraphs will be on renewable-based aromatic
polyesters and their co-polyesters from furans (e.g. FDCA) and
lignin (e.g. vanillic acid, caeic acid and ferulic acid).
4.1

Furan-based polyesters

The realm of furan polyesters has reached remarkable proportions in terms of the number of investigations, as pointed out by
comprehensive reviews available elsewhere.1,2,192 Hence, this
section does not intend to cover exhaustively the numerous
mentions to furan polyesters but rather to highlight the latest
and most relevant examples among the vast catalogue.
Aer the ground-breaking work of Moore and Kelly about the
synthesis of a number of linear polyesters from FDCA or the
corresponding dichloride,193,194 some groups have devoted
considerable attention to furan polyesters in the last couple of
years.84,85,195208 For example, Gandini et al.200202 have demonstrated that FDCA can be used in polycondensation reactions to
produce PEF and other FDCA based (co)polyesters. In the case
of PEF, the similarity with PET in terms of thermal properties
(e.g. glass transition, melting and the onset of thermal decomposition temperatures around 80, 215 and 300  C, respectively)
was unambiguously demonstrated.200
The realm of FDCA can be extended to a more vast number of
polyesters besides PEF, such as the ones prepared by using
FDCA and three isomers of 1,4:3,6-dianhydrohexitol, PDO or
hydroquinone.202 Moreover, the polycondensation using
dierent linear diols (C2C8)203 or the random copolymerizations of FDCA with dierent amounts of ethylene glycol and
butanediol204 have also been reported.
Notwithstanding the enormous potential of the polyesters
based on FDCA, recently renewed interest in terephthalic acid
(TPA) emerged because of the sugar-based p-xylene route to
prepare 100% renewable TPA.3,34 Thus, Sousa et al.84 reported
the partial substitution of non-renewable PET with copolymers
of TPA and FDCA derivatives, as epitomized in Scheme 14.

Scheme 14 Polytransesterication reaction of bis(2-hydroxyethyl) terephthalate


and bis(hydroxyethyl)-2,5-furandicarboxylate to obtain PET-ran-PEF copolyesters.84

Polym. Chem.

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Polymer Chemistry

Review

Scheme 16

Scheme 15 Multistep conversion of furfural into the poly(butylene 2,5furandicarboxylate).208

These PET-ran-PEF copolyesters showed average molecular


weights, thermal and mechanical properties comparable to
petrochemical derived materials.
Other recent studies in this eld include the work of Zhu
et al.206 on polyesters based on FDCA and BDO that might be
promising furan alternatives to PBT, since poly(butylene 2,5furan dicarboxylate)s possess a crystal structure, mechanical
and thermal properties similar to those of PBT. Moreover, Wu
et al.207 prepared bio-based aliphaticaromatic copolyesters,
namely poly(butylene succinate-co-butylene furandicarboxylate)s (PBSFs), ranging from crystalline thermoplastics to
amorphous elastomer-like polymers depending on the
composition, and possessing excellent thermal and mechanical
properties.207 Gubbels et al.85 prepared polyesters from FDCA or
2,5-dimethyl-furandicarboxylate and 2,3-butanediol via bulk
polycondensation, which may be suitable for coating applications. The fully bio-based polyesters displayed number average
molecular weights from 2 to 7 kDa and glass transition
temperature values between 70 and 110  C, and are thermally
stable up to 270300  C.
As a nal and noteworthy contribution, Pan et al.208 developed
a route to produce a FDCA-based polyester fully derived from
furfural and with total carbon utilization. The four step route
comprised the (i) oxidation of furfural into furoate; (ii) disproportionation of furoate to furan and 2,5-furandicarboxylate; (iii)
hydrogenation and hydrolysis of furan to 1,4-butanediol; and (iv)
polymerization of 1,4-butanediol with FDCA to poly(butylene
2,5-furandicarboxylate), as illustrated in Scheme 15.
Lastly, it is worth mentioning that the studies published so
far bear out the potential of furan-based polyesters as promising
chemicals from renewable resources to replace those obtained
from non-renewable feedstock.
4.2

Synthesis of lignin-based PET analogue poly(dihydroferulic acid).209

biorenewable poly(dihydroferulic acid) (PHFA) (Scheme 16), a


polyester with thermal properties and functionally simulating
those of PET,210 and therefore demonstrates the huge potential
of lignin-derived monomers for the preparation of aromatic
polymers.
Recently, and as an extension of these investigations, Mialon
et al.100 prepared other biorenewable aromatic polyesters using
aromatic hydroxyacids obtained from lignin. Poly(alkylene
4-hydroxybenzoate)s, poly(alkylene vanillate)s, and poly(alkylene
syringate)s were prepared via bulk polycondensation of the
corresponding alkylene hydroxybenzoates and exhibited a high
thermal robustness in the temperature range of 407478  C.100
Firdaus and Meier211 prepared thermoplastic aromatic polyesters with good thermal stability using vanillin together with
fatty acid derivatives via three dierent polymerization reactions, namely acyclic diene metathesis (ADMET), thiolene
addition, and polycondensation. They reported that the polyesters obtained by ADMET polymerizations presented higher
molecular weights (up to 50 kDa) than the ones obtained by
thiolene polyaddition and polycondensation (15 and 17 kDa,
respectively).211
Lovell et al.212 produced a series of random aromatic
aliphatic thermotropic copolyesters derived from p-hydroxybenzoic acid, vanillic acid, 4,40 -sulfonylbis(2-methylphenol),
and adipic acid. The investigation of their phase behaviour,
morphology, and molecular relaxation behaviours showed their
potential use as bioresorbable polymers for orthopaedic
medical applications.
The investigations of Dong et al.213 on caeic and lithocholic
acid-based copolyesters (Scheme 17) led to the development of
branched and biodegradable polymers with controllable photocrosslinkability, i.e. poly(caeic acid-co-lithocholic acid) with

Lignin-based polyesters

In this section, attention will be focused on a couple of monomer precursors, such as vanillic, caeic and ferulic acids, that
globally will behave as hydroxyacids in what concerns to polyester synthesis.
In a very original study, Mialon et al.209,210 used vanillin
obtained from lignin and acetic anhydride to prepare 4-acetyl
dihydroferulic acid which was then used to prepare the

Polym. Chem.

Scheme 17

Synthesis of poly(caeic acid-co-lithocholic acid).213

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Review

Polymer Chemistry

Scheme 19

Synthesis of poly(glycerol-sebacate-curcumin) polyester.217

Scheme 20

PBT obtained by polycondensation of bio-based TPA and BDO.218

Scheme 18 Aromaticaliphatic copolyesters from rapeseed-derived renewable


starting materials.215

high thermal stability. The same authors also reported the


synthesis of poly(caeic acid-co-hydroxycapric acid).214 Both
copolyesters may nd applications in the environmental and
biomedical areas given their high thermal stability, good
mechanical properties, and photocrosslinkability.213,214
Very recently, Kreye et al.215 prepared (co)polyesters from
ferulic acid- and rapeseed-derived starting monomers via basecatalyzed transesterication reactions (Scheme 18) under the
requirements of green chemistry. The ensuing thioetherbearing (co)polyesters were then oxidized to sulfone analogues
with adjustable thermal properties.
4.3

Other aromatic polyesters

Recent studies also reported the preparation of aromatic polyesters from renewable resources other than furans and lignin
derivatives. For instance, Matsumi et al.216 used curcumin as an
inherent natural diol monomer for the synthesis of polyesters
bearing b-diketone functionalities in their main chain. The
resulting copolyesters should be valuable building blocks for a
variety of functional polymermetal complexes, as demonstrated
by the prepared polyesterruthenium complex. In a dierent
vein, Sun et al.217 also used curcumin but now to prepare
poly(glycerol-sebacate-curcumin) polyesters (Scheme 19), which
show in vitro antitumor properties against brain cancer cells due
to their inherent biocompatibility and degradability which
enhance the availability of the active component, i.e. curcumin.
Other studies either make use of renewable-based TPA from
e.g. limonene, used to prepare PBT,218 or fossil-based TPA which
is then polymerized with bio-based monomers e.g. sugar-based
compounds,219224 and fatty acid derivatives.225
Colonna et al.218 synthesized terephthalate polyesters from
renewable resources, namely dimethyl terephthalate from
limonene and BDO from SA (Scheme 20). As expected, these biobased polymers possess thermal properties comparable to those
of commercial terephthalate polyesters of non-renewable origin.

This journal is The Royal Society of Chemistry 2013

Scheme 21 Polycondensation of DAHs with various aliphaticaromatic acid


dichlorides.219

Abderrazak et al.219 prepared cyclic and acyclic aliphatic


aromatic polyesters based on 1,4:3,6-dianhydrohexitols and
various aliphaticaromatic acid dichlorides with dierent
alkenyl chain lengths (Scheme 21). Although the use of
hazardous reagents and solvents was not minimized during the
solution polycondensation step, the novel polyesters with exible and rigid chains were obtained in high yields with high
inherent viscosities and glass transition temperatures.
Chatti et al.220 synthesized a new class of biodegradable
copolyesters from isosorbide, SA and isophthalic acid, whose
glass transition temperatures (90160  C) can rival those of
most commercial amorphous engineered plastics. Japu et al.221
prepared D-glucose-derived PET copolyesters via bulk polycondensation of 2,4:3,5-di-O-methylene-D-glucitol and dimethyl
2,4:3,5-di-O-methylene-D-glucarate with dimethyl terephthalate
and ethylene glycol. These copolyesters exhibited an enhanced

Polym. Chem.

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Polymer Chemistry

Review

Scheme 22 Polymerization reactions leading to statistical copolyesters from


BDO, 2,4:3,5-di-O-methylene-D-mannitol and dimethyl terephthalate.223

glass transition temperature, a higher hydrolysis rate than


PET homopolymer and an appreciable susceptibility towards
biodegradation.
Other original additions come from Lavilla et al.222224 that
also devoted some attention to sugar-based aromatic polyesters.
First, they prepared semicrystalline and thermally stable
random poly(hexamethylene terephthalate-co-galactarate) and
poly(dodecamethylene
terephthalate-co-galactarate)
copolyesters with weight-average molecular weights of 3050 kDa by
melt polycondensation.222 This study showed that the incorporation of sugar units in aromatic polyesters yielded copolyesters
more easily degradable and also biodegradable while retaining
their thermal and mechanical properties at acceptable values.222
In another study, a set of statistical copolyesters were successfully synthesized from dimethyl terephthalate, BDO and
the biobased 2,4:3,5-di-O-methylene-D-mannitol monomer
(Scheme 22).223 The use of this sustainable compound yielded
copolyesters with satisfactory molecular weights and with
improved thermal and mechanical properties. Furthermore,
they carried out a comparative study regarding the use of isosorbide and diacetalized alditols as sustainable comonomers of
dimethyl terephthalate and BDO in terms of their reactivity and
of the properties of the resulting copolyesters.224 The higher
reactivity of diacetalized diols compared to isosorbide was
demonstrated and the partial replacement of BDO by sugarbased diols leads to copolyesters with increased glass transition
temperatures when compared to those of PBT.224
A dierent contribution was reported by Gross's group who
explored a unique family of copolyesters consisting of units
derived from u-hydroxytetradecanoic acid, 1,4-butanediol and
dimethyl terephthalate.225 The thermal and mechanical properties of the resulting random copolyesters could be ne-tuned
to achieve the desired balance of material rigidity, ductility,
melting point and bio-based content.

5.

Renewable miscellaneous polyesters

There is clearly great interest in developing novel biomaterials


as emphasised by the number of publications on bio-based
polyesters over the years. The purpose of this section is not to
cover exhaustively the numerous scattered mentions to polyesters besides the aliphatic and aromatic ones, but rather to

Polym. Chem.

Scheme 23 Synthesis of poly(ester-urethane)s containing pendant photoresponsive moieties.228

choose a couple of them which produced interesting materials


such as the potentially biodegradable polymers with both ester
and other heteroatom-containing linkages in the backbone
structure, and the green polyester blends and sustainable
composite materials.
Among the polymers containing ester linkages as key sites
for biodegradation together with other heteroatom-containing
linkages, namely poly(ester-urethane)s,226228 poly(esteramide)s229231 and poly(ester-imide)s232234 are the most interesting ones. Representative examples of the ongoing research in
this eld include the work of Lluch et al.227 on bio-based multiblock polymers containing both ester and urethane linkages.
They reported the synthesis of allyl ester of undecenoic acid and
its subsequent polymerization via UV-initiated thiolene addition reactions to obtain a series of well-dened telechelics with
hydroxyl, carboxyl, or trimethoxysilyl end-groups. The telechelics containing hydroxy end-groups were then reacted with a
diisocyanate, yielding bio-based multi-block poly(esterurethane)s. Rochette and Ashby228 investigated poly(esterurethane)s containing pendant photoresponsive moieties for
tailorable shape memory bio-based materials (Scheme 23).
These biodegradable poly(ester urethane) thermosets with a
range of tunable thermal and mechanical properties possess
the ability to undergo light-induced shape memory which
would be advantageous for various applications including
implants and biomedical devices.
Poly(ester-amide)s from nontoxic natural gallic acid and
amino acids were reported by Li and co-workers.230 According to
their work, hyperbranched poly(ester-amide)s bearing terminal
acetyl groups were prepared via the self-polycondensation of a
series of AB3 monomers derived from gallic acid, glycine,
alanine and aminobutyric acid, originating poly(ester-amide)s
that are hydrolytically and enzymatically degradable. Karimi
et al.231 synthesized poly(ester-amide)s derived from the amino
acids phenylalanine and methionine by interfacial polymerization. The resulting polymers, susceptible to either hydrolytic or
enzymatic degradation and with excellent lm-forming properties, are potential candidates for tissue engineering and drug
delivery applications.

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Review

Scheme 24 Non-linear
derivatives.234

Polymer Chemistry

poly(ester-imide)s

from

castor

oil

and

furan

Vilela et al.232234 reported an innovative approach in which


thermally labile polymers containing both ester and imide linkages were obtained via the DielsAlder (DA) reaction (Scheme 24).
Linear232,233 and non-linear234 thermoreversible polymers based
on two renewable resources, i.e. castor oil and furan derivatives,
were prepared. Two click chemistry approaches were applied
in these macromolecular syntheses, namely a thiolene
coupling to convert unsaturated plant oil derivatives into
bifunctional and trifunctional monomers, and their DA polycondensation to obtain, respectively, linear and non-linear
thermoreversible materials. The retro-DA depolymerization of
these polymers at 110  C conrmed the clean-cut return to their
respective starting monomers. This distinctive feature opens the
way to original macromolecular materials with interesting
applications, such as self-mendability and recyclability.
Mallakpour et al.235,236 also investigated the preparation of
polymers containing both ester and imide linkages but now
derived from amino acid monomers. First, the authors
produced several poly(ester-imide)s via direct polyesterication
of a biodegradable optically active phenolic diol and diacids
containing dierent amino acids and the phthalimide group in
the side chain.235 These polymers are interesting candidates for
applications such as chiral materials for asymmetric synthesis
and chiral stationary phases for high-performance liquid
chromatography (HPLC). In another study, the same research
group conrmed that, as expected, their previously synthesised
poly(ester-imide)s are indeed biodegradable.236
Like several petroleum-based polyesters, many properties of
renewable-based polyesters can be enhanced through blending
and composite formation, leading to novel materials with
broader application elds. There have been many reports on
blends and composites of bio-based polyesters with renewablebased materials,237247 however, only the latest relevant contributions are highlighted here.
Konwar et al.240 reported the preparation of blends of highly
branched polyester and epoxy resins from Mesua ferrea L. seed
oil, with enhanced physical, thermal, and mechanical properties. These blends were then used as matrices for the preparation of nanocomposites of organophilic montmorillonite
nanoclay, with potential to be used as multipurpose advanced
thin-lm materials.240 Nurkhamidah et al.241 investigated the
phase behaviour and crystal morphology of blends of poly(ethylene succinate) and tannins, which are naturally biocompatible and biodegradable polyphenols. In a dierent vein,
Cock et al.242 prepared blends of two commercial biodegradable
polyesters, namely PLA (Ingeo produced by NatureWorks LLC)
and PCL (Capa manufactured by Perstorp), and showed that

This journal is The Royal Society of Chemistry 2013

the material's thermal/thermo-rheological properties and, thus,


its microstructure, are deeply dependent on its thermal history.
Regarding composite materials, Wu et al.243 explored the
biodegradability, morphology, mechanical, and thermal properties of biocomposites of PBT and sisal bres. Their results
pointed out that the biocomposites exhibited superior
mechanical properties and lower melting points, and the degree
of biodegradation increased with increasing bre content.
Furthermore, Wu et al.244246 also investigated the biodegradability, morphology, and mechanical properties of other
composite materials consisting of polyesters and agricultural
residues namely rice husk,244 sesame husk,245 and wheat bran.246
A dierent study is reported by Rahman et al.247 and it concerns
the addition of lignin in the preparation of exible biocomposites based on plasticized PLLA for possible applications
in packaging with improved shelf-life.

6.

Sustainable routes to polyesters

One of the major challenges dealt with in polymer syntheses is


indeed to use renewable resources, but also concomitantly to
approach convenient synthetic pathways avoiding hazard and
toxic substances and using more eco-friendly conditions.248
Thus, sustainable catalytic approaches in polymer chemistry
became an emerging research area with a tremendous impact
on environmental and even on economic issues.249 This section
draws the attention towards the possibility of synthesizing
polyesters via sustainable routes, namely through the use of
green catalytic systems, mild polymerization conditions and/or
atom ecient approaches.
One example of a slow turnover in polyester chemistry is the
progressive replacement of the highly ecient (poly)transesterication antimony catalysts, conventionally used in PET
production,250 by other less toxic catalysts, such as the titaniumbased Lewis acid catalysts, among other metal ones.250 For
example, very recently several furanic-based polyesters, already
mentioned before, were synthesised by transesterication followed by polytransesterication using titanium(IV) tetraisopropoxide, titanium(IV) butoxide, tin(II) 2-ethyl hexanoate, or
zirconium(IV) butoxide (Table 1).85,203207
However, the use of these less aggressive catalysts in polyester synthesis is still hampered by the need for using very high
temperatures and reduced pressure to achieve high conversion
yields and molecular weights. One approach which could
partially circumvent these hard conditions both applied to
aromatic and aliphatic polyester syntheses is ROP,251 which
does not imply the vacuum removal of by-products (e.g. water,
methanol, etc), on the contrary, the reaction occurs at atmospheric pressure. This has been tested for the synthesis of high
molecular weight PET251 and is routinely used in the production
of PLA as patented by Cargill (US) in 1992.104,252 In a typical
experiment, Nagahata et al.253 reported the synthesis of PET
using a macrocyclic ethylene terephthalate dimer and dierent
transesterication catalysts, conducted at atmospheric pressure
but using relatively high temperatures. Among the various
catalysts, 1,3-dichloro-1,1,3,3-tetrabutyldistannoxane promoted
the highest catalytic activity, and a colourless polymer with a

Polym. Chem.

View Article Online

Polymer Chemistry

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Table 1

Review

Titanium- and tin-based catalysts applied to the synthesis of novel renewable polyesters of FDCA

Polyester

Catalyst

T a/ C

t b/h

Mwc/kDa

Ref.d

Poly(ethylene 2,5-furandicarboxylate)
Poly(trimethylene 2,5furandicarboxylate)
Poly(butylene 2,5-furandicarboxylate)

Titanium(IV) butoxide
Titanium(IV) butoxide

210245
210245

>8
>8

250
90

203
203

Titanium(IV) butoxide
Titanium(IV) butoxide
Titanium(IV) tetraisopropoxide
Titanium(IV) butoxide
Tin(II) ethylhexanoate
Titanium(IV) butoxide
Titanium(IV) butoxide
Titanium(IV) butoxide

210245
160200
150200
180220
180220
210245
210245
190250

>2
8
26
28
28
>3
>5
9

42
26e
965
3.6
26
67
48
49

203
206
85
85
203
203
207

Titanium(IV) butoxide

160200

204

Poly(hexylene 2,5-furandicarboxylate)
Poly(octylene 2,5-furandicarboxylate)
Poly(butylene succinate-co-butylene 2,5furandicarboxylate)
Poly(ethylene 2,5-furandicarboxylate-cobutylene 2,5-furandicarboxylate)
a
d

Reaction temperature. b Esterication and polytransesterication reaction time.


Reference. e Mw determined by 1H NMR in triuoroacetic acid-d.

weight average molecular weight of 36 kDa was obtained in


100% isolation yield, by heating only for 3 min, at 200  C. The
ROP approach although representing a noteworthy advance in
polyester synthesis, especially in what concerns atom eciency,
is still limited. The relatively high temperatures used in some
cases and the necessity of preparing the oligomeric macrocyclic
precursors could represent the main disadvantages.
Another interesting approach, applied in this case to the
polycondensation (or just condensation) of aliphatic monomers, is conducted in water in the presence of a Brnsted acid
surfactant catalyst, in which the (poly)condensation is attained
at the interface of the emulsion.145,147,254256 This approach is
particularly interesting because of the low reaction temperature
usually adopted, avoiding undesirable side reactions. Also, the
use of aqueous media is quite attractive, since it is obviously a
safe, environmentally benign and very cheap solvent. One
example147 is the polycondensation of renewable-based longchain aliphatic suberin monomers. The reaction was conducted
in water at 80  C, for dierent reaction times, and using
p-dodecylbenzenesulfonic acid as a surfactant/catalyst. The
ensuing polyesters were, however, isolated in moderate yields
(between 6 and 49%).
Metal triuoromethanesulfonates (Lewis acid catalysts),
usually designated as triates, have reached a very important
place in the arsenal of proposed polycondensation catalysts.127,146,257259 Thus, bismuth(III) triate-mediated polycondensation reactions of aliphatic monomers have attracted some
interest throughout scientic communities due to their low
toxicity, ease of handling, low cost, and stability.127,146,257,258
Takasu et al.259 studied the chemoselective direct polycondensation of dierent dicarboxylic acids and alcohols
including the renewable monomers SA, glycerol, and sorbitol
and explored a one-step procedure involving the selective
reaction of primary hydroxyl groups. The polycondensation was
conducted under mild temperatures and using scandium triate as a catalyst. The reaction proceeded to aord linear
polyesters with pendant hydroxyl groups in their backbone in
excellent isolation yields (99%), and average molecular weights

Polym. Chem.

7.5

Weight average molecular weight determined by SEC.

between 4 and 26 kDa. More recently, Kricheldorf et al.257,258,260


published several articles on the use of metal triate catalysts in
the synthesis of polyesters, also by direct polycondensation. The
authors tested several triate catalysts including sodium,
magnesium, aluminium, zinc, tin, scandium, lanthanum,
samarium, yttrium, and hafnium, and used dierent aliphatic
monomers, namely C3C10 diols and dicarboxylic acids (see e.g.
in Scheme 25 one of the systems studied258). The reactions were
conducted in the bulk, at a moderate temperature (80  C), and
under reduced pressure. It was found that bismuth(III) triate
was one of the most convenient in terms of the extent of reactions (around 95%), average molecular weights of the ensuing
polymers (up to 35 kDa) and also because Bi(III) is the less toxic
among the heavy metals.
An inspiring strategy to polyester synthesis is denitely
enzymatic polymerization, which is a highly selective reaction,
occurs under mild and neutral conditions, at low temperature
and in a quantitative conversion.249 Hence, they sound the ideal
conditions for any chemical reaction and are perfectly in tune
with the green chemistry concept.
Several authors throughout the world studied the mild
enzymatic polycondensation or polytransesterication catalysed by lipases.114,128,147,150,153,249,261266 Lipases are hydrolases
that catalyse in vivo the hydrolysis of lipids into fatty acids and
glycerol in an aqueous emulsion environment (at the oilwater
interface).267 The pertinent discovery of lipases which catalyse
esterication reactions in vitro at relatively high temperatures
and in organic media made biocatalysis a valuable tool in

Scheme 25 Polycondensation reaction between 1,6-hexanediol and decanedioic acid catalysed by metal triates.258

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Review
polymer chemistry.267 Lipases can be isolated from a variety of
sources, for example, porcine pancreas, and many microorganisms. In particular, the commercially available Candida
antarctica lipase B (CALB) is denitely the most used in polyesterications. It is produced by a genetically modied Aspergillus oryzae microorganism and immobilised on poly(methyl
methacrylate-co-butyl methacrylate) resin, supplied by Bayer.
This enzyme is active for the polycondensation (or polytransesterication) and for the ROP, both in the bulk or organic
media, leading to the formation of polyesters in reasonably high
yields.249,265,268,269
One of the major advantages of CALB-catalysed reactions is
their selectivity towards primary hydroxyl groups (reactions
through secondary hydroxyl groups proceed slowly), which
enables one-pot synthesis of polyesters of diols with extra
secondary OH groups without tedious protection/deprotection
steps of the latter.269 Indeed, the conversion of multifunctional
monomers to linear or nearly linear homo- and copolymers,
together with linear unsaturated and linear epoxidized polyesters were reported.249
Kulshrestha et al.270 reported the bulk polycondensation of
adipic acid, 1,8-octanediol, and glycerol (Scheme 26), at 70  C,
under vacuum, using CALB as a catalyst. The ensuing products
were obtained in average molecular weights ranging from 27 to
81 kDa and polydispersity values between 2 and 8, depending
on the adopted reaction conditions. Furthermore, CALB's
regioselectivity converted the multifunctional monomers to
linear copolymers for polymerizations up to 18 h. However, as
the reaction time was extended to 42 h, products formed
became increasingly branched. Thus, at 42 h, a hyperbranched
polymer, resulting from the reaction of the secondary OH
group, was obtained in 90% yield with 19 mol % dendritic
glycerol repeating units, average molecular weight and polydispersity values of 76 kDa and 3.1, respectively.
Olsson et al.,150 as already mentioned above, studied the
polycondensation of 18-hydroxy-9,10-epoxyoctadecanoic acid
(Scheme 6) from birch outer bark suberin using the CALB
lipase. These reactions were carried out either in the bulk or in
the presence of an organic solvent. Interestingly, bulk

Scheme 26 Synthesis of linear and hyperbranched glycerol copolyesters using


CALB as a catalyst.270

This journal is The Royal Society of Chemistry 2013

Polymer Chemistry
polycondensations gave molecular weights comparable to those
obtained with toluene (15 and 20 kDa, respectively), at a much
shorter reaction time. The reactions were conducted at
moderate temperatures (75 or 85  C), but during a reasonable
period of time (48 h), and using a relatively high amount of
lipase (25% w/w). The most important aspect of this study was
denitely the success in synthesizing the linear epoxidized
poly(cis-9,10-epoxy-18-hydroxyoctadecanoate).
The versatility of CALB-catalysed synthesis extends to the use
of aromatic diacid compounds. For example, Linko et al.271
studied the polymerizations of isophthalic acid with BDO that
yielded oligomers, and a similar reaction between the 1,6-hexanediol and isophthalic acid at 70  C yielded a polymer with an
average molecular weight of 55 kDa.
Although these enzymatic-catalysed reactions proceed under
sustainable pathways, they are not yet considered economically
competitive,272 essentially due to the high costs of the enzyme
and the long reaction times. Hence, even here, in such a
promising polyester strategy, industry and scientists have to
gain balance to nd cheap and viable processes to enable
enzyme use at a large scale in polyester synthesis.
Other promising approaches that have gained interest
recently comprise the use of ionic liquids as reaction/catalyst
media,273 or even the use of click-chemistry274 in polyester
synthesis. Nevertheless, these approaches are still limited in
some cases by the use of non-ecofriendly conditions.

7.

Conclusions and insights into the future

In recent years much has been written about renewable


resources and their use in polymer science.6 The raison d'etre for
this interest is intimately related to the nite supply of petroleum
and, regardless of the time frame of the end of cheap oil, also
because society is genuinely concerned with the environmental
impacts associated with the augment of fossil CO2 emissions.6
Several pertinent questions regarding renewable resources
and polyesters from renewable resources have been raised over
the years: Which chemicals and polymeric materials (polyesters included) can be produced at competitive costs from
renewable resources? Are polyesters from renewable resources
valuable alternatives to the petrochemical counterparts?. This
state of aairs has been scrutinized by several authors
including, for example, Mecking in the review about Nature or
petrochemistry? biologically degradable materials;275 Mathers in the highlight under the title How well can renewable
resources mimic commodity monomers and polymers?;3 and
Miller in the viewpoint on Sustainable polymers: opportunities
for the next decade.190
The future challenges that must be addressed regarding
renewable resource based production of polyesters, and bioplastics in general, are (i) the sustainable management of raw
material resources, (ii) the sustainable production of chemicals
and polymeric materials from biomass, (iii) the performance of
the biobased polyesters and, ultimately, (iv) their production
costs.32 These are indeed the major challenges for the sustainable
development of society based on the future bioreneries. The
major impediments to the growth of sustainable polyesters are,

Polym. Chem.

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Polymer Chemistry
on the one hand, the petroleum industry lobbies given the wellestablished petroleum-based polyesters, and on the other hand,
the competition of raw materials with food and feed supply.276
The former is giving bioplastics a hard competition, and unless
local and national governments pass legislation encompassing
economic/tax benets that stimulate the integration of a growing
share of bioplastics in the industrial supply (similar to what has
been done in several countries to stimulate biobased transportation fuel production and usage), the transition will be slow
and dicult.276 The latter can be relatively simple to overcome by
using non-food sources such as cellulose, lignin and vegetable
oils as renewable resources through the concept of biorenery,3
bearing in mind that adequate management of soil resources and
biodiversity protection should imperatively be taken into account
for the whole chain to be sustainable in the long-term.
Although there is still a long way to go in terms of R&D in
these elds, the expansion of renewable-based polyesters is also
hindered by economical issues related to cost-competitiveness.
Having in consideration that currently renewable-based chemicals can hardly compete with petroleum counterparts, two
approaches seem logical in the short-medium term to stimulate
its growth: (i) to make bio-versions of the existing fossil-based
polyesters, e.g. DuPont is commercializing PTT, trade name
Sorona EP, based on renewable 1,3-propanediol and petrochemical TPA,26 and (ii) to progressively replace petroleumbased homologues with their renewable-based homologues, e.g.
Avantium, which produces FDCA, has announced the development of the rst PEF bottle.25
In this regard, the use of biomass in the chemical industry is
already a feasible reality, although on a small scale. Nevertheless, the continuous skyrocketing consumer demand for
advanced plastics will denitely contribute to the steady
conversion from petrochemical-based polyesters to sustainable
and renewable-based polyesters.
So globally, the development of sustainable polyesters from
renewable resources, as part of the general concern for
sustainability, will denitely continue to ourish since they are
serious candidates to replace oil-based polymers such as polyethylene, polypropylene, PET, among others. Notwithstanding
the promising aspect of investigations related to polyesters
from renewable resources, it is reasonable to assume that great
challenges remain and there are plenty of possibilities in the
future for innovation and environmentally friendlier polyesters.

Acknowledgements
The authors wish to thank FCT (Funda~ao para a Ci
encia e
Tecnologia) and POPH/FSE for the postdoctoral grants to CV
(SFRH/BPD/84168/2012) and AFS (SFRH/BPD/73383/2010),
and for Associate Laboratory CICECO funding (PEst-C/CTM/
LA0011/2013).

References
1 M. N. Belgacem and A. Gandini, Monomers, Polymers and
Composites from renewable resources, Elsevier, Amsterdam,
1st edn, 2008.

Polym. Chem.

Review
2 A. Gandini, Green Chem., 2011, 13, 1061.
3 R. T. Mathers, J. Polym. Sci., Part A: Polym. Chem., 2012, 50, 1.
4 P. A. Wilbon, F. Chu and C. Tang, Macromol. Rapid
Commun., 2013, 34, 8.
5 Towards Sustainable Development, in Our Common Future,
ch. 2, http://www.un-documents.net/ocf-02.htm, accessed
August 2013.
6 A. J. Ragauskas, C. K. Williams, B. H. Davison, G. Britovsek,
J. Cairney, C. A. Eckert, W. J. Frederick Jr, J. P. Hallett,
D. J. Leak, C. L. Liotta, J. R. Mielenz, R. Murphy,
R. Templer and T. Tschaplinski, Science, 2006, 311, 484.
7 D. R. Dodds and R. A. Gross, Science, 2007, 318, 1250.
8 B. P. Mooney, Biochem. J., 2009, 418, 219.
9 C. S. Marvel, J. Chem. Educ., 1981, 58, 535.
10 A. Gandini, in Biocatalysis in Polymer Chemistry, ed. K. Loos,
Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany,
2011, pp. 133.
11 European Bioplastics, Bioplastics: facts and gures, http://
en.european-bioplastics.org/, accessed August 2013.
12 J. M. G. Cowie and V. Arrighi, Polymers: chemistry & physics
of modern materials, CRC Press, Boca Raton, 3rd edn, 2008.
13 U. Edlund and A.-C. Albertsson, Adv. Drug Delivery Rev.,
2003, 55, 585.
14 J. Scheirs and T. E. Long, Modern Polyesters: Chemistry and
Technology of Polyesters and Copolyesters, John Wiley &
Sons, Ltd, West Sussex, England, 2003.
15 L. Shen, J. Haufe, and M. K. Patel, PRO-BIP, 2009, http://
www.plastice.org/leadmin/les/PROBIP2009_Final_June_
2009.pdf, accessed August 2013.
16 C. Williams and M. Hillmyer, Polym. Rev., 2008, 48, 1.
17 National Renewable Energy Laboratory (NREL), What Is a
Biorenery?, http://www.nrel.gov/biomass/biorenery.html,
accessed August 2013.
18 H. R. Ghatak, Renewable Sustainable Energy Rev., 2011, 15,
4042.
19 V. Menon and M. Rao, Prog. Energy Combust. Sci., 2012, 38,
522.
20 B. Kamm, Angew. Chem., Int. Ed., 2007, 46, 5056.
21 J. J. Bozell and G. R. Petersen, Green Chem., 2010, 12, 539.
22 NatureWorks LLC, The Ingeo Journey, http://www.nature
worksllc.com//media/News_and_Events/NatureWorks_
TheIngeoJourney_pdf.pdf, accessed August 2013.
23 CornFlower, PLA Tableware, http://www.cornower.com.tw/
English.htm, accessed August 2013.
24 Showa Denko, Bionolle Biodegradable aliphatic polyester,
http://www.showa-denko.com/leadmin/template/img/
Bionolle/Bionelle_2013_reduced_size_PDF.pdf, accessed
August 2013.
25 Avantium,
FDCA,
http://avantium.com/yxy/productsapplications/fdca.html, accessed August 2013.
26 DuPont Performance Polymers, http://www2.dupont.com/
Plastics/en_US/assets/downloads/uses_apps/DPM_Renewa
bly_Sourced_5pBrochure_RSE_A10904_00_A1010_PDF.pdf,
accessed August 2013.
27 The Coca-Cola Company, PlantBottle Packaging, http://
www.coca-cola.com/content-store/en_US/SC/PlantBottle/,
accessed August 2013.

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Review
28 M. Okada, Prog. Polym. Sci., 2002, 27, 87.
29 J. N. Hoskins and S. M. Grayson, Polym. Chem., 2011, 2, 289.
30 D. J. A. Cameron and M. P. Shaver, Chem. Soc. Rev., 2011, 40,
1761.
31 J. Twibanire and T. B. Grindley, Polymers, 2012, 4, 794.
32 R. P. Babu, K. O'Connor and R. Seeram, Prog. Biomat., 2013,
2, 8.
33 A. Tsui, Z. C. Wright and C. W. Frank, Annu. Rev. Chem.
Biomol. Eng., 2013, 4, 143.
34 G.-Q. Chen and M. K. Patel, Chem. Rev., 2012, 112, 2082.
35 K. Yao and C. Tang, Macromolecules, 2013, 46, 1689.
36 A.-L. Marshall and P. J. Alaimo, Chem.Eur. J., 2010, 16, 4970.
37 H. Kobayashi and A. Fukuoka, Green Chem., 2013, 15,
1740.
38 T. Werpy and G. Petersen, Top Value Added Chemicals from
Biomass Volume IResults of Screening for Potential
Candidates from Sugars and Synthesis Gas, U.S.
Department of Energy, Washington, DC, 2004.
39 F. Fenouillot, A. Rousseau, G. Colomines, R. Saint-Loup and
J.-P. Pascault, Prog. Polym. Sci., 2010, 35, 578.
40 J. A. Galbis and M. G. Garca-Martn, in Monomers, Polymers
and Composites from Renewable Resources, ed.
M. N. Belgacem and A. Gandini, Elsevier Ltd, Amsterdam,
2008, pp. 89114.
41 H. Song and S. Y. Lee, Enzyme Microb. Technol., 2006, 39, 352.
42 K.-K. Cheng, X.-B. Zhao, J. Zeng and J.-A. Zhang, Biofuels,
Bioprod. Bioren., 2012, 6, 302.
43 T. Kurzrock and D. Weuster-Botz, Biotechnol. Lett., 2010, 32,
331.
44 J. J. Beauprez, M. De Mey and W. K. Soetaert, Process
Biochem., 2010, 45, 1103.
45 J. McKinlay, C. Vieille and J. G. Zeikus, Appl. Microbiol.
Biotechnol., 2007, 76, 727.
46 J. Xu and B.-H. Guo, Biotechnol. J., 2010, 5, 1149.
47 P. Harmsen and M. Hackmann, Green building blocks for
biobased
plastics,
http://www.groenegrondstoen.nl/
downloads/Boekjes/16GreenBuildingblocks.pdf, accessed
August 2013.
48 I. Goldberg, J. S. Rokem and O. Pines, J. Chem. Technol.
Biotechnol., 2006, 81, 1601.
49 G. T. Tsao, N. J. Cao, J. Du and C. S. Gong, in Recent Progress
in Bioconversion of Lignocellulosics, ed. G. T. Tsao, Springer,
Berlin Heidelberg, 1999, vol. 65, pp. 243280.
50 C. Engel, A. Straathof, T. Zijlmans, W. van Gulik and L. van
der Wielen, Appl. Microbiol. Biotechnol., 2008, 78, 379.
51 Q. Xu, S. Li, H. Huang and J. Wen, Biotechnol. Adv., 2012, 30,
1685.
52 G. Xu, W. Zou, N. Xu, L. Liu and J. Chen, PLoS One, 2012, 7,
e52086.
53 E. Kaclkov
a, T. Lachowicz, Y. Gbelsk
a and J. Subk, FEMS
Microbiol. Lett., 1992, 91, 101.
54 M. Okabe, D. Lies, S. Kanamasa and E. Y. Park, Appl.
Microbiol. Biotechnol., 2009, 84, 597.
55 A. Kuenz, Y. Gallenmuller, T. Wilke and K.-D. Vorlop, Appl.
Microbiol. Biotechnol., 2012, 96, 1209.
56 D. Barrett, T. Merkel, J. Lu and M. Yousaf, Macromolecules,
2010, 43, 9660.

This journal is The Royal Society of Chemistry 2013

Polymer Chemistry
57 B. Guo, Y. Chen, Y. Lei, L. Zhang, W. Y. Zhou, A. B. M. Rabie
and J. Zhao, Biomacromolecules, 2011, 12, 1312.
58 O. Goerz and H. Ritter, Polym. Int., 2013, 62, 709.
59 M. G. Steiger, M. L. Blumho, D. Mattanovich and M. Sauer,
Front. Microbiol., 2013, 4, 23.
60 G. Xie and T. P. West, Lett. Appl. Microbiol., 2009, 48, 639.
61 E. Alben and O. Erkmen, Food Technol. Biotechnol., 2004,
42, 19.
62 A. R. Angumeenal and D. Venkappayya, LWTFood Sci.
Technol., 2013, 50, 367.
63 X. Feng, A. East, W. Hammond and M. Jae, in
Contemporary Science of Polymeric Materials, ed.
L.
Korugic-Karasz,
American
Chemical
Society,
Washington, DC, 2010, pp. 327.
64 I. S. Risti
c, N. Vukic, S. Caki
c, V. Simendic, O. Risti
c and
J. Budinski-Simendc, J. Polym. Environ., 2012, 20, 519.
65 L. Jasinska and C. E. Koning, J. Polym. Sci., Part A: Polym.
Chem., 2010, 48, 2885.
66 B. A. J. Noordover, V. G. van Staalduinen, R. Duchateau,
C. E. Koning, R. A. T. M. van Benthem, M. Mak, A. Heise,
A. E. Frissen and J. van Haveren, Biomacromolecules, 2006,
7, 3406.
67 G. Kaur, A. K. Srivastava and S. Chand, Biochem. Eng. J.,
2012, 64, 106.
68 H. Liu, Y. Xu, Z. Zheng and D. Liu, Biotechnol. J., 2010, 5,
1137.
69 D. Szymanowska-Powalowska, A. Drozdzynska and
N. Remszel, Advances in Microbiology, 2013, 3, 171.
70 R. K. Saxena, P. Anand, S. Saran and J. Isar, Biotechnol. Adv.,
2009, 27, 895.
71 S. K. Hoekman, A. Broch, C. Robbins, E. Ceniceros and
M. Natarajan, Renewable Sustainable Energy Rev., 2012, 16,
143.
72 C. E. Nakamura and G. M. Whited, Curr. Opin. Biotechnol.,
2003, 14, 454.
73 Z. Ma, X. Shentu, Y. Bian and X. Yu, Eng. Life Sci., 2012, 12,
553.
74 H. Yim, R. Haselbeck, W. Niu, C. Pujol-Baxley, A. Burgard,
J. Boldt, J. Khandurina, J. D. Trawick, R. E. Osterhout,
R. Stephen, J. Estadilla, S. Teisan, H. B. Schreyer,
S. Andrae, T. H. Yang, S. Y. Lee, M. J. Burk and S. Van
Dien, Nat. Chem. Biol., 2011, 7, 445.
75 V. Sharma and P. P. Kundu, Prog. Polym. Sci., 2006, 31, 983.
76 Y. Lu and R. C. Larock, ChemSusChem, 2009, 2, 136.
77 M. N. Belgacem and A. Gandini, in Monomers, Polymers and
Composites from Renewable Resources, ed. N. M. Belgacem
and A. Gandini, Elsevier Ltd, Amsterdam, 2008, pp. 3966.
78 Elevance Renewable Sciences, http://www.elevance.com/
platforms/engineered-polymers-and-coatings/inherent-c18diacid/, accessed September 2013.
79 U. Biermann, W. Friedt, S. Lang, W. L
uhs, G. Machm
uller,
J. O. Metzger, M. R
usch gen. Klaas, H. J. Sch
afer and
M. P. Schneider, Angew. Chem., Int. Ed., 2000, 39, 2206.
80 U. Biermann, U. Bornscheuer, M. A. R. Meier, J. O. Metzger
and H. J. Sch
afer, Angew. Chem., Int. Ed., 2011, 50, 3854.
81 A. Gandini, C. Pascoal Neto and A. J. D. Silvestre, Prog.
Polym. Sci., 2006, 31, 878.

Polym. Chem.

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Polymer Chemistry
82 A. Heredia, Biochim. Biophys. Acta, 2003, 1620, 1.
83 J. E. Holladay, J. F. White, J. J. Bozell and D. Johnson, Top
Value-Added Chemicals from Biomass Volume IIResults of
Screening for Potential Candidates from Biorenery Lignin,
U.S. Department of Energy, Washington, DC, 2007, vol. II.
84 A. F. Sousa, M. Matos, C. S. R. Freire, A. J. D. Silvestre and
J. F. J. Coelho, Polymer, 2013, 54, 513.
85 E. Gubbels, L. Jasinska-Walc and C. E. Koning, J. Polym. Sci.,
Part A: Polym. Chem., 2013, 51, 890.
86 X. Tong, Y. Ma and Y. Li, Appl. Catal., A, 2010, 385, 1.
87 M. Bicker, J. Hirth and H. Vogel, Green Chem., 2003, 5, 280.
88 C. Triebl, V. Nikolakis and M. Ierapetritou, Comput. Chem.
Eng., 2013, 52, 26.
89 M. J. Climent, A. Corma and S. Iborra, Green Chem., 2011,
13, 520.
90 A. A. Rosatella, S. P. Simeonov, R. F. M. Frade and
C. A. M. Afonso, Green Chem., 2011, 13, 754.
91 F. Koopman, N. Wierckx, J. H. de Winde and
H. J. Ruijssenaars, Bioresour. Technol., 2010, 101, 6291.
92 G. Gellerstedt and G. Henriksson, in Monomers, Polymers
and Composites from Renewable Resources, ed. M. N.
Belgacem and A. Gandini, Elsevier Ltd., Amsterdam, 2008,
pp. 201224.
93 D. Havkin-Frenkel and F. C. Belanger, in Biotechnology
in Flavor Production, ed. D. Havkin-Frenkel and F. C.
Belanger, Blackwell Publishing Ltd., Oxford, UK, 2009,
pp. 83103.
94 T. Voitl and V. Rohr, Ind. Eng. Chem. Res., 2010, 49, 520.
95 J. D. P. Ara
ujo, C. A. Grande and A. E. Rodrigues, Chem. Eng.
Res. Des., 2010, 88, 1024.
96 P. C. R. O. Pinto, C. E. Costa and A. E. Rodrigues, Ind. Eng.
Chem. Res., 2013, 52, 4421.
97 H. Priefert, J. Rabenhorst and A. Steinb
uchel, Appl.
Microbiol. Biotechnol., 2001, 56, 296.
98 F. Perestelo, M. A. Dalcon and G. de la Fuente, Appl. Environ.
Microbiol., 1989, 55, 1660.
99 Y. Sazanov, M. Goykhman, I. Podeshvo, G. Fedorova,
G. Mikhailov and V. Kudriavtsev, J. Therm. Anal. Calorim.,
1999, 55, 721.
100 L. Mialon, R. Vanderhenst, A. G. Pemba and S. A. Miller,
Macromol. Rapid Commun., 2011, 32, 1386.
101 S. Ghosh, A. Sachan and A. Mitra, Curr. Sci., 2006, 90, 825.
102 C. Civolani, P. Barghini, A. R. Roncetti, M. Ruzzi and
A. Schiesser, Appl. Environ. Microbiol., 2000, 66, 2311.
103 M. Jamshidian, E. A. Tehrany, M. Imran, M. Jacquot and
S. Desobry, Compr. Rev. Food Sci. FoodSaf., 2010, 9, 552.
104 L. Av
erous, in Monomers, polymers and composites from
renewable resources, ed. M. N. Belgacem and A. Gandini,
Elsevier Ltd, Amsterdam, 2008, pp. 433450.
105 R. Mehta, V. Kumar, H. Bhunia and S. N. Upadhyay,
J. Macromol. Sci., Part A: Pure Appl.Chem., 2005, 45, 325.
106 R. M. Rasal, A. V. Janorkar and D. E. Hirt, Prog. Polym. Sci.,
2010, 35, 338.
107 M. N. Somleva, O. P. Peoples and K. D. Snell, Plant
Biotechnol. J., 2013, 11, 233.
108 Bio-on, MINERV-PHA, http://bio-on.it/what.php, accessed
August 2013.

Polym. Chem.

Review
109 G.-Q. Chen, Chem. Soc. Rev., 2009, 38, 2434.
110 M. Kolitz, N. Cohen-Arazi, I. Hagag, J. Katzhendler and
A. J. Domb, Macromolecules, 2009, 42, 4520.
111 N. Cohen-Arazi, A. J. Domb and J. Katzhendler, Polymers,
2010, 2, 418.
112 J. R. Lowe, W. B. Tolman and M. A. Hillmyer,
Biomacromolecules, 2009, 10, 2003.
113 J. R. Lowe, M. T. Martello, W. B. Tolman and
M. A. Hillmyer, Polym. Chem., 2011, 2, 702.
114 H. Ebata, K. Toshima and S. Matsumura, Macromol. Biosci.,
2007, 7, 798.
115 H. Ebata, K. Toshima and S. Matsumura, J. Synth. Org.
Chem., Jpn., 2008, 66, 673.
116 Z. S. Petrovi
c, J. Milic, Y. Xu and I. Cvetkovic,
Macromolecules, 2010, 43, 4120.
117 R. Slivniak and A. J. Domb, Macromolecules, 2005, 38, 5545.
118 M. Sokolsky-Papkov and A. J. Domb, Polym. Adv. Technol.,
2008, 19, 671.
119 M. Sokolsky-Papkov, L. Golovanevski, A. J. Domb and
C. F. Weiniger, Pharm. Res., 2009, 26, 32.
120 J. G. Hiremath, V. K. Devi, K. Devi and A. J. Domb, J. Appl.
Polym. Sci., 2007, 107, 2745.
121 A. Shikanov, B. Vaisman, S. Shikanov and A. J. Domb,
J. Biomed. Mater. Res., Part A, 2010, 92, 1283.
122 D. Quinzler and S. Mecking, Angew. Chem., Int. Ed., 2010,
49, 4306.
123 J. Trzaskowski, D. Quinzler, C. B
ahrle and S. Mecking,
Macromol. Rapid Commun., 2011, 32, 1352.
124 F. Stempe, D. Quinzler, I. Heckler and S. Mecking,
Macromolecules, 2011, 44, 4159.
125 F. Stempe, P. Ortmann and S. Mecking, Macromol. Rapid
Commun., 2013, 34, 47.
126 D. Quinzler and S. Mecking, Chem. Commun., 2009, 5400.
127 C. Vilela, A. J. D. Silvestre and M. A. R. Meier, Macromol.
Chem. Phys., 2012, 213, 2220.
128 Y. Yang, W. Lu, X. Zhang, W. Xie, M. Cai and R. A. Gross,
Biomacromolecules, 2010, 11, 259.
129 Y.-R. Zhang, S. Spinella, W. Xie, J. Cai, Y. Yang, Y.-Z. Wang
and R. A. Gross, Eur. Polym. J., 2013, 49, 793.
130 P.-J. Roumanet, F. La`
eche, N. Jarroux, Y. Raoul, S. Claude
and P. Gu
egan, Eur. Polym. J., 2013, 49, 813.
131 O. T
ur
un and M. A. R. Meier, Eur. J. Lipid Sci. Technol.,
2013, 115, 41.
132 M. A. R. Meier, Macromol. Chem. Phys., 2009, 210, 1073.
133 C. O. Akintayo, H. Mutlu, M. Kempf, M. Wilhelm and
M. A. R. Meier, Macromol. Chem. Phys., 2012, 213, 87.
134 O. T
ur
un and M. A. R. Meier, Macromol. Rapid Commun.,
2010, 31, 1822.
135 A. Corma, S. Iborra and A. Velty, Chem. Rev., 2007, 107,
2411.
136 A. Behr, J. Eilting, K. Irawadi, J. Leschinski and F. Lindner,
Green Chem., 2008, 10, 13.
137 F. Yang, M. A. Hanna and R. Sun, Biotechnol. Biofuels, 2012,
5, 13.
138 R. Dobson, V. Gray and K. Rumbold, J. Ind. Microbiol.
Biotechnol., 2012, 39, 217.
139 Y. Gu and F. J
er
ome, Green Chem., 2010, 12, 1127.

This journal is The Royal Society of Chemistry 2013

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Review
140 S. Salehpour and M. A. Dub
e, J. Macromol. Sci., Part A: Pure
Appl.Chem., 2012, 49, 103.
141 S. Salehpour, C. J. Zuliani and M. A. Dub
e, Eur. J. Lipid Sci.
Technol., 2012, 114, 92.
142 S. J. Yoon, Y.-C. Choi, Y.-I. Son, S.-H. Lee and J.-G. Lee,
Bioresour. Technol., 2010, 101, 1227.
143 P. G. Parzuchowski, M. Grabowska, M. Jaroch and
M. Kusznerczuk, J. Polym. Sci., Part A: Polym. Chem., 2009,
47, 3860.
144 X. Zhao, L. Liu, H. Dai, C. Ma, X. Tan and R. Yu, J. Appl.
Polym. Sci., 2009, 113, 3376.
145 A. F. Sousa, A. J. D. Silvestre, A. Gandini and C. P. Neto, High
Perform. Polym., 2012, 24, 4.
146 A. F. Sousa, A. Gandini, A. J. D. Silvestre and C. P. Neto,
J. Polym. Sci., Part A: Polym. Chem., 2011, 49, 2281.
147 A. F. Sousa, A. Gandini, A. J. D. Silvestre and C. Pascoal
Neto, ChemSusChem, 2008, 1, 1020.
148 J. A. Heredia-Guerrero, M. A. San-Miguel, M. S. P. Sansom,
A. Heredia and J. J. Bentez, Langmuir, 2009, 25, 6869.
149 J. A. Heredia-Guerrero, A. Heredia, R. Garca-Segura and
J. J. Bentez, Polymer, 2009, 50, 5633.
150 A. Olsson, M. Lindstro and T. Iversen, Biomacromolecules,
2007, 8, 757.
151 P. C. R. O. Pinto, A. F. Sousa, A. J. D. Silvestre, C. P. Neto,
A. Gandini, C. Eckerman and B. Holmbom, Ind. Crops
Prod., 2009, 29, 126.
152 A. F. Sousa, A. Gandini, A. J. D. Silvestre and C. Pascoal
Neto, Appl. Spectrosc., 2009, 63, 873.
153 H. Ebata, K. Toshima and S. Matsumura, Macromol. Biosci.,
2008, 8, 38.
154 H. R. Kricheldorf, S. Chatti, G. Schwarz and R.-P. Kr
uger,
J. Polym. Sci., Part A: Polym. Chem., 2003, 41, 3414.
155 W. J. Yoon, K. S. Oh, J. M. Koo, J. R. Kim, K. J. Lee and
S. S. Im, Macromolecules, 2013, 46, 2930.
156 M. Garaleh, T. Yashiro, H. R. Kricheldorf, P. Simon and
S. Chatti, Macromol. Chem. Phys., 2010, 211, 1206.
157 J. Wu, P. Eduard, L. Jasinska-Walc, A. Rozanski,
B. A. J. Noordover, D. S. van Es and C. E. Koning,
Macromolecules, 2013, 46, 384.
158 C. Lavilla, A. Alla, A. M. de Ilarduya, E. Benito,
M. G. Garca-Martn, J. A. Galbis and S. Mu~
noz-Guerra,
Biomacromolecules, 2011, 12, 2642.
159 C. Lavilla, A. Alla, A. M. de Ilarduya, E. Benito, M. G. GarcaMartn, J. A. Galbis and S. Mu~
noz-Guerra, J. Polym. Sci., Part
A: Polym. Chem., 2012, 50, 1591.
160 C. Lavilla, A. Alla, A. Martnez de Ilarduya and S. Mu~
nozGuerra, Biomacromolecules, 2013, 14, 781.
161 M. Tang, A. J. P. White, M. M. Stevens and C. K. Williams,
Chem. Commun., 2009, 941.
162 S. Kakasi-Zsurka, A. Todea, A. But, C. Paul, C. G. Boeriu,
Kuki, S. K
C. Davidescu, L. Nagy, A.
eki and F. P
eter,
J. Mol. Catal. B: Enzym., 2011, 71, 22.
163 I. Bechthold, K. Bretz, S. Kabasci, R. Kopitzky and
A. Springer, Chem. Eng. Technol., 2008, 31, 647.
164 B. A. J. Noordover, A. Heise, P. Malanowksi, D. Senatore,
M. Mak, L. Molhoek, R. Duchateau, C. E. Koning and
R. A. T. M. van Benthem, Prog. Org. Coat., 2009, 65, 187.

This journal is The Royal Society of Chemistry 2013

Polymer Chemistry
165 R. Mincheva, A. Delangre, J.-M. Raquez, R. Narayan and
P. Dubois, Biomacromolecules, 2013, 14, 890.
166 T. Okuda, K. Ishimoto, H. Ohara and S. Kobayashi,
Macromolecules, 2012, 45, 4166.
167 R. Marn, A. Alla and S. Mu~
noz-Guerra, Macromol. Rapid
Commun., 2006, 27, 1955.
168 R. Marn, A. Martnez de Ilarduya, P. Romero, J. R. Sarasua,
E. Meaurio, E. Zuza and S. Mu~
noz-Guerra, Macromolecules,
2008, 41, 3734.
169 M. Yokoe, K. Aoi and M. Okada, J. Polym. Sci., Part A: Polym.
Chem., 2005, 43, 3909.
170 R. Wu, T. F. Al-Azemi and K. S. Bisht, Macromolecules, 2009,
42, 2401.
171 S. Dhamaniya and J. Jacob, Polymer, 2010, 51, 5392.
172 S. Dhamaniya and J. Jacob, Polym. Bull., 2012, 68, 1287.
173 S. Dhamaniya, D. Das, B. K. Satapathy and J. Jacob, Polymer,
2012, 53, 4662.
174 S. Dhamaniya, H. S. Jaggi, M. Nimiya, S. Sharma,
B. K. Satapathy and J. Jacob, Polym. Int., 2013, DOI:
10.1002/pi.4569.
175 L. J. DiBenedetto and S. J. Huang, Polym. Degrad. Stab.,
1994, 45, 249.
176 F. Barroso-Bujans, R. Martnez and P. Ortiz, J. Appl. Polym.
Sci., 2003, 88, 302.
177 N. Tsutsumi, M. Oya and W. Sakai, Macromolecules, 2004,
37, 5971.
178 J. Yang, A. R. Webb, S. J. Pickerill, G. Hageman and
G. A. Ameer, Biomaterials, 2006, 27, 1889.
179 L. V. Thomas, U. Arun, S. Remya and P. D. Nair, J. Mater.
Sci.: Mater. Med., 2009, 20, S259.
180 Y. Chandorkar, G. Madras and B. Basu, J. Mater. Chem. B,
2013, 1, 865.
181 N. Mithil Kumar, K. Varaprasad, K. Madhusudana Rao,
A. Suresh Babu, M. Srinivasulu and S. Venkata Naidu,
J. Polym. Environ., 2011, 20, 17.
182 B. A. J. Noordover, R. Duchateau, R. A. T. M. van Benthem,
W. Ming and C. E. Koning, Biomacromolecules, 2007, 8,
3860.
183 B. Tisserat, R. H. O'kuru, H. Hwang, A. A. Mohamed and
R. Holser, J. Appl. Polym. Sci., 2012, 125, 3429.
184 J. M. Halpern, R. Urbanski, A. K. Weinstock, D. F. Iwig,
R. T. Mathers and H. A. von Recum, J. Biomed. Mater.
Res., Part A, 2013, DOI: 10.1002/jbm.a.34821.
185 M. Hakkarainen and A. Albertsson, Adv. Polym. Sci., 2008,
211, 85.
186 BASF, Ecoex Biodegradable Plastic Overview, http://
www.bioplastics.basf.com/ecoex.html, accessed August
2013.
187 F. Trinh Tan, D. G. Cooper, M. Mari
c and J. A. Nicell, Polym.
Degrad. Stab., 2008, 93, 1479.
188 T. Kijchavengkul, R. Auras, M. Rubino, S. Selke,
M. Ngouajio and R. T. Fernandez, Polym. Degrad. Stab.,
2010, 95, 2641.
189 A. A. Shah, T. Eguchi, D. Mayumi, S. Kato, N. Shintani,
N. R. Kamini and T. Nakajima-Kambe, Polym. Degrad.
Stab., 2013, 98, 609.
190 S. A. Miller, ACS Macro Lett., 2013, 2, 550.

Polym. Chem.

View Article Online

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

Polymer Chemistry
191 A. Kaplan, Awaiting PlantBottle 2.0, http://www.beverage
world.com/articles/full/15300/awaiting-plantbottle-2.0,
accessed August 2013.
192 A. Gandini, Polym. Chem., 2010, 1, 245.
193 J. A. Moore and J. E. Kelly, Macromolecules, 1978, 11, 568.
194 J. A. Moore and J. E. Kelly, Polymer, 1979, 20, 627.
195 A. Khrouf, S. Bou, R. El Gharbi, N. M. Belgacem and
A. Gandini, Polym. Bull., 1996, 37, 589.
196 L. Gharbi, M. Abid, R. El Gharbi and A. Gandini, Macromol.
Chem. Phys., 1998, 199, 2755.
197 A. Khrouf, S. Bou, R. El Gharbi and A. Gandini, Polym. Int.,
1999, 48, 649.
198 S. Gharbi, J. Andreolety and A. Gandini, Eur. Polym. J., 2000,
36, 463.
199 M. Okada, K. Tachikawa and K. Aoi, J. Polym. Sci., Part A:
Polym. Chem., 1997, 35, 2729.
200 A. Gandini, A. J. D. Silvestre, C. P. Neto and A. F. Sousa,
J. Polym. Sci., Part A: Polym. Chem., 2009, 47, 295.
201 A. Gandini, D. Coelho, M. Gomes, B. Reis and A. Silvestre,
J. Mater. Chem., 2009, 19, 8656.
202 M. Gomes, A. Gandini, A. J. D. Silvestre and B. Reis,
J. Polym. Sci., Part A: Polym. Chem., 2011, 49, 3759.
203 M. Jiang, Q. Liu, Q. Zhang, C. Ye and G. Zhou, J. Polym. Sci.,
Part A: Polym. Chem., 2012, 50, 1026.
204 J. Ma, Y. Pang, M. Wang, J. Xu, H. Ma and X. Nie, J. Mater.
Chem., 2012, 22, 3457.
205 J. Ma, X. Yu, J. Xu and Y. Pang, Polymer, 2012, 53, 4145.
206 J. Zhu, J. Cai, W. Xie, P.-H. Chen, M. Gazzano, M. Scandola
and R. A. Gross, Macromolecules, 2013, 46, 796.
207 L. Wu, R. Mincheva, Y. Xu, J.-M. Raquez and P. Dubois,
Biomacromolecules, 2012, 13, 2973.
208 T. Pan, J. Deng, Q. Xu, Y. Zuo, Q.-X. Guo and Y. Fu,
ChemSusChem, 2013, 6, 47.
209 L. Mialon, A. G. Pemba and S. A. Miller, Green Chem., 2010,
12, 1704.
210 S. A. Miller and L. Mialon, US 2013/0137847 A1, 2013.
211 M. Firdaus and M. A. R. Meier, Eur. Polym. J., 2013, 49, 156.
212 C. S. Lovell, M. E. Ries, I. M. Ward, H. Montes de Oca and
D. Farrar, Macromolecules, 2013, 46, 1201.
213 W. Dong, H. Li, M. Chen, Z. Ni, J. Zhao, H. Yang and
P. Gijsman, J. Polym. Res., 2011, 18, 1239.
214 W. Dong, J. Ren, L. Lin, D. Shi, Z. Ni and M. Chen, Polym.
Degrad. Stab., 2012, 97, 578.
215 O. Kreye, S. Oelmann and M. A. R. Meier, Macromol. Chem.
Phys., 2013, 214, 1452.
216 N. Matsumi, N. Nakamura and K. Aoi, Polym. J., 2008, 40,
400.
217 Z.-J. Sun, B. Sun, R.-B. Tao, X. Xie, X.-L. Lu and D.-L. Dong,
J. Biomed. Mater. Res., Part A, 2013, 101, 253.
218 M. Colonna, C. Berti, M. Fiorini, E. Binassi, M. Mazzacurati,
M. Vannini and S. Karanam, Green Chem., 2011, 13,
2543.
219 H. B. Abderrazak, A. Fildier, S. Marque, D. Prim,
H. B. Romdhane, H. R. Kricheldorf and S. Chatti, Eur.
Polym. J., 2011, 47, 2097.
220 S. Chatti, S. M. Weidner, A. Fildier and H. R. Kricheldorf,
J. Polym. Sci., Part A: Polym. Chem., 2013, 51, 2464.

Polym. Chem.

Review
221 C. Japu, A. Martnez de Ilarduya, A. Alla, M. G. GarcaMartn, J. A. Galbis and S. Mu~
noz-Guerra, Polym. Chem.,
2013, 4, 3524.
222 C. Lavilla, A. Alla, A. Martnez de Ilarduya, E. Benito,
M. G. Garca-Martn, J. A. Galbis and S. Mu~
noz-Guerra,
J. Polym. Sci., Part A: Polym. Chem., 2012, 50, 3393.
223 C. Lavilla, A. Martnez de Ilarduya, A. Alla, M. G. GarcaMartn, J. A. Galbis and S. Mu~
noz-Guerra, Macromolecules,
2012, 45, 8257.
224 C. Lavilla and S. Mu~
noz-Guerra, Green Chem., 2013, 15,
144.
225 A. Celli, P. Marchese, S. Sullalti, J. Cai and R. A. Gross,
Polymer, 2013, 54, 3774.
226 Y. Liu, M. S. Lindlad, E. Ranucci and A.-C. Albertsson,
J. Polym. Sci., Part A: Polym. Chem., 2001, 39, 630.
227 C. Lluch, J. C. Ronda, M. Gali`
a, G. Lligadas and V. C
adiz,
Biomacromolecules, 2010, 11, 1646.
228 J. M. Rochette and V. S. Ashby, Macromolecules, 2013, 46,
2134.
229 M. Okada, M. Yamada, M. Yokoe and K. Aoi, J. Appl. Polym.
Sci., 2001, 81, 2721.
230 X. Li, Y. Su, Q. Chen, Y. Lin, Y. Tong and Y. Li,
Biomacromolecules, 2005, 6, 3181.
231 P. Karimi, A. S. Rizkalla and K. Mequanint, Materials, 2010,
3, 2346.
232 C. Vilela, L. Cruciani, A. J. D. Silvestre and A. Gandini,
Macromol. Rapid Commun., 2011, 32, 1319.
233 C. Vilela, L. Cruciani, A. J. D. Silvestre and A. Gandini, RSC
Adv., 2012, 2, 2966.
234 C. Vilela, A. J. D. Silvestre and A. Gandini, J. Polym. Sci., Part
A: Polym. Chem., 2013, 51, 2260.
235 S. Mallakpour and P. Asadi, Colloid Polym. Sci., 2010, 288,
1341.
236 S. Mallakpour, P. Asadi and M. R. Sabzalian, Amino Acids,
2011, 41, 1215.
237 Y. Parulekar and A. K. Mohanty, Green Chem., 2006, 8, 206.
238 L. Yu, S. Petinakis, K. Dean, A. Bilyk and D. Wu, Macromol.
Symp., 2007, 249-250, 535.
239 R. Bhatt, D. Shah, K. C. Patel and U. Trivedi, Bioresour.
Technol., 2008, 99, 4615.
240 U. Konwar, G. Das and N. Karak, J. Appl. Polym. Sci., 2011,
121, 1076.
241 S. Nurkhamidah, E. M. Woo, I.-H. Huang and C. C. Su,
Colloid Polym. Sci., 2011, 289, 1563.
242 F. Cock, A. A. Cuadri, M. Garca-Morales and P. Partal,
Polym. Test., 2013, 32, 716.
243 C.-S. Wu, F.-S. Yen and C.-Y. Wang, Polym. Bull., 2011, 67,
1605.
244 C.-S. Wu, Polym. Degrad. Stab., 2012, 97, 64.
245 C.-S. Wu, Y.-C. Hsu, J. Yeh, H.-T. Liao, J.-J. Jhang and
Y.-Y. Sie, Carbohydr. Polym., 2013, 94, 584.
246 C.-S. Wu, J. Polym. Environ., 2013, 21, 421.
247 M. A. Rahman, D. De Santis, G. Spagnoli, G. Ramorino,
M. Penco, V. T. Phuong and A. Lazzeri, J. Appl. Polym. Sci.,
2013, 129, 202.
248 R. T. Mathers and M. A. R. Meier, Green Polymerization
Methods: Renewable Starting Materials, Catalysis and Waste

This journal is The Royal Society of Chemistry 2013

View Article Online

Review

Published on 01 November 2013. Downloaded by TECHNISCHE UNIVERSITEIT EINDHOVEN on 11/11/2013 21:25:50.

249
250
251
252

253
254
255
256
257
258
259
260
261

Reduction, Wiley-VCH Verlag GmbH & Co. KGaA,


Weinheim, Germany, 2011.
K. Loos, Biocatalysis in Polymer Chemistry, Wiley-VCH
Verlag GmbH & Co. KGaA, Weinheim, Germany, 2010.
U. K. Thiele, Chem. Fibers Int., 2004, 54, 162.
K. Pang, R. Kotek and A. Tonelli, Prog. Polym. Sci., 2006, 31,
1009.
Y. Doi and A. Steinb
uchel, Biopolymers, Applications and
Commercial Products Polyesters III, Wiley-VCH, Weiheim,
2002.
R. Nagahata, J.-I. Sugiyama, M. Goyal, M. Asai, M. Ueda and
K. Takeuchi, J. Polym. Sci., Part A: Polym. Chem., 2000, 38, 3360.
A. Takasu, A. Takemoto and T. Hirabayashi,
Biomacromolecules, 2006, 7, 6.
K. Manabe, S. Iimura, X. M. Sun and S. Kobayashi, J. Am.
Chem. Soc., 2002, 124, 11971.
K. Manabe, X.-M. Sun and S. Kobayashi, J. Am. Chem. Soc.,
2001, 123, 10101.
T. Yashiro, H. R. Kricheldorf and S. Huijser, Macromol.
Chem. Phys., 2009, 210, 1607.
P. Buzin, L. Mohammed, G. Schwarz and H. R. Kricheldorf,
Macromolecules, 2008, 41, 8491.
A. Takasu, Y. Shibata, Y. Narukawa and T. Hirabayashi,
Macromolecules, 2007, 40, 151.
T. Yashiro, H. R. Kricheldorf and S. Huijser, J. Macromol.
Sci., Part A: Pure Appl.Chem., 2010, 47, 202.
M. de Geus, I. van der Meulen, B. Goderis, K. van Hecke,
M. Dorschu, H. van der Wer, C. E. Koning and A. Heise,
Polym. Chem., 2010, 1, 525.

This journal is The Royal Society of Chemistry 2013

Polymer Chemistry
262 J. Cai, C. Liu, M. Cai, J. Zhu, F. Zuo, B. S. Hsiao and
R. A. Gross, Polymer, 2010, 51, 1088.
263 M. A. J. Veld, A. R. A. Palmans and E. W. Meijer, J. Polym.
Sci., Part A: Polym. Chem., 2007, 45, 5968.
264 A. Mahapatro, A. Kumar and R. A. Gross,
Biomacromolecules, 2004, 5, 62.
265 A. Mahapatro, B. Kalra, A. Kumar and R. A. Gross,
Biomacromolecules, 2003, 4, 544.
266 T. Takamoto, P. Kerep, H. Uyama and S. Kobayashi,
Macromol. Biosci., 2001, 1, 223.
267 C.-H. Wong and G. Whitesides, Enzymes in synthetic organic
chemistry, Pergamon, Oxford, 1994.
268 Y. Yu, D. Wu, C. Liu, Z. Zhao, Y. Yang and Q. Li, Process
Biochem., 2012, 47, 1027.
269 R. A. Gross, M. Ganesh and W. Lu, Trends Biotechnol., 2010,
28, 435.
270 A. S. Kulshrestha, W. Gao and R. A. Gross, Macromolecules,
2005, 38, 3193.
271 Y.-Y. Linko, M. L
ams
a, X. Wu, E. Uosukainen, J. Sepp
al
a
and P. Linko, J. Biotechnol., 1998, 66, 41.
272 S. N. Vouyiouka, E. Topakas, A. Katsini, C. D. Papaspyrides
and P. Christakopoulos, Macromol. Mater. Eng., 2013, 298,
679.
273 E.-M. Dukuzeyezu, H. Lefebvre, M. Tessier and A. Fradet,
Polymer, 2010, 51, 1218.
274 X. Q. Xiong and Y. H. Xu, Polym. Bull., 2010, 65, 455.
275 S. Mecking, Angew. Chem., Int. Ed., 2004, 43, 1078.
276 A. U. B. Queiroz and F. P. Collares-Queiroz, Polym. Rev.,
2009, 49, 65.

Polym. Chem.

Вам также может понравиться