Você está na página 1de 47

See

discussions, stats, and author profiles for this publication at:


https://www.researchgate.net/publication/234076748

The chemomechanical properties of


microbial polyhydroxyalkanoates

Conference Paper in Progress in Polymer Science · January 2012


Impact Factor: 26.93 · DOI: 10.1016/j.progpolymsci.2012.06.003

CITATIONS READS

89 180

5 authors, including:

Peter J. Halley Steven Pratt


University of Queensland University of Queensland
272 PUBLICATIONS 3,651 CITATIONS 69 PUBLICATIONS 1,181 CITATIONS

SEE PROFILE SEE PROFILE

Alan G Werker Paul Lant


AnoxKaldnes AB University of Queensland
68 PUBLICATIONS 1,021 CITATIONS 135 PUBLICATIONS 3,758 CITATIONS

SEE PROFILE SEE PROFILE

Available from: Bronwyn Laycock


Retrieved on: 04 May 2016
Progress in Polymer Science 39 (2014) 397–442

Contents lists available at ScienceDirect

Progress in Polymer Science


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

Review

The chemomechanical properties of microbial


polyhydroxyalkanoates
Bronwyn Laycock a,b,∗ , Peter Halley a,b,1 , Steven Pratt a,2 ,
Alan Werker c,3 , Paul Lant a,4
a
School of Chemical Engineering, The University of Queensland, St Lucia, Qld 4072, Australia
b
Australian Institute for Bioengineering and Nanotechnology, The University of Queensland, Brisbane, Qld 4072, Australia
c
AnoxKaldnes AB, Klosterängsvägen 11A, SE-226 47 Lund, Sweden

a r t i c l e i n f o a b s t r a c t

Article history: Microbially produced polyhydroxyalkanoates (PHAs) are fully biodegradable biopolyesters
Received 22 May 2013 that have attracted much attention recently as alternative polymeric materials that can be
Accepted 3 June 2013
produced from biorenewable and biowaste resources. The properties of these biological
Available online 21 August 2013
polymers are affected by the same fundamental principles as those of fossil-fuel derived
polyolefins, with a broad range of compositions available based on the incorporation of
Keywords:
different monomers into the PHA polymer structure, and with this broad range tailoring
Biopolymer
Polyhydroxyalkanoates subsequent properties. This review comprehensively covers current understanding with
Microbial polyesters respect to PHA biosynthesis and crystallinity, and the effect of composition, microstructure
Poly((R)-3-hydroxybutyrate) and supramacromolecular structures on chemomechanical properties. While polymer com-
Poly((R)-3-hydroxybutyrate-co-3- position and microstructure are shown to affect these properties, the review also finds that
hydroxyvalerate) a key driver for determining polymer performance properties is compositional distribution.
Mechanical properties From this review it follows that PHA–PHA blend compositions are industrially important,
and the performance properties of such blends are discussed. A particular need is identified
for further research into the effect of chemical compositional distribution on macromolec-
ular structure and end-use properties, advanced modeling of the PHA accumulation process
and chain growth kinetics for better process control.
© 2013 Elsevier Ltd. All rights reserved.

1. Introduction

Bacterial polyhydroxyalkanoates (PHAs) are a unique


family of polymers that act as a carbon/energy store
for more than 300 species of Gram-positive and Gram-
DOI of the original article: negative bacteria as well as a wide range of archaea [1].
http://dx.doi.org/10.1016/j.progpolymsci.2012.06.003.
∗ Corresponding author at: School of Chemical Engineering, The Uni-
Synthesized intracellularly as insoluble cytoplasmic inclu-
versity of Queensland, St Lucia, Qld 4072, Australia. Tel.: +61 7 3346 3188;
sions in the presence of excess carbon when other essential
fax: +61 7 3346 3973. nutrients such as oxygen, phosphorous or nitrogen are
E-mail addresses: b.laycock@uq.edu.au (B. Laycock), limited, these polymeric materials are able to be stored at
p.halley@uq.edu.au (P. Halley), s.pratt@uq.edu.au (S. Pratt), high concentrations within the cell since they do not sub-
Alan.Werker@anoxkaldnes.com (A. Werker), paul.lant@uq.edu.au
stantially alter its osmotic state [2]. The resulting polymers
(P. Lant).
1
Fax: +61 7 3346 3973.
are piezoelectric and perfectly isotactic/optically active
2
Fax: +61 7 3346 7843. (having only the (R)-configuration). They are hydrophobic,
3
Fax: +46 46 133201. water-insoluble, inert and indefinitely stable in air, and
4
Fax: +61 7 3365 4728. are also thermoplastic and/or elastomeric, non-toxic and

0079-6700/$ – see front matter © 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.progpolymsci.2013.06.008
398 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

Nomenclature
kDa kilodalton
2D two-dimensional lc lamellar core distance
3D three-dimensional Lp long period distance
13 C carbon-13 Mn number-average molecular weight
3HB 3-hydroxybutyrate Mw weight-average molecular weight
3HD 3-hydroxydecanoate mcl medium chain length
3HDD 3-hydroxydodecanoate MDa megadalton
3HHx 3-hydroxyhexanoate MPa megapascals
3H2MB 3-hydroxy-2-methylbutyrate ␮m micrometer
3H2MV 3-hydroxy-2-methylvalerate nA Avrami index
3H4MV 3-hydroxy-4-methylvalerate NADH nicotinamide adenine dinucleotide
3HO 3-hydroxyoctanoate (reduced form)
3HP 3-hydroxypropionate NADPH nicotinamide adenine dinucleotide phos-
3HPE 3-hydroxy-4-pentenoate phate (reduced form)
3HTD 3-hydroxytetradecanoate nm nanometer
3HV 3-hydroxyvalerate NMR nuclear magnetic resonance
4HB 4-hydroxybutyrate P(3HB) poly(3-hydroxybutyrate)
ACP acyl carrier protein P(3HB-co-3HV) poly(3-hydroxybutyrate-co-3-
Ae aerobic hydroxyvalerate)
ADF aerodynamic feeding P(3HB-co-4HB) poly(3-hydroxybutyrate-co-4-
AFM atomic force microscopy hydroxybutyrate)
An anaerobic P(3HB-co-3HHx) poly(3-hydroxybutyrate-co-3-
BOD biochemical oxygen demand hydroxyhexanoate)
CCD chemical compositional distribution PAOs polyphosphate accumulating organisms
CH2 methylene group PDI polydispersity index
CoA coenzyme A PEG polyethylene glycol
COD chemical oxygen demand PHA polyhydroxyalkanoate
Da Dalton PHO polyhydroxyoctanoate
DMTA dynamic mechanical thermal analysis Pn number average degree of polymerization
DO dissolved oxygen R R enantiomeric form
DOC dissolved organic carbon concentration RNA ribonucleic acid
DSC differential scanning calorimetry SBR sequencing batch reactor
DP degree of polymerization scl short chain length
Ea activation energy SDS-PAGE sodium dodecyl sulfate polyacrylamide
EBPR enhanced biological phosphorus removal gel electrophoresis
ESI-MS electrospray ionization multistage mass t time
spectrometry Tc crystallization temperature
FBB fraction of the BB diad sequence in a P(3HB- Tcc cold crystallization temperature
co-3HV) copolymer Tg glass transition temperature
FBV fraction of the BV diad sequence in a P(3HB- Tm melting temperature
co-3HV) copolymer Tm0 equilibrium melting temperature
FVB fraction of the VB diad sequence in a P(3HB- TCA tricarboxylic acid cycle
co-3HV) copolymer TEM transmission electron microscopy
FVV fraction of the VV diad sequence in a P(3HB- UHMW ultra high molecular weight
co-3HV) copolymer UV ultraviolet
FAB-MS fast atom bombardment mass spectroscopy VFAs volatile fatty acids
FT-IR Fourier-transform infrared WAXD wide angle X-ray diffraction
g gram WAXS wide angle X-ray scattering
G growth rate of spherulites WWTP wastewater treatment plant
GAOs glycogen accumulating organisms Xt percent crystallinity
GPa gigopascals XRD X-ray diffraction
GPC gel permeation chromatography
Hf heat of fusion of the melting peak
Hf 100% heat of fusion of a 100% crystalline material have very high purity within the cell [3–8]. PHA has a
HB hydroxybutyrate much better resistance to UV degradation than polypro-
HDPE high density polyethylene pylene but is less solvent resistant. Most importantly,
HV hydroxyvalerate these biopolymers are completely biodegradable. To
kA crystallization rate constant date more than 150 different monomer units have been
incorporated into biological PHA [3–8], and the polymer
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 399

properties available within this family are as a result very [22], with an inoculum of bacteria being introduced into
broad. In general, PHAs can be divided into two main a sterile solution of trace metal nutrients and a suitable
groups, these being the short-chain-length PHAs (scl- carbon source and nutrients in the first (growth) stage. In
PHAs) that contain monomer units with 3–5 carbon atoms, the second stage, an essential nutrient (such as N, P or O2 )
and the medium-chain-length PHAs (mcl-PHAs) that is deliberately limited and PHA accumulation takes place.
contain monomer units of 6–18 carbon atoms. The most The properties of the final polymer depends on the mix of
common PHAs are poly(3-hydroxybutyrate) (P3HB) and carbon feedstocks fed during accumulation, the metabolic
poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (P(3HB- pathways that the bacteria use for the following conver-
co-3HV)). These materials have mechanical properties that sion into precursors, and the substrate specificities of the
are comparable to those of polypropylene and polyethyl- enzymes involved [39].
ene, although they have much lower elongation to break
and are more brittle. 2.2. PHA production in mixed cultures
There are a number of reviews of PHA polymers and
their crystallinity [9–26]. However, the effect of micro- PHA production based on open mixed cultures has
structure and composition, and in particular the influence been proposed as a means of lowering production costs
of compositional distribution/blending, on mechanical [40,41]. No reactor sterilization is necessary and the cul-
properties has not been a focus. This is addressed in this ture is able to adapt to various complex (cheap) waste
review. Special attention is paid to higher HV content feedstocks [40–46]. Gurieff and Lant [47] undertook a life
copolymers (>20 mol% 3HV), the properties of which have cycle assessment and financial analysis of mixed culture
been infrequently reported. The fundamentals of PHA bio- PHA production using an industrial wastewater as the feed-
chemistry and crystallinity are given as a background, stock, and found that PHA production was preferable to
with the scope of the review being summarized in Fig. 1. biogas production and was financially attractive in com-
It is believed that a clear understanding of the relation- parison to pure culture PHA production, while both PHA
ships between PHA crystallinity and polymer composition, production processes had similar environmental impacts
compositional distribution, microstructure and blend com- that were significantly lower than HDPE production.
position will help inform future polymer development The synthesis of PHA in mixed cultures was first
within this field. observed in 1974 in wastewater treatment plants designed
for biological phosphorus removal (EBPR) [48], when
2. Overview of PHA production Wallen and Rohwedder reported PHA heteropolymers in
the chloroform extracts of the activated sewage sludge.
2.1. PHA production in pure cultures Most mixed culture production relies on ecological selec-
tion pressures that favor organisms with elevated PHA
The French bacteriologist Lemoigne first isolated storage capacity, thus engineering the microbial consor-
and characterized poly-3-hydroxybutyrate (P(3HB)) from tium [45,49,50]. One such method is to enrich a culture
bacteria between 1923 and 1927 [27–33], and showed in PHA-accumulating organisms by taking the culture
that this extract could be cast into a transparent film. through a number of aerobic feast/famine cycles (known as
Although it was some time before this discovery was the aerobic dynamic feeding, or ADF, process) [51,52]. ADF
turned into a practical outcome, PHA production using relies on the removal of a portion of the consortium dur-
pure cultures still has a long history. Stanier and Wilkin- ing each cycle – the organisms that are able to store carbon
son and their co-workers were responsible for some of during feast condition and then use that stored carbon to
the initial fundamental research into the mechanisms of grow during famine conditions are selectively enriched. A
PHA biosynthesis, beginning in 1957 [34,35]. This was fol- modified approach is to use alternating aerobic/anaerobic
lowed in 1959 by the first attempted commercialization, conditions; in this case it is the polyphosphate and glyco-
when W.R. Grace and Company patented a P(3HB) pro- gen accumulating organisms (PAOs and GAOs) that are
duction process using bacteria [36], although production selectively enriched and therefore responsible for PHA
inefficiencies, poor thermal stabilities and a lack of avail- accumulation [53,54].
able extraction technologies limited this application. In Volatile fatty acids (VFAs) can be readily stored as PHA
1970, Imperial Chemical Industries Ltd. commercialized by mixed cultures, so most studies on PHA production
the production of P(3HB-co-3HV) under the trade name of by mixed cultures have been carried out using organic
BiopolTM [37], with the technology since being sold to Mon- acids (such as acetate, propionate, butyrate, and valerate)
santo and then to Metabolix. Since that time, there have [55–58] as feedstocks. Fatty acids are energetically advan-
been a range of new technologies developed, and recent tageous substrates from a metabolic perspective since their
focus within pure culture production has been on the syn- complete ␤-oxidation generates more equivalent chemical
thesis of alternative copolymers such as P(3HB-co-3HHx) energy than the complete oxidation of a molar equiva-
(commercialized as NodaxTM ) [38], a wide range of func- lent of glucose [59]. A major advantage of mixed culture
tionalized PHAs (incorporating monomer units with novel PHA production is the opportunity to use real fermented
and active functionalities on the side chains) and also on wastes as feedstock as opposed to synthetic VFAs, since
the use of metabolic engineering to reengineer the cen- substrate costs are known to be a critical factor in deter-
tral metabolism of PHA producers for more efficient PHA mining the economics of PHA production. The use of low
production. In its simplest form, PHA production using cost agricultural or industrial waste feedstocks such as
pure cultures adopts a two-stage batch production process fermented sugar cane molasses [42,53,60–62], fermented
400 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

Fig. 1. Overview of PHA synthesis: (a) schematic depiction of chain polymerisation catalyzed by enzymes. (b) Schematic depiction of a PHA granule with
granule associated proteins. Reproduced with permission, from Rehm (2003), Biochemical Journal, vol. 376, page 27 © the Biochemical Society [1]. (c)
Schematic depiction of the ␣- and ␤-forms of the PHB polymer chain. (d) Schematic depiction of semi-crystalline polymer structure. (e) AFM image of
PHBV film, courtesy of JPK Instruments AG. (f) Final plastic products.

paper mill effluents [43], saponified sunflower oil [63], fruit Further research is required to understand the effect of
[64] and tomato [65] cannery effluents, fermented munici- key controlling factors such as feeding strategy (continu-
pal sludge [66,67] fermented food industrial and domestic ous or pulse) [79], pH control [80] and microstructure [81]
wastewaters [68,69], bio-oil (a pyrolysis by-product) [70] in mixed cultures, as they may have quite different effects
and fermented olive oil mill effluents [71,72] have all so far under these production conditions compared to pure cul-
resulted in the successful accumulation of PHA, and there ture PHA production. Given the multiple organisms (and
is a growing knowledge base in the use of other feedstocks possibly also PHA production pathways) present in a mixed
such as alcohols and glycerol [11]. culture environment, one of the key questions will be the
A broad range of PHA compositions can be eas- compositional distribution of the polymer produced and
ily produced using mixed cultures based on differ- effect of these blend compositions on chemomechanical
ing feedstocks, with other monomer units such as properties.
3-hydroxyvalerate (3HV), 3-hydroxy-2-methylbutyrate
(3H2MB), 3-hydroxy-2-methylvalerate (3H2MV), and 3-
hydroxyhexanoate (3-HHx) being common components 2.3. Other production techniques
in P(3HB)-based copolymers [58,62,74–76]. For example,
P(3HB-co-3HV) copolymers with mol% HV ranging from Plant-based (photosynthesis fuelled) production sys-
17 to 85% have been recorded based on the propionate tems for PHA have been under intensive investigation
fraction in acetate/propionate feed [58,75] (see Appendix over recent years [82,83]. Poirier first demonstrated
A). Pure cultures by contrast often need large amounts of this approach in 1992 [84] using transgenic Arabidopsis
co-substrates to produce polymers with relatively small expressing acetoacetyl-CoA reductase and PHB synthase
fractions of monomers other than P(3HB) (see for exam- from Ralstonia eutropha. (There have been a series of name
ple [76] where P(3HB-co-3HV) with only 2–8 mol% HV was changes for many of the organisms used for PHA accumula-
produced from mixtures containing equal amounts of veg- tion. For example, Ralstonia eutropha (the most extensively
etable oils and propionic acid (used to produce HV)). This studied bacterium) was originally named Hydrogenomonas
may be due to the fact that mixed cultures contain a diver- eutrophus, then renamed Alcaligenes eutropha, and has
sity of organisms that are likely to employ a range of PHA since been renamed.) The commercial development of
production pathways. Cellular PHA contents and produc- plant-based PHA production processes is now underway,
tion rates comparable to or superior to those of pure culture although there is still much research to be done. For exam-
have now been achieved by Johnson et al. [77,78] using ple, Metabolix has been developing technologies in a range
mixed cultures, with this team producing P(3HB) at 89 wt% of transgenic plants, such as tobacco (Nicotiana tabacum)
of the dried biomass in less than 8 h. [85] and switchgrass [86]. Preliminary investigations are
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 401

also underway into the production of PHA using microalgae acyl-CoA oxidase and enoyl-CoA hydratase. Pathway III
[87]. involves 3-hydroxyacyl-ACP-CoA transferase (encoded by
Finally, PHAs can be synthesized from the relevant sub- PhaG) and malonyl-CoA-ACP transacylase (FabD); sub-
stituted propiolactones using, for example, aluminum or strates are converted to 3-hydroxyacyl-ACP which can
zinc alkyl catalysts with water as a cocatalyst (see [88,89] then form 3-hydroxyacyl-CoA and thus PHA. Pathway IV
for example). However, this process is not cost-competitive uses NADPH-dependent acetoacetyl-CoA reductase to oxi-
for large-scale production and involves the use of toxic dize (S)-(+)-3-hydroxybutyryl-CoA. The other pathways
reagents. are used for the synthesis of alternative copolymers; for
example, pathways V and VII are used to synthesize P(4HB)
2.4. Conclusions containing PHAs (and are used by Clostridium kluyveri and
A. hydrophila 4AK4 respectively).
There are many approaches to the production of PHA As Kessler and Witholt state: “Regulation of PHA
with potential variation in composition and compositional metabolism can take place at different levels: (1) activa-
distribution depending on the microbial diversity of the tion of pha gene expression due to specific environmental
culture used, etc. Regardless of the production strategy signals, such as nutrient starvation; (2) activation of
being adopted, it is of fundamental importance to under- the PHA synthetic enzymes by specific cell components
stand the underlying biological processes and chemistries or metabolic intermediates; (3) inhibition of metabolic
that affect the chemomechanical properties of the final enzymes of competing pathways and therefore enrichment
polymer product. These are reviewed in the following sec- of required intermediates for PHA synthesis; or (4) a combi-
tions. nation of these.” [23]. For example, during normal bacterial
growth, the ␤-ketothiolase in Pathway I is inhibited by
3. PHA fundamentals free coenzyme-A coming out of the Krebs (or TCA) cycle.
But when nutrients other than carbon are limited, acetyl-
3.1. Biosynthesis CoA is restricted from entering into the Krebs cycle and
the excess acetyl-CoA is channeled into PHA biosynthesis
The biology of PHA bioplastics is complex; PHA accu- [95]. If growth is limited for other reasons, protein syn-
mulation is controlled by many genes that encode a range thesis stops, which leads to high concentrations of NADH
of enzymes that are directly or indirectly involved in and NADPH. This inhibits citrate synthase and isocitrate
PHA synthesis [5,10,13,23,25,26,90]. So far, biosynthesis dehydrogenase, which again slows down the Krebs cycle,
of PHA can be summarized in eight pathways [12]. The directing acetyl-CoA towards PHA synthesis [96].
first pathway (Pathway I) is well-studied, and is used The synthesis of the PHA polymer chain takes place
by the model organism Ralstonia eutropha; it involves within the cytoplasm of the bacterial cell, within inclusions
the three key enzymes ␤-ketothiolase, NADPH-dependent known as granules – the biochemistry of this synthetic pro-
acetoacetyl-CoA reductase, and PHA synthase, encoded by cess and of these granules in turn influences the properties
genes phaA, phaB and phaC, respectively. These three genes of the final polymer product.
are found together on an operon whose expression is rela-
tively constant during PHA production [91,92]. The carbon 3.2. Granule structure and development
source is initially converted into coenzyme A thioesters
of (R)-hydroxyalkanoic acid. ␤-Ketothiolase can then cat- Granules in microorganisms were observed under the
alyze the condensation of two coenzyme A thioester microscope by Beijerinck as early as 1888 (reported in [97])
monomers (such as an acetyl-CoA and a propionyl-CoA and may have been known before this. An image of granules
monomer). This is followed by an (R)-specific reduction to in Azotobacter chroococcum (taken using a transmission
give (R)-3-hydroxybutyryl-CoA (or (R)-3-hydroxyvaleryl- electron microscope) is given in Fig. 2 as an example [98].
CoA) (catalyzed by acetoacetyl-CoA reductase), which is The genetically manipulable model organism in PHA
then converted by PHA synthase into PHA [91,93]. At research, Ralstonia eutropha, can accumulate P(3HB) in the
least 88 PHA synthases have been sequenced, with four form of multiple granules of 0.2–0.5 ␮m in size, and can
major classes being identified [94]. Class I uses fatty acids have a polyester content of more than 90% of the cell
with 3–5 carbon atoms; class II uses those with 6–14 dry matter [99], with an average final number of gran-
carbons; and Classes III and IV synthesize short chain ules in a typical PHA-producing cell of around 10 [2,100].
length PHA. In the depolymerization reaction the accu- Steinbüchel et al. [101] presented some calculations on
mulated PHA is hydrolyzed into 3-hydroxybutyrate (3HB) P(3HB) in granules: assuming a relative molecular mass of
by depolymerase (encoded by phaZ), and can then be 3 × 106 Da, then a single P(3HB) molecule would be made
converted back into acetoacetyl-coenzyme A. Pathway up of approximately 35,000 3HB units, with a total length
II is associated with fatty acid uptake by microorgan- of 24,000 nm. Based on the typical density of 1.2 g cm−3 ,
isms and can be used for the synthesis of mcl-PHA. the diameter of a granule of a single P(3HB) molecule
Following fatty acid ␤-oxidation to give acyl-CoA, the pre- would be around 9.7 nm. In a typical granule of 350 nm
cursor is then converted to 3-hydroxyacyl-CoA which can diameter, therefore, there would be around 43,000 P(3HB)
then form PHA under synthase catalysis. Known enzymes molecules. A granule has an amorphous polyester core (see
involved in this pathway include 3-ketoacyl-CoA reduc- Section 3.3). A distinct boundary layer at least 3–4 nm thick
tase, epimerase and (R)-enoyl-CoA hydratase/enoyl-CoA was noted in early thin-section electron microscopy and
hydratase I. Two others are also believed to be involved: was thought to comprise a phospholipid monolayer with
402 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

PHA accumulation. They are however known to control the


amount [106,108], size and number of granules [111], and
possibly also help prevent PHA crystallization [112]. In the
absence of phasins, cells accumulate only one very large
granule, taking up all available space in the cell [111]. It has
also been shown that overexpression of the phasin leads to
smaller granules [99]. Phasins can bind to any hydropho-
bic inclusion regardless of its composition [3] and appear
to bind to the synthase when conditions are not right for
PHA accumulation. Cells have been found to generally stop
increasing their P(3HB) content at around 80–90% of the
cell dry mass. It is known that PHA synthase activity and
intracellular monomer concentration remain high at this
point, so it is likely that physical constraints are at least
one limiting factor for polymer accumulation [113].
Three models at least have been developed to describe
Fig. 2. Transmission electron micrograph of ultrathin section of Azotobac- PHA granule formation in vivo [94,114,115]. One is that
ter chroococcum cell treated with phenylacetic acid.
Reprinted from Nuti et al. [98], with permission © Canadian Science Pub-
micelles form as the polymer grows and the constituents
lishing or its licensors. of the boundary layer become incorporated as the size
increases [113]. A modified version of this model has
recently been proposed [116] in which the initial nucle-
embedded and attached proteins; later results suggested ating species consists of a PhaC dimer covalently attached
that this layer may be as thick as 14 nm [94,102,103]. Other to a PHB chain associated with several phasins that are
results obtained using atomic force microscopy (AFM) on attached to maintain solubility. It is hypothesized that sev-
lysed cells seemed to suggest a smooth outer envelope eral of these nucleating species fuse together due to the
(lipid) with an inner network structure ∼2–4 nm thick hydrophobic patches on the PHB chains, and a soluble gran-
based on PhaP (protein) and a crystalline layer beneath this ule precursor is thus formed. In a second model, a synthase
[104]. However, Western blot and immunogold labeling localizes to the inner face of the cytoplasmic membrane
experiments have shown that the surface of the granule with biosynthesis in the inner membrane space with the
has a high protein content. Electron cryotomography has granules eventually budding off [94,115]. In other studies,
recently been used [105] to examine granule synthesis and nascent PHA granules have been found to localize to the
development in a “near-native”, “frozen-hydrated” state in cell poles or occasionally to the cell poles and to the center
intact cells. All had a discontinuous surface layer more con- of the cells (i.e. the future cell poles) [117,118]. However,
sistent with a partial protein coating than a lipid mono- or fluorescence [118] and immunocytochemical localization
bi-layer [105]. It was proposed that the results of earlier studies [119] have shown that PhaC is located randomly
studies may have been in part affected by artifacts of the throughout the cytoplasm at the surface of the granule, and
preparation methods used as well as differences in nucleoid biosynthetic pathways associated with phospholipid pro-
condensation. duction are downregulated during PHB production [116].
Granule associated proteins can be divided into 4 A third model [109,120] proposes that mediation elements
groups [3]: the PHA synthases (see Section 3.3), PHA serve as scaffolds for the initiation of granule formation,
depolymerases, regulatory proteins (such as PhaR, the tran- which is incompatible with much of the evidence for the
scription factor that negatively regulates PhaP expression second model but for which other recent evidence has been
[105,106]) and phasins (PhaP). Phasins are the most abun- obtained [121]. However, in the recent study by Beeby et al.
dant protein at the granule surface, and are predicted to [105], no granules were found within the cell membranes
have high (∼90%) proportions of an ␣-helical structure nor were scaffolds observed. Instead, all of the granules
[107]. Their expression is induced by PHA accumulation were found towards the center of the cytoplasm. Further
[106], with expression levels increasing and decreasing in work is required to determine which model or combination
parallel with PHA production [108,109]. They can consti- of models is correct.
tute as much as 5% of the total cellular protein, and have
been variously categorized as constituting between 14 and 3.3. Mechanism of polymerization
54% of the granule surface [107,109], although Neumann
et al. [107] suggest that the coverage may be higher than Various mechanisms of PHA polymerization have been
this. Each granule has 1–2 molecules of phasin per PHB proposed. It has been shown that Class 1 PHA synthase from
chain [109]. Phasins are amphiphilic, with a hydropho- R. eutropha exists in both monomeric and dimeric forms
bic domain exposed to the PHA and a hydrophilic domain [122–126]. Artificial priming of the synthase with a trimer
exposed to the cell cytoplasm [3]. They are low molecular of 3-hydroxybutyryl-CoA shifts the equilibrium towards
weight proteins (between 11 and 25 kDa), are non-related the dimeric form and reduces the lag phase for polymer
and share no sequence homology [110]. Phasins are only synthesis, suggesting that the two PHA synthase molecules
produced during PHA accumulation and are thought to combine to form a homodimer with two thiol groups as the
positively influence PHA accumulation although the mech- catalytically active sites [100,122–126]. It is also known,
anism is not clear and they are apparently not essential for from kinetic studies, that the lag phase is dependent on
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 403

synthase concentration [116] and that the elongation rate termination proposed that a nucleophilic “chain transfer
is much faster than the initiation rate [127,128]. In a recent agent” would attack at the active site thioester, thereby
publication [116], Cho et al. reported the first isolation of releasing PHB and terminating elongation [134–138]. Mad-
a Strep-2 tagged PHB synthase (PhaC) directly from native den et al. [135] found through end group analysis by
R. eutropha as opposed to expression from recombinant E. 31 P NMR of polymer derivatised with 2-chloro-4,4,5,5-

coli, and found that it unexpectedly copurifies with PHB and tetramethyl-1,3,2-dioxaphospholane that all groups were
the phasin protein PhaP1 in a soluble form. This enzyme covalently linked to PHA at the carboxyl terminus, and that
showed no lag phase in CoA production, unlike expressed all chain transfer agents possessed one or more hydroxyl
PhaC, and it would seem that it is already primed with poly- groups. These agents may be diol compounds such as ethyl-
meric PHB (HBn where n is large). It was again found that ene glycol, polyethylene glycol (PEG) and propylene glycol
the polymerization rate was faster than the initiation rate. as well as 3-hydroxybutyryl-CoA and water; they may
Several independent studies suggest that the synthase and also be enzymes with a water molecule in its active form
depolymerase are simultaneously present during polymer [134] or 3-hydroxybutyric acid, which is not enzyme bound
synthesis [109]. [135]. The carboxyl group at the terminal of the polymer
It has also been shown [129] that synthases possess a chain is believed to be esterified with the hydroxyl moi-
conserved catalytic triad comprising three residues (cys- ety of the chain transfer agents [137–144]. In support of
teine, aspartic acid and histidine) and that the cysteine this hypothesis, molecular weight is substantially lowered
bonds are involved in covalent catalysis. It is suggested when various alcohols or poly(ethylene glycol)s are added
that the histidine activates the cysteine for nucleophilic to the culture medium [139–141,145]. Tian et al. [109]
attack on 3-hydroxybutyryl-CoA, while the aspartic acid also showed that a surface-exposed amino acid was able
may function as a general base catalyst to activate the to hydrolyze the polyester chain in a chain transfer reac-
hydroxyl group of HBCoA for ester formation [7]. tion. Likewise, an increasing concentration of methanol has
Two mechanisms have been proposed for chain elon- been shown to have a large negative effect on the molec-
gation to fit with these observations [130] (Fig. 3). Both ular weight of P(3HB) produced by Protomonas extorquens
involve chain elongation from acylated cysteine. The first [145].
requires only one active cysteine site, and involves both However, Lawrence et al. [92] proposed alternative
covalently and noncovalently bound HBn (CoA) interme- models, based on the use of an alternative substrate ((R)-3-
diates. The second requires that the active site be at the hydroxybutyryl-N-acetylcysteamine, HB-NAC). One option
interface of two PhaC monomers (with PHA synthase being is that chain transfer actually occurs through cleavage of an
present in a dimeric form), with the chain being always internal ester bond within the covalently bound polyester
covalently attached to one of the two synthase units, chain rather than at the thioester linkage; the liberated NAC
switching between them with the addition of each new functional group could act as chain transfer/termination
HA unit. In both mechanisms, the cysteine is activated for agent at either site. Another option is that the noncova-
nucleophilic attack by a base residue such as histidine in lently bound polymer intermediate (as in Fig. 3) covalently
the active site. The authors found through labeling studies linked to NAC could dissociate from the active site. In
using the slower synthase mutant C149S-PhaEC that non- addition, it was proposed that the alcohols used for molec-
covalently bound species were present in the early stages ular weight control may in fact have been associated with
of synthesis, and that these intermediates functioned in the granule and then been involved in a transesterifica-
a chemically and kinetically competent fashion, lending tion reaction during the long fermentation, inducing chain
weight to the possibility that the first mechanism is more hydrolysis [92].
likely. In another proposal, Yamanaka et al. [131] suggested It was further noted [92] that there was a probable
that an active site of PHA synthase could be first capped change in the system during polymerization (for which
with succinyl-CoA (formed by the condensation reaction of there is yet no explanation) that allows termination only
3HB-CoA with succinyl thiolate), although the MALDI evi- after several hundred monomers have been incorporated.
dence on which this hypothesis is partly based could have Results suggest [130] that as the HB chain grows longer,
different interpretations such as a 3HV-terminal unit. the rate of hydrolysis of the covalent linkage to PhaC is
It is well established that microbial PHB (and PHA in dramatically reduced. The cyclic process of simultaneous
general) is amazingly monodisperse [109]. Doi et al. [132] accumulation and turnover of P(3HB) in bacterial cells has
estimated the concentration of PHA synthase in a cell (for been demonstrated by Doi et al. [146] and Taidi et al. [147]
R. eutropha) to be 5 ± 1 × 10−7 mol L−1 , making the number under nitrogen limitation conditions [148].
of PHA synthase molecules per cell approximately 18,000 In 1986 it was believed that the P(3HB) in granules was
(i.e. a ratio of around 42:1 P(3HB) molecules to P(3HB) present as a crystalline solid, so the observation (using
synthase molecules based on Steinbüchel’s result ([101], solution state NMR) that P(3HB) in live cells was in an
Section 3.2)). Tian et al. [109] also found that the ratio of amorphous state was unexpected [149–152]. It was found
PHB chains to PhaC class I synthase was ∼60, with 1–2 that treating the cells to deactivate the granules also led
molecules of PhaP1 (phasin) per chain [109,116]. Hence it to a loss of high resolution in the NMR spectrum, with the
has been supposed that there are many chain termination simultaneous appearance of crystalline P(3HB), as judged
and reinitiation events during granule formation [111]. by X-ray diffraction (XRD). Purified PHAs and even isolated
The mechanism of chain transfer or chain termination is PHA granules are known to crystallize rapidly under some
not well understood at this point [105,116]. Kawaguchi and conditions. This suggests an important role for the granule
Doi [133] in the original model for the mechanism of chain surface layer in preventing crystal nucleation [19,112].
404 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

Fig. 3. Proposed polymerisation mechanism for the synthesis of PHA.


Reprinted (adapted) with permission from Li et al. [130] Copyright (2009) American Chemical Society.

There are two general hypotheses to explain this result. volume and therefore on the third power of the granule
One is that the amorphous state of P(3HB) in vivo can be radius. The upper limit for the rate constant of spontaneous
caused by a simple physical-kinetic limitation [153,154] crystal nucleation in isolated P(3HB) at 30 ◦ C (the tempera-
within granules based on their small size (stabilized by an ture of bacterial growth) is 2.5 events mm−3 s−1 [155,156].
amphiphilic shell). This assumes that the rate of P(3HB) The corresponding rate of nucleation for a typical storage
crystallization in vivo is governed by the rate of sponta- granule of diameter 0.25 ␮m is 2.0 × 10−11 events s−1 . Pro-
neous nucleation events, which depends upon the granule vided the granules do not coalesce and are not exposed
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 405

to foreign nucleating agents, native granules should have chains within the unit cell are nearly at the same level,
a crystallization half-life of at least 3.4 × 1010 s, or >l000 and the angles between dipoles was calculated to be
years [112]. Even at a maximum nucleation rate of less than ∼60◦ . From this it was concluded that the dipole–dipole
100 events mm−3 s−1 , the average lifetime for a granule interaction is one of the main factors governing the molec-
before crystallization should still exceed several months. ular packing [164]. P(3HV) is also orthorhombic with a
The in vivo state is therefore proposed to be under space group of P21 21 21 (D42 ), this time with dimensions
kinetic, not thermodynamic control [112,154]. Horowitz a = 0.592 nm, b = 1.008 nm and c (fiber axis) = 0.556 nm
and Sanders simulated the formation of granules using [165,166]. It also has a twofold screw left-handed sym-
synthetic surfactant to coat the surface of submicron-size metry along the molecular axis. Crystallographic data for
artificial granules, which were shown to be amorphous and a range of mcl-PHA copolymers is reported in Sudesh and
stable in suspension, and which had the same size, den- Abe [9].
sity, morphology and molecular mobility as native granules The crystalline morphology of folded single chain crys-
[112,157]. tals has been shown to conform with the classical folding
The other theory is that the water present in the granule behavior of synthetic linear aliphatic/thermoplastics; the
at around 5–10% of the total content could act as a plasti- crystallographic a-axis runs along the direction of each
cizer, forming hydrogen bonds with the carbonyl groups crystal and on the crystal surface carboxyl groups and
of P(3HB) [19]. The proposal is that this hydrogen bonding hydroxyl groups seem to exist alternately [167]. PHA
prevents strong dipole–dipole interactions from occurring adopts a helical conformation due to the formation of a
between the C O and CH3 groups in P(3HB) [19,158,159]. favorable carbonyl oxygen/methyl interaction, unlike pol-
Once this P(3HB)-bound water is removed, in situ crystal- ypropylene for which the driving force is unfavorable
lization can then occur due to increased hydrogen bonding, methyl/methyl (van der Waals) interactions [168,169].
in a similar manner to the crystallization process on cooling Because of this bipolar interaction between continuous
for pure P(3HB) [19,158]. anti-parallel helical chains, stabilized by the ester moieties,
Once released from the granule structure, PHA crys- no intramolecular hydrogen bonds are needed to stabi-
tallinity then develops and is core to the overall end-use lize the helix, unlike proteins [16]. The crystal structure
properties of PHA including their physical, chemical, at the surface of P(3HB) thin films has been studied using
mechanical and rheological characteristics [148]. infrared reflectance-absorption and transmission spectra;
the results show that the C H· · ·O C hydrogen bonding of
4. Crystal structures of PHA P(3HB) exists along the a-axis (not the b-axis) between the
CH3 group of one helix and the C O group of another helix
In plates and films and similar bulky forms, P(3HB) is [170].
usually a multilamellar system that has been aggregated Single crystals with well-defined structures are
into multi-oriented lamellar crystals [160]. Spherulites are the “monolamellar” form of PHA. Single crystals of
usually formed when P(3HB) chains are crystallized from P(3HB) have been prepared from a dilute solution of
the melt in bulk materials; the lamellar crystals grow radi- polymer in many kinds of organic solvents, such as
ally with twisting. The crystallographic a axis is radial with chloroform/ethanol [169,171], propylene carbonate
b and c axes rotating about it, and as a result the spherulites [160,172–174], poly(ethylene glycol) [175], etc. The thick-
usually have a banded texture [155,161]. The patterns of ness is dependent on many factors including the molecular
this banding will change depending both on the crystal- weight, solvent, and crystallization temperature. Typi-
lization temperature and the molecular weight. Lamellar cally P(3HB) forms lathe-shaped single crystals, around
thicknesses in spherulites typically range from 4 nm to ∼5–10 ␮m long and ∼0.3–2 ␮m wide. These single crystals
10 nm depending on the crystallization temperature [162]. are usually 4–10 nm thick, depending on the molecular
In order to understand the nature of these macrostructures weight, solvent and crystallization temperature. Solution
though, it is important to understand the basic crystal units grown lamellar crystals of P(3HV) (both synthetic and
from which the lamellae are constructed, and the influence bacterial) are square-shaped with a width of 2–4 ␮m and
of comonomer composition on those crystals. a thickness of 5–6 nm [168]. From the electron diffraction
diagram, the polymer chains are aligned perpendicular to
4.1. The ˛-form crystals of PHA the base of the crystal. “The observation of a screw dislo-
cation morphology suggests that the single crystals appear
The most common crystal structure of the P(3HB) to be formed by the superposition of several monolamellar
molecule is the ␣-form, which is produced under the crystals” [168].
typical conditions of melt, cold or solution crystalliza- Single crystals of P(3HB-co-3HV) have also been pre-
tion. In oriented fibers it has been characterized using pared [16,167,172,176]. Mitomo et al. [172] looked at single
X-ray diffraction analysis, and has been shown to adopt crystals with six different 3HV contents (at <30% HV). Up
a left-handed 21 helix (G2 T2 )2 (␣-form). The unit cells to 10 mol% of HV, the crystals were very similar in mor-
are orthorhombic with a space group of P21 21 21 (D42 ) and phology to those of P(3HB). However, between 17 and
with dimensions a = 0.576 nm, b = 1.320 nm and c (the fiber 30 mol% HV, the crystals had morphological irregularities
period) = 0.596 nm [163]. Two molecules of the polyester and a bumpy surface. In P(3HB-co-3HV) spherulites, the
pass through the unit cell and the chains are anti-parallel. long spacing (Lp ) and lamellar core thickness (lc ) values
Conformational analysis based on potential energy calcu- are 7–24 nm and 4–17 nm respectively when the (R)-3HV
lations indicate that the ester groups of the antiparallel content is <21 mol%.
406 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

According to Marchessault et al. [177], mcl-PHAs could crystals in the FT-IR (Fourier-transform infrared) spec-
be classified as thermotropic liquid crystals. The conforma- trum of stretched P(3HB), with bands at 966, 935, 908,
tion was found to be “herringbone” or “comblike” – and it 858, and 1735 cm−1 . For two-step drawn films these peaks
was not found to be possible to distinguish between the two are relatively weak and the crystallinity is small indicat-
possibilities. Polyhydroxyoctanoate (PHO) was found to be ing a disordered structure. These peaks disappeared with
about 25% crystalline. The structure was pseudo-network annealing at higher temperatures.
in character with an initial modulus around 10–20 times The structure of P(3HB-co-3HV) at a clay surface in a
greater than that of a slightly crosslinked rubber. composite is similar to the ␤-form but more ordered, with
chains in a planar zigzag conformation in parallel with
4.2. The ˇ-form crystals of PHA the silica layers. The WAXS reflection (d = 0.480 nm) cor-
responds to the mean distance of P(3HB-co-3HV) chains
The ␤-form of the P(3HB) crystal is a strain-induced in a pseudohexagonal structure. This crystal form disap-
paracrystalline structure of highly extended chains with pears with annealing [188]. Antipov et al. [189] showed
a fiber period of 0.46 nm corresponding to a twisted pla- that when gel spinning was used to obtain highly oriented
nar zigzag conformation [178]. It has an orthorhombic fibers, a third phase (in addition to the crystalline and
crystal system with unit cell parameters of a = 0.576 nm, amorphous) was present and was identified as a columnar
b = 1.320 nm, and c (fiber axis) = 0.598 nm and space group mesophase with 2D-pseudohexagonal packing of the cen-
P21 21 21 . It can be obtained under such conditions as further tral axes of the polymers; it is reversibly strain induced and
drawing of uniaxially stretched films, and is metastable – it is a thermally stable phase formed only under mechanical
can be annealed back to the ␣-form at 130 ◦ C, with a result- stress.
ing increase in crystallinity. The ␤-form was first observed
in a stretched P(3HB) film by Yokouchi et al. [163], and also
4.3. Isodimorphism in P(3HB-co-3HV) copolymers
obtained in uniaxially cold-stretched films of P(3HB) and
P(3HB-co-3HV) [179]. From the calculated value of the fiber
Semicrystalline random copolymers can on rare occa-
repeat length for fully extended P(3HB), the ␤-form has a
sions be isodimorphic in nature, whereby two or more
near completely extended chain conformation. The shift to
comonomer units can exist in the crystal unit structure
the ␤-form did not require prior alignment of the ␣-form
of the other monomer units in the copolymer. As a result
lamellae, but followed orientation of the free chains in the
there is approximately the same high degree of crystallinity
amorphous regions between the crystals [180]. In partic-
across the compositional range, with little resulting dis-
ular, the tie molecules between the lamellar crystals were
ruption to the crystal unit dimensions. There is usually
strongly extended, which leads to high mechanical proper-
pseudoeutectic melting behavior and a structural transi-
ties. Techniques such as melt spinning and drawing of PHA
tion occurs near the pseudoeutectic composition [190,191].
fibers or one- or two-stage cold drawing of films followed
P(3HB-co-3HV) is one such isodimorphic copolymer. A
by annealing have been shown to dramatically increase the
relatively high crystallinity (>50%) has generally been
tensile strength, Young’s modulus and elongation to break
observed across the entire range of HV content (0–95 mol%)
of PHA materials and are associated with the introduc-
in P(3HB-co-3HV) (by XRD); when the materials have been
tion of a ␤-crystalline form as at least part of the lamellar
carefully fractionated and allowed time to fully develop
structure [180]. These properties remain unchanged after
their crystallinity [192] there is evidence of a minimum at
several months suggesting that secondary crystallization
the pseudoeutectic (Fig. 4).
is suppressed when materials are highly orientated and
crystalline [181].
A different mechanism for the formation of the ␤-form
crystals was proposed for one-step-drawn P(3HB-co-3HV)
fibers after isothermal crystallization [182]; it was sug-
gested that many small crystals grow in the amorphous
region during slow isothermal crystallization near the Tg ;
the ␤-structure is then developed by the stretching of
molecular chains in the constrained amorphous region.
Iwata et al. [183] have also reported a 3D crystal structure of
the ␤-structure that has lattice parameters of a = 0.528 nm,
b = 0.920 nm, and c (fiber axis) = 0.469 nm as an ortho-
rhombic crystal system. Etching with enzymes suggests
that the structure of stretched films is stretched chain core
in a shish-kebab morphology with the plates normal to the
stretching direction [184,185].
Nishiyama et al. [186] used a 2D NMR 13 C off-magic-
angle spinning technique to look at the ␣- and ␤-forms of
the P(3HB) crystal. The orientation of the carbonyl carbons
Fig. 4. Effect of 3HV content in P(3HB-co-3HV) on polymer crystallinity
was shown to be in good agreement with the molecu-
by XRD.
lar models proposed by Marchessault and others. Sato Reprinted (adapted) with permission from Wang et al. [201]. Copyright
and coworkers [187] also found evidence for the ␤-form (2001) American Chemical Society.
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 407

the non-crystalline region. This pseudoeutectic has instead


been attributed to the disturbance of crystal packing by
the inclusion of HV units in the P(3HB) lattice and vice
versa [190]. At the pseudoeutectic, the polymer structure
in the crystalline region shifts from a P(3HB) type crys-
tal (fiber repeat 0.596 nm), to a P(3HV) crystal one (fiber
repeat of 0.556 nm) (by XRD), with the middle region orig-
inally being supposed to have both crystal types present
[190]; however later work [201] showed that this was due
to blends being present. It was originally proposed that the
3HV units could sit in the P(3HB) crystal without distorting
the crystalline structure or lattice significantly, and like-
wise for the HB units in the HV crystal. However, solution
grown single crystals of P(3HB) and P(3HB-co-3HV) (up to
30% HV) showed a linear increase in melting points and a
hyperbolic increase in long spacings as the crystallization
temperature increases [172]. As the HV content increases,
the lattice indexes for the HB crystal expand up to the max-
imum (a = 0.602 nm, b = 1.343 nm and c = 0.604 nm) and the
crystal is particularly elongated in the a and b parame-
ters; the d spacing in the HB crystal lattice increases. By
contrast, the HV crystal structure slightly contracts in the
b axis but essentially remains unchanged throughout the
range [165,190,194,201,203], so that the HB unit can sit
in the HV crystal lattice without distorting it. Iwata et al.
[204] prepared single crystals of P(3HB-co-8 mol% 3HV)
and analyzed their electron diffraction diagrams; it was
found that the d-spacings for the P(3HB) homopolymer
and the P(3HB-co-8 mol% 3HV) copolymer were the same,
implying that (R)-3HV units are excluded from the crys-
tals and exist mainly on the crystal surfaces during slow
crystallization, with a small proportion acting as a defect
group in the crystal. In this case, the ethyl side chains of
the HV component cause the (1 1 0) plane of the HB crystal
to expand due to steric effects [172,192]. These results are
confirmed by other studies. By solid state NMR, for exam-
ple, it has been shown that at low HV contents the crystal
almost exclusively comprised the HB crystal unit with HV
effectively excluded, and the less ordered regions being
enriched in HV, and at higher levels of 3HV content the
3HV crystal structure dominated. A plot of mol% 3HV in the
Fig. 5. Effect of comonomer content in PHA on melting temperature, heat crystal and amorphous phase was produced (Fig. 6). At 47%
of fusion and glass transition temperature. Plot shows (a) melting temper- HV, the less ordered portion of HV increases on stretching
ature (Tm ), (b) heat of fusion (Hm ) and, (c) glass transition temperature while that for HB disappears. On annealing this behavior
(Tg ) versus content (mol%) of second monomer unit (3HV, 3HP, 3HHx and
is reversed [153,190]. This work was confirmed by other
4HB) for well fractionated copolyesters; 䊉 P(3HB); : P(3HB-co-3HV); :
P(3HB-co-3HP); : P(3HB-co-3HHx); : P(3HB-co-4HB). NMR studies using very carefully fractionated copolymers,
Reprinted (adapted) with permission from Ishida et al. [196], Copyright which showed that the HB units are easier to crystallize in
(2005) International Union of Pure and Applied Chemistry. the HV unit than vice versa [205]. The 3HV units are mostly
excluded from the crystalline region if the HV content is
<20% [205,206].
A minimum in the melting temperatures and in the Density measurements also support this proposal. The
enthalpy of melting at around 30–55% HV has also been density of the amorphous PHA is 1.177 g cm−3 and of the
consistently demonstrated [165,190,193–200] with more crystalline phase is 1.260 g cm−3 [160]. Barker et al. [207]
recent literature proposing that it is actually a narrow win- showed that the density of P(3HB-co-3HV) in both the
dow around 47–52% HV once the copolymers are carefully crystalline and amorphous states decreased as HV content
solvent fractionated [201,202] (Fig. 5). increased. Assuming that the degree of crystallinity of the
Melting point depression in polymer mixtures is usu- copolymers of different HV content is similar, then there
ally explained by the Flory equation, but for P(3HB-co-3HV) must be partial exclusion of HV units from the crystals
this is not applicable since it was derived based on the [207]. Mitomo and Doi [195] isolated the crystalline core
assumption that copolymer crystals are composed of only using aminolysis (with 20% methylamine), and showed
one type of comonomer component and the other is in that the HV content of the core was smaller than the
408 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

co-crystallization therefore decreases with an increase in


crystallization temperature because more arrangement of
the chain can take place during crystallization at the higher
temperature, providing the opportunity to approach the
equilibrium state. The temperature dependence of the
crystal structure and the weak intermolecular interactions
in P(3HB-co-3HV) has also been analyzed using FT-IR [213].
It was found that for HB-rich copolymers, the CH3 · · ·O C
hydrogen bond between the CH3 group and the C O
group is found only in the HB part of the crystal structure,
while the high HV content materials bond through the
CH2 · · ·O C groups. For P(3HB-co-3HV) up to 21 mol% HV,
there was a sudden drop in crystallinity above 170 ◦ C. By
contrast, for 28.8 mol% HV, the change began much earlier
at around 70 ◦ C. The P(3HV) homopolymer lost crystallinity
earlier and at lower temperatures than P(3HB), and at
Fig. 6. Composition dependence of HV content of P(3HB-co-3HV) in the the eutectic (HV = 58.4 mol%) crystallinity began to be
crystalline phase; 䊉 HV content of crystalline phase;  HV content of lost even at room temperature. WAXD spectra were also
amorphous phase. followed with increasing temperature and showed similar
Reprinted with permission from Bonthrone et al. [154], Copyright (1992)
Wiley-Blackwell. results. It was proposed that the lamellar structure in
these copolymers deformed with increasing temperature,
for steric reasons, and that these weak hydrogen bonds
overall content in P(3HB-co-3HV) copolymer, indicating are very important for the stabilization of the lamellae. In
some exclusion of HV. addition it has been found that there is a rigid component
By using compositionally fractionated and labeled in the non-crystalline region of P(3HB) and P(3HB-co-3HV)
copolymers, Yoshie et al. [208] also showed there is a struc- at room temperature, which decreases with increasing
tural transition in the crystal state at around 10 mol% HV. It temperature and HV content [214]. This may be due to the
was confirmed that above 10 mol% HV the d(1 1 0) spacing tightly packed structures adjacent to the lamellae.
increased with an increase in HV, and a small increase in When the 3HV comonomers are included in the P(3HB)-
d(0 2 0) was also observed. The d(0 0 2) spacing remained type lattice, they will result in “less-perfect” crystals with
constant, confirming the expansion in the ab plane. Below more defects, smaller crystalline domains and less brit-
10 mol% HV, the crystal was the same as for P(3HB). Two tleness [215]. At higher HV contents the spherulites are
models for P(3HB-co-3HV) crystals were introduced: (1) small as the rate of crystallization is slower and/or there are
the sandwich model, which applies below 10 mol% HV: in relatively more nucleation sites. Less perfection in the crys-
this case the core of the lamella is pure HB units, and the talline domain in turn corresponds to films with greater
HV units are at the crystal edges and in the amorphous flexibility and toughness [215].
regions; and (2) the uniform lamella model in which the By contrast, none of the other comonomer units found
HV units are uniformly distributed through the crystal. It in PHA have demonstrated any degree of co-crystallization.
was of note that crystallization rates dropped rapidly above Mitomo and Doi [195] found that there was no change in
10 mol% HV as well. lattice parameters in P(3HB-co-4HB) with increasing 4HB
From Molecular Mechanics (MM2) modeling [191], the content, implying that there was no inclusion of 4HB in
incorporation of the 3HV unit into the P(3HB) type lattice the 3HB crystal, and Iawata et al. [204] also demonstrated
has a substantial influence. This is attributed to the steric that the second monomer units for P(3HB-co-6HHx) and
interaction between the ethyl side group and the nearest P(3HB-co-4HB) copolymers are largely found on the crys-
carboxyl oxygen atom. However, it was also found that tal surfaces (and also that these surfaces are irregular with
replacement of a 3HB unit by a 3HV unit only affected the loose loop chain foldings). Xie et al. [216] likewise found
conformational angles of one of the nearest neighboring HB that HHx is excluded from the HB crystal (at 7.2% HHx),
units; therefore the overall 21 helical structure remained and high-resolution solid-state 13 C NMR of composition-
unchanged. The effect of crystallization temperature on ally fractionated bacterial P(3HB-co-3HP) [217] showed
the extent of co-crystallization has also been investigated that co-crystallization did not occur in the P(3HB-co-3HP)
[156,209,210]. The 3HV content in the P(3HB) lattice has system. From the lineshape analyzes of the 3HB methyl car-
been shown to decrease as the crystallization temperature bon peak it was also shown that the presence of minor 3HP
increases; likewise at higher HV contents when the P(3HV) comonomer units significantly suppressed the crystalliza-
lattice is present, the 3HB content decreases with tem- tion of a 3HB-rich semicrystalline copolyester; there was a
perature. Barham and coworkers [182] pointed out that correlation between HB block length and the crystallizabil-
the mechanism for inclusion of comonomer units into the ity of the copolymers.
P(3HB) crystal unit is by kinetic (i.e. dependent on a combi- The glass transition temperature does not show pseu-
nation of time and temperature), rather than equilibrium, doeutectic behavior, as it is reflective of molecular motion
methods. According to Yoshie et al. [212] co-crystallization in the non-crystalline phase; instead it shows a linear rela-
results from the entrapment of P(3HB-co-3HV) chains tionship, decreasing with an increase in comonomer unit
near the growing front of the crystals. The extent of content [197] (Fig. 5). The Tg of P(3HA)s with long side
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 409

chains is about −40 ◦ C, which decreases with an increase the extreme cases of complete inclusion or exclusion of the
in the average side chain length [218]. comonomers from the crystals” [211] – see also [207].
Melting temperatures and the enthalpy of melting are
known to be dependent on the thickness of the crystalline
5. Notes on the analysis of crystallinity in PHA lamellae. This thickness varies with crystallization condi-
copolymers tions. Therefore the melting temperature as detected in
the DSC depends on crystallization conditions such as crys-
With respect to the previous discussion on composition, tallization temperature, crystallization time and annealing
crystallinity and crystal structure, there are notes of caution treatment.
to be made. The analysis of crystallinity is not straightfor- One of the standard methods for determining the crys-
ward for PHA copolymers. tallinity (X) of semi-crystalline polymers is by comparing
Crystallinity can be expressed in three ways depend- the area of the melting peak (Hf ) with the melting
ing on the analytical approach, and can include the volume enthalpy of a 100% crystalline material (Hf 100% ):
fraction of crystal, the mass fraction of crystal, or the molar
fraction of crystal [207]. WAXD methods of analysis mea- Hf
X=
sure the intermolecular order as a result of chain packing Hf 100%
(effectively the mass fraction crystallinity) and are the
method of choice for determining relative crystallinity of An estimate of the heat of fusion of an infinite crystal
these materials, particularly if the Ruland method (which of P(3HB) was obtained using a slowly annealed film of
requires diffraction data over a wide angular range and 86% crystallinity and was found to be 146 J g−1 by den-
correction for air scattering, incoherent scattering, and sity measurement [160], which is the usual value used in
diffuse scattering arising from thermal fluctuations and this calculation. Bauer and Owen [221] obtained a higher
paracrystalline disorder) is used [218]. A number of other value of 170 J g−1 for P(3HB). Steinbüchel and Schmack
techniques have also been used over the years, and most [222] found that P(3HV) had a Tm of 108 ◦ C with H
give results that are consistent with XRD results. For of around 91 J g−1 . Pearce and Marchessault [223] used
example, by 13 C NMR analysis using single-pulse excita- Hf 100% for 95 mol% HV of 128 J g−1 (from [165]) and for
tion with magic-angle spinning, the crystallinity of P(3HB) 100 mol% P(3HV), 131 J g−1 (from [224]). For other PHBV
and P(3HB-co-3HV) (2.7 and 6.5 mol% 3HV) was found to copolymers of lower HV content, a common value used for
be 68%, 60% and 56% respectively; this compared well Hf 100% is 109 J g−1 (obtained from WAXS and DSC analy-
with WAXD data of 60, 55, and 53% [219]. Similarly, in sis of PHBV containing 10 mol% 3HV) [225]. However, some
another NMR-based study it was found that the degree authors have used slightly modified techniques to calcu-
of crystallinity was 60%, 60%, 58% and 56% for P(3HB) and late percent crystallinity in these copolymers. Cheng and
P(3HB-co-3HV) (at 4.33, 6.96 and 10.5 mol% 3HV respec- Sun [226] took Hf 100% as 146 * mol fraction HB J g−1 ; this
tively) [214]. Kansiz et al. [220] found by FT-IR that was used since at low HV content (<10%) the sequences of
P(3HB-co-3HV) (14 mol% HV) had 47% crystallinity at full HV units are excluded from the lamella. By contrast Chan
development of crystallinity (which was slightly lower et al. [227], in blends of P(3HB-co-3HV) (12 mol% 3HV), took
than found by the other techniques but still in line) – this H = HPHB * wt% PHB + HPHBV * (1 − wt% PHB), where
method measures the molar fraction of crystals and takes HPHB = 146 J g−1 and HPHBV = 109 J g−1 . An alternative
into account short-range order (such as in tie molecules) measure for Hf 100% for P(3HB-co-3HV) = 3427 cal mol−1
[207]. (or 14.3 kJ mol−1 ) of crystalline repeat unit according to
Differential scanning calorimetry (DSC) is an extremely Orts et al. [228]. (Values given in kJ mol−1 are not con-
useful tool for the characterization of PHA materials, par- verted into J g−1 given the variation in molecular weight
ticularly in light of the significance of rate of crystallization depending on repeat unit composition.) Kamiya et al. [205]
with respect to polymer properties and also of the impact used Hf 100% for P(3HB) = 11.2 kJ mol−1 (Tm = 175 ◦ C) while
of melt temperature on polymer processing, thermal sta- Hf 100% for P(3HV) = 10 kJ mol−1 (Tm = 108 ◦ C). Specific val-
bility, etc. The significant features of the trace include the ues of H◦ for a range of P(3HB-co-3HV) copolymers were
melt temperature (Tm ) and the heat of fusion (Hf ), the also obtained using a combination of heat of fusion from
crystallization temperature (Tc ), the cold crystallization DSC and crystallinity from XRD [165], with values ranging
temperature (Tcc ), and the glass transition temperature from a maximum of 130 J g−1 for P(3HB) to a minimum of
(Tg ). However, the use of DSC for the measurement of crys- 92 J g−1 for 34 mol% 3HV. Most of these methods bring the
tallinity in PHA copolymers based on heat of fusion needs to crystallinities as assessed by DSC more in line with XRD
be approached with much care. According to Barham et al. results; however, the enthalpy of fusion is often still lower
[211], “the measurement and definition of crystallinity is than expected (see Fig. 5 for the enthalpy of fusion of frac-
not a simple matter even in homopolymers; in copolymers tionated samples, and data in [203]).
it becomes very complex”. While there is adequate agree- The major uncertainty of methods such as DSC is that
ment between the results obtained for different methods some knowledge of the values of a given property (e.g.,
such as XRD and DSC for the P(3HB) homopolymer, in PHBV density or heat of fusion) of the 100% crystalline polymer
copolymers, “it is quite difficult to measure a true crys- must be assumed. Furthermore it is usually assumed that
tallinity, since the density and heat of fusion measurements this property varies linearly with both the degree of crys-
are affected by the partitioning of the HV units between the tallinity and the copolymer composition. This is invalid
crystalline and amorphous regions and can only be used for where co-crystallization occurs [228]. “The borderline of
410 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

copolymer composition between isomorphism and lattice fraction, (5) different molecular weight species [234]
defects is ambiguous.” [229]. and (6) orientation effects, and so on [234–236].
Orts et al. [228] compared the crystallinity obtained Melt/recrystallization melting peaks are typically due
using inverse gas chromatography with the results to the primary crystals formed in crystallization process
obtained from XRD and DSC. In this method, a column was (the lower peak) with higher melting peaks formed due to
packed with diatomaceous silica coated with P(3HB-co- crystals that have recrystallized during heating.
3HV) (7, 21 and 27% HV), and the retention of a “probe” As a further note, it has been demonstrated that the rate
chemical (decane) was tested as a function of tempera- of crystallization of pure P(3HB-co-3HV) copolymers of 18
ture. Crystallinities of around 60% for all copolymers were and 30 mol% HV is extraordinarily slow and can take weeks
observed, which was consistent with XRD results. By con- to fully develop [172]. In any DSC studies of blends, not pure
trast, the crystallinities determined from the heat of fusion random copolymers, it is important to take into account the
showed a large drop from 59% to 24%. It was found that the fact that preferential crystallization of the faster crystalliz-
decrease in the heat of fusion with increasing HV content ing phase will exclude the separate, slower crystallizing
was not describing a decrease in crystallinity but instead components into a separate phase where, with time, these
was reflective of a decrease in crystallite size and/or crys- components can also slowly crystallize. Without allowing
tal perfection. It was also noted that as the HV content sufficient time (or the best temperature for annealing of the
increases, melting begins at a lower temperature, getting materials) the DSC analysis will not be reflective of the final
increasingly close to room temp. This is likely due to sam- polymer crystallinity. This is also true for pure copolymers
ples of lower HV content crystallizing more readily into of high HV content.
highly ordered crystallites with a relatively large crystal The increased segmental mobility in the amorphous
size. It is also likely that the inclusion of this “other” co- phase for P(3HB-co-3HV) is responsible for the observed
unit in the P(3HB) crystalline phase systematically affects decrease in Tg [196]. However, there is a very low enthalpy
the heat of fusion due to defect formation. for the glass transition of P(3HB) of 0.43 J ◦ C−1 g−1 as
Chen and Hwang [230] have shown that crystallinity reported in Scandola et al. [237], so as the percentage
obtained by Hf /Hf 100% can be called the ‘equivalent of the P(3HB) crystal phase increases, so the Tg for this
weight crystallinity’, signifying “the amount of perfect phase becomes increasingly difficult to observe, and DMTA
crystals that can be melted by the measured enthalpy of becomes a more reliable method of analysis for this param-
melting. The equivalent weight crystallinity does not only eter.
depend on the amount of the crystals, but also on the per-
fection of the crystals: a crystal with higher perfection has 6. Crystallization kinetics
larger equivalent weight” [230]. Samples of P(3HB-co-3HV)
crystals (of 17 and 30 mol% 3HV) grown isothermally from The crystallization kinetics of PHA polymers are very
solution have been shown to be very small and irregular important to understand, as one of the potentially limiting
[172]. The fold surface also became more irregular with factors for the commercial use of these materials is the
an increase in HV content, which was attributed to inho- slow rate of crystallization for some of the copolymers.
mogeneous distribution of HV units in the crystalline and There are three types of crystallization behaviors (regimes
amorphous phases. It is also known that lamellar thick- I–III) according to the Hoffman theory of crystallization
ening with increased melting temperatures and spherulite [238]. In regime I (at low supercoolings i.e. at a temper-
size has been observed when PHA copolymers are crystal- ature that is not much below the melting temperature),
lized at higher temperatures [162,195,231]. the radial growth rate is large relative to the nucleation
Poley et al. [232] in an alternative explanation proposed rate: growth following each nucleation event is rapid and
that the difference may be due to a decrease in the amor- complete before the next nucleation event. In regime II,
phous phase density. In addition, the presence of air in at lower temperatures, the nucleation rate and the radial
the free volume of the polymers may be a consideration; growth rate are comparable, so that both the formation of
given the results of oxygen and carbon dioxide mass dif- new nuclei and the growth of crystals take place together.
fusion, they assumed that the air solubility of the samples In regime III (at high supercoolings), the nucleation rate is
grows with increasing HV. As air-specific heat capacity is very large in comparison with the growth rate so that there
1.17 × 10−3 J cm−3 K−1 , three orders of magnitude lower is little distance between the nuclei.
than the values in the typical range of polymers, there-
fore it was suggested that air could under these conditions 6.1. Equilibrium melting temperature
contribute to a stronger reduction in heat capacity values.
It is important to distinguish multiple peaks that are The equilibrium melting temperature, Tm 0 , is defined as

due to phase-separated structures from those that are a “the melting temperature of an infinite stack of extended
result of annealing during heating in the DSC. This can be chain crystals, large in directions perpendicular to the chain
done by changing the rate of heating [233]. The possibil- axis and where the chain ends have established an equi-
ities contributing to double or multiple melting behavior librium state of pairing” [239]. This parameter is “one of
include: (1) melting, recrystallization, and remelting the most important thermodynamic properties of crystal-
during heating, (2) the presence of more than one crystal lizable polymers, as it is the reference temperature from
modifications (polymorphism), (3) different morphologies which the driving force for crystallization is defined” [239].
(lamellar thickness, distribution, perfection or stability), (4) According to the Hoffman–Weeks method, the equilib-
physical aging or/and relaxation of the rigid amorphous rium melting temperature Tm 0 is obtained by extrapolating
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 411

a linear regression of the observed melting temperature


Tm of crystals formed at a range of crystallization tem-
0 values have
peratures (Tc ) to the line Tm = Tc . Various Tm
been calculated for P(3HB) and its copolymers. Tm 0 was

estimated to be 197 ◦ C for P(3HB) and 143 ◦ C for P(3HB-co-


3HV) (30 mol% 3HV) by Mitomo et al. [172], and was found
to be 195 ± 5 ◦ C as judged by one method, 200 ◦ C by another
and 197 ◦ C by a third according to Barham et al. [160]. Organ
[240] obtained a range of 176 ≤ Tm 0 ≤ 188 ◦ C for P(3HB-

co-3HV) (4 mol% 3HV), compared with 179 ≤ Tm 0 ≤ 198 ◦ C

for P(3HB), depending on molecular weight. Other authors


also report values within this range [227]. The equilib-
rium melting temperature for P(3HB) was found to have
a strong and unexpected dependence on molecular weight
[240,241]. For 8 mol% HV, Tm 0 decreases as molecular weight

decreases (from 188 ◦ C to 176 ◦ C as molecular weight falls


from 220,000 to 56,000). The behavior is more complex
Fig. 7. Variation of crystallisation rate for a range of melt-quenched
for the 18 mol% HV material, suggesting that the degree of
P(3HB-co-3HV) copolymers as followed by XRD. 䊉 0% HV;  4% HV; 
exclusion of HV from the crystal lattice varies with molecu- 7% HV;  8% HV;  17% HV;  20% HV.
lar weight. There was a hypothesis that there is a different, Reprinted (adapted) with permission from Bloembergen et al. [243], Copy-
higher melting crystal structure in polymers above 200,000 right (1986) American Chemical Society.

molecular weight; however, this was not supported by XRD


[240,241]. similar result was obtained by Kunioka et al. [192]. This
The Tm0 of P(3HV) was found to be 130 ◦ C, with a T slow rate was attributed to the increasing difficulty for
g
value of −16 ◦ C (suggesting that the previously reported either crystal lattice to accommodate increasing amounts
−11 ◦ C was due to the partial crystallization of P(3HV) on of guest monomer units during cocrystallization [165]; the
supercooling) [223]. (R)-HV units have to be regarded as energetically unfavor-
able defects in the P(3HB) crystalline phase and vice versa.
6.2. PHA crystal growth kinetics It was found [243] that the crystallization rate of P(3HB-
co-3HV) copolymers from the melt was faster than that
The kinetics of the growth rate of the P(3HB) spherulite of solution-cast samples, and it was much faster still from
has been evaluated at various crystallization temperatures solution-precipitated samples (as analyzed using FT-IR and
and shows a complex curve against temperature and a XRD). The variation of composition in the crystalline phase
maximum value of about 3–4 ␮m s−1 at ∼90 ◦ C [160]. It with temperature is also more pronounced in the copoly-
has been reported that the regime of crystallization for mers with mid-level HV contents. Bloembergen et al. [243]
P(3HB-co-3HV) and its blends is assigned to be regime III noted that these materials have an induction period, and
(j = 4) when Tc is below 130 ◦ C [156], with the transition there was a suggestion that the initial formation of highly
from regime II to regime III also occurring at ∼130 ◦ C for ordered crystals was followed by the formation of much
P(3HB) [160]. Regime I growth has yet to be observed in less perfect crystals with lower melting temperatures. It
P(3HB) [9]. P(3HB) can undergo heterogeneous nucleation was found that the crystallite size as judged by the width
without addition of nucleating agents provided it has com- at half height of the crystalline peaks (2 ≈ 17◦ ) remained
pletely melted. The nucleation rate is obtained by counting constant for melt quenched films, but for solution cast sam-
the number of spherulites formed during the isothermal ples the initial thickness was higher but fell with time.
crystallization process. The maximum rate of nucleation This implies that crystallization is mostly nucleation rate
occurs at a lower temperature than that of crystal growth limited in the first phase, then crystal growth rate limited
rate for P(3HB), and in the pure polymer it is sporadic in the second.
[155,156,211]. Therefore the overall crystallization rate for It has been suggested that CH· · ·O hydrogen bonding
P(3HB) is at a maximum at between 60 and 90 ◦ C [242]. between the C O and CH3 groups in P(3HB) helps to
The spherulitic growth rates for P(3HB-co-3HV) copoly- drive crystal formation [244,245]. During crystallization,
mers are markedly reduced as the comonomer content the P(3HB) molecules become constrained in their move-
increases to the pseudoeutectic (at ∼50 mol% 3HV) ment due to close packing into a crystalline state. When the
[165,192,211,217]. Crystallization curves are shifted to crystals melt, the CH· · ·O hydrogen bonds weaken, increas-
lower temperature, and nucleation rates also decrease, ing molecular mobility.
with the overall rate being some four orders of magnitude In P(3HB-co-4HB), the rate of crystallization also
slower at the pseudoeutectic. Above the pseudoeutec- decreases with increasing comonomer content [192,246].
tic, the rate again increases; at 83 mol% 3HV it is about The crystallization of mcl-P(3HA)s with long side-chains
the same as for P(3HB) [165]. For example, Bloember- is very slow [247,248], taking days to weeks to reach full
gen et al. [243] followed melt quenched P(3HB-co-3HV) crystallinity, with an equilibrium crystallinity of about 25%,
(20 mol% HV) films by WAXD (Fig. 7), which was still which is relatively small compared with that of P(3HB-co-
amorphous after 15 min and took days to fully crystal- 3HV) [249]. Abe et al. [88] also looked at isotactic/atactic
lize. The ultimate crystallinity was around 65% [243]. A blends; again the radial growth rate decreased as isotactic
412 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

P(3HB) content increased. The SAXS data showed that the influences the end-use properties, it is important to under-
lamellar periodicity of P(3HB) increased also. The decrease stand the source of this variability, which may be a result
in crystallinity was fully accounted for by the amorphous of inherent uncertainties in the method but may also be a
addition of the atactic component, i.e. it did not affect result of variability in the properties of the materials being
the crystallinity of the crystalline component, implying analyzed, such as chemical compositional distribution and
that the atactic P(3HB) is found in the amorphous regions levels of impurities.
between the lamellae. The activation energy of crystallization was found to be
Detailed studies have shown that the high degree of 110 ± 14 kJ mol−1 for P(3HB-co-3HV) (8 mol% HV), larger
crystallization and relatively rapid crystallization rate of than that of P(3HB) [254], meaning that the crystallization
P(3HB) generates pores and protrusions on the P(3HB) ability of P(3HB-co-3HV) is lower than that of P(3HB). The
film surface, which could prohibit mammalian cell growth work of chain-folding was also lower than that for P(3HB),
[250]. The presence of P(3HB-co-3HHx) in P(3HB) polymer implying that the chains of P(3HB-co-3HV) were more flex-
strongly reduced both the degree of crystallization and the ible.
crystallization rate of P(3HB). The low degree of crystalliza- Non-isothermal crystallization kinetics can also be ana-
tion of P(3HB-co-3HHx)/P(3HB) blends provided films with lyzed by looking at the relationship between the relative
a fairly regular and smooth surface. degree of crystallinity (Xt ) as a function of temperature
at different cooling rates. Ziaee and Supaphol [258] have
6.3. Analysis of crystallization kinetics and mechanism reviewed the kinetics of P(3HB) crystallization under non-
isothermal conditions, and found that the Tobin and Ozawa
The Avrami equation describes the overall crystalliza- methods gave comparable results to the Avrami method
tion rate under isothermal conditions, assuming that the under these conditions.
relative degree of crystallization Xt is developed with crys-
tallization time t: 6.4. Effect of blends on crystallization rates
nA
1 − Xt = exp(−kA t ) (1)
Finally, the crystallization kinetics of PHA copolymers
which is the double-logarithmic form of can be strongly affected by the presence of blended mate-
rials. According to Kammer et al. [259], the sensitivity
ln[− ln(1 − Xt )] = ln kA + nA ln t (2)
of conventional DSC (Tg ) to heterogeneities in polymeric
where kA is the crystallization rate constant (min−1 ) materials is approximately 50 nm domain size, although
and is temperature dependent; nA , an integral, is the others say that it can provide information on miscibility
Avrami index, and contains information about nucleation and phase structure on the 10–30 nm scale [260,261]. The
and growth geometry. Although Avrami analysis can also crystallization kinetics can be monitored through the crys-
be used to describe the non-isothermal processes, the tallization exotherm, with the initial slope being related to
obtained values of kA and nA do not have the same physi- the nucleation rate. If this crystallization rate is depressed,
cal meaning as in isothermal processes due to the fact that then this is typically due to (1) a reduction of the driving
under nonisothermal conditions the temperature changes force for crystallization due to the changes in the equilib-
constantly. There have been a range of values obtained for rium melting point, (2) a dilution effect (i.e. a change in
the Avrami index, nA . For example, Chanprateep et al. [251] the composition of the not-yet-crystallized melt, which is
and Chua et al. [252] obtained values for P(3HB) of the order enriched in components that do not crystallize as rapidly
of 3, and that of P(3HB-co-3HV) was found to be between as those in the developing crystal), which lowers the
2.5 and 3 (for 2.6 mol% and 8 mol% HV) [253,254]. In general, probability of a critical nucleus forming in front of the
nA = 3 corresponds to two different possible crystallization growing spherulites and (3) entropic factors that produce
mechanisms. One is three-dimensional growth and instan- an increase in the energy related to the formation of a
taneous nucleation; the other is two-dimensional growth nucleus of a critical size [262].
and homogeneous nucleation [254]. By contrast, Gunaratne
and Schanks [234] obtained nA ∼ 2 for commercial P(3HB) 7. Molecular weight control
and P(3HB-co-3HV) (5–12 mol% HV) which corresponds to
2D crystal growth and heterogeneous nucleation. Liu et al. The molecular weight distribution is also, to a
[255] also obtained nA values of 2.0–2.2 for P(3HB-co-3HV) large extent, responsible for the end-use properties of
(6.6 mol% HV). Chan et al. [227], however, obtained nA val- biopolymers, via its control of macromolecular and supra-
ues across a range of HV contents of 2.35–2.7 and Saad et al. macromolecular structures. It is known that the molecular
[256] obtained an exponent of 3.8 (consistent with sporadic weight of PHA synthesized by biological means is much
3D spherulitic growth) for thin P(3HB) films. Meanwhile higher than that achieved chemically [12]. The mechani-
Xu et al. [257] applied the Avrami equation to IR data for cal properties of PHA deteriorate when the weight average
isothermal crystallization of P(3HB) and Nodax (P(3HB- molecular weight (Mw ) is lower than 0.4 × 106 Da [263],
co-3HHx)) and obtained an nA value of 1.72 and 2.08 and for thermoplastic applications the value of Mw should
respectively, indicating that a heterogeneous nucleation be higher than ∼0.6 × 106 Da [264]. Molecular weights typ-
mechanism exists in this case. There are other examples for ically range between 0.2 × 106 and 3 × 106 Da [39], and
blends, but there would seem to be considerable variability different bacteria are known to produce P(3HB) of different
in the isothermal crystallization kinetics for P(3HB/V) and molecular weights [71]. Substrate type and concentration,
similar materials. Since the crystallization kinetics strongly nutrient availability and growth conditions such as pH
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 413

and temperature are also important [265–267], and val- in an overall increase in the average molecular weight. It
ues of Mn as high as 2 × 107 Da have also been reported is believed to be the balance between these processes that
in mutant strains [134]. There is also a low molecular affects the length of the polymer chain [19].
weight form of P(3HB) (of ∼13,000 Da) which is commonly There have been two approaches to the control of
found in bacterial cells and is a constituent of the cyto- the molecular weight of PHA through polymer synthesis
plasm and cytoplasm membrane, where it may be part of a to date. The first involves control of the level and time
P(3HB)/calcium polyphosphate channel [268]. of expression of PHA synthase [282]. A lower molecular
Accurate determination of the true molecular weight weight is obtained when high levels of synthase expression
requires solution viscosity measurement. However, the are induced; likewise a higher molecular weight results
Mark–Houwink–Sakurada parameters (which relate the when low levels of synthase are expressed [282,283].
intrinsic viscosity to the molecular weight) have only However, the molecular weight of P(3HB) produced by
been reported for the P(3HB) homopolymer [269,270]. In recombinant R. eutropha was found to be independent of
general, the molecular weight data for PHA samples are PHA synthase activity [136]. The second method of control
obtained for comparison purposes by means of GPC anal- of PHA molar mass involves the addition of poly(ethylene
ysis, where polystyrene standards are used to construct a glycol) (PEG) to the fermentation medium (see Section 3.3
calibration curve. for a discussion of the effect of water or alcohol addition
A number of integrated metabolic/polymerization mod- on the molecular weight of PHA chains). PEG is a neu-
els have been developed for the prediction of P(3HB) tral, water-soluble polyether that is relatively nontoxic to
concentration and molecular weight distribution in bacte- cellular systems [88], and associates with membrane phos-
rial cultures [271–276]. Some models have assumed chain pholipid head groups [284,285].
transfer and polymer degradation reactions to be neg- In other studies, it has been shown that a potassium
ligible, which does not hold true in wild-type bacteria deficiency can lead to the biosynthesis of a high Mw P(3HB)
[148]. More recent approaches have attempted to integrate (>3 × 106 Da) in a methane-utilizing mixed culture [286]. It
current metabolic models containing a realistic descrip- has also been shown that as the carbon source concentra-
tion of cell metabolism (i.e. monomer concentration is not tion increases, so the average Mw decreases [134,145,287].
assumed to be constant) with a polymerization kinetic In summary, since molecular weight is one of the pri-
model comprising initiation, propagation, chain trans- mary drivers controlling end-use properties, it is evident
fer and degradation reactions [124,131,135,274,276,277]. that the mechanisms controlling molecular weight need
Most models also assume that the concentrations of to be well-defined and tools for modeling and/or control-
polymerase, chain transfer and depolymerase molecules ling molecular weight need to be further developed and
remain constant during the polymer synthesis, and that the integrated into PHA production processes.
selectivity of the various P(3HA) polymerases for different
monomers does not change [148]. 8. Control of monomer distribution and
It has been calculated [132] that approximately two microstructure
molecules of (R)-3-hydroxybutyryl-CoA are added every 1 s
into a propagating chain of P(3HB), and that the average Another approach to the modification of chemome-
lifetime of a growing chain should therefore be about 1 h chanical properties of PHA copolymers is to manipulate the
at 30 ◦ C. It has also been shown that the Mn of P(3HB) poly- sequences of monomer units within the polymer chain to
mers increases rapidly at first then can decrease with time produce non-random distributions. This will in turn influ-
[100,278]. Shimizu et al. [279] likewise found that in Alca- ence the spherulitic microstructure of the polymer. Block
ligenes eutrophus H16 (ATCC 17699) fed with butyric acid, copolymers are of particular interest as microphase sep-
the maximum average molecular weight of P(3HB) was aration may lead to periodic nanostructures of differing
already reached within 5 h and was found to be lower when morphology, with unique mechanical properties as a result.
the butyric acid concentration was increased. Hiraishi et al. It should be noted however that in biological systems
[280] used AFM to study granule formation in vitro on the synthesis of a block copolymer is not straightforward.
graphite; in the initial stages, flexible fibrillar P(3HB) chains The kinetics of chain synthesis are fast (see Section 7 and
(0.3 ␮m diameter) were shown to be growing from the below), and the intracellular and/or intra-granular concen-
immobilized synthase particles, and 150–400 nm globules trations of free acids are not necessarily the same as in the
were formed within 5 min, which then coalesced to form bulk reactor. Likewise, the enzyme packages present within
large granules. the bacterial cells will differ depending on bacterial type,
The polydispersity (Mw /Mn ) meanwhile increases (typi- culture/feeding history and so on. The high degree of diver-
cally from 1.5 to 2.0) during accumulation [100,278]. These sity found in biological systems suggests a widely varying
results could be due to either chain transfer reactions or population of polymer chains may be synthesized within
depolymerase activity during granule formation or to the the bacteria [288]. The final product could well be a com-
random decay of active synthase molecules [278,281]. plex mixture of random copolymers and/or homopolymers
The cyclic nature of P(3HB) metabolism can also be used blended with copolymeric materials of a blocky character,
to manipulate molecular weight – for example Shimizu and care is needed in characterizing the final products.
et al. [279] found that, in the presence of a nitrogen Core–shell structures of PHA have been prepared by
source, P(3HB) production and degradation in Alcaligenes a number of authors. Curley et al. [289] showed that
eutrophus took place simultaneously, with the lower a core–shell structure resulted when P. oleovorans was
molecular weight P(3HB) being degraded fastest resulting grown on a mixture of nonanoic acid and 5-phenylvaleric
414 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

acid, with poly(3-hydroxynonanoate) in the core and n-pentanol concentrations. In a subsequent study [296],
poly(3-hydroxy-5-phenylvalerate) around this core. From both the carbon to nitrogen ratio and the feeding strategy
SDS-PAGE it was shown that both polymers resulted from were controlled. The crystallization mechanisms for the as-
the same enzyme system. A random copolymer could obtained “blocky” P(3HB-co-3HV) were similar to that for
not be obtained. Kelley and Srienc [290] also produced P(3HB), explained as being due to the P(3HB) crystal com-
core–shell polymer structures in Ralstonia eutropha by ponent dominating. The rate of non-isothermal crystalliza-
growing the bacteria on fructose and limited quantities tion of this blocky P(3HB-co-3HV) was found to be faster
(1 g L−1 ) of valeric acid; RuO4 staining was used to see the than that of random P(3HB-co-3HV) but slower than that of
structures (with P(3HB-co-3HV) copolymer taking up more P(3HB). Both normal and blocky P(3HB-co-3HV) materials
stain than P(3HB), being less crystalline). Yoon and Choi were fractionated; the former had D values (by NMR) typi-
[291] studied intracellular degradation rates for PHA inclu- cal of random copolymers (see Section 9.1) while the three
sions in Hydrogenophaga pseudoflava cells, where a range fractions from the latter had values compatible with P(3HB-
of structures varying from homopolymer blends to copoly- co-3HV)s rich in long blocks of 3HB units. The nominal block
mers to single homopolymer were formed. It was noted copolymer was shown to have slower biodegradation rates.
that a blend type structure (or core–shell or similar) was Pederson, McChalicher and Srienc [297] in a subsequent
formed with sequential feeding of glucose for 22 h followed study used off-gas mass spectroscopy to control valerate
by butyrolactone (for P(3HB) and P(4HB) blocks), or by additions for PHA block copolymer synthesis using valeric
feeding glucose first and then valerolactone (for P(3HB) and acid and fructose addition rates based on oxygen uptake
P(3HV) blocks). Pederson and Srienc [292] also produced rate and carbon dioxide evolution rate. Alternating feeding
a core–shell granule structure, with either P(3HB-co-3HV) times of 8, 10, 14 and 32 min gave in theory 21/22, 26/23,
core and P(3HB) shell (produced using either a 4 h or 7 h 36/24 and 48/10% di/triblocks respectively. The production
valerate feed followed by fructose feeding) or P(3HB) core of triblock copolymer with equal HB/HV content was tar-
with P(3HB-co-3HV) shell (produced with either a 5 h or geted using 12.5 min of mixed valerate and fructose feed
10 h feed of fructose followed by valerate feeding). Stain- alternating with 25 min of fructose. The block copolymer
ing demonstrated that two phase structures were achieved. approach was said to be complicated by the fact that syn-
This work was backed by model predictions. It was noted thesis does not occur simultaneously on all polymer chains
by these authors that block copolymer structures could be because of random chain initiation and termination events.
achieved if feeding pulses are on the timescale of individual About 30% of the total polymer sample revealed melting
polymer chain synthesis, which was estimated to be 1 h or properties suggestive of block copolymers.
less. Mumtaz et al. [121] used TEM to examine P(3HB-co-
Madden et al. [293] examined the kinetics of PHA 3HV) inclusions in wild type Comamonas sp. EB172, a local
synthesis: PHA synthase was said to catalyze 6.7 propa- isolate, grown on mixed organic acids obtained from palm
gation reactions per second. As a result they calculated oil mill wastewater. Core–shell structures were evident
that it should take 0.3 h (18 min) to complete a 0.7 × 106 Da through RuO4 staining; the authors proposed the forma-
chain. Therefore to get a block copolymer, it was proposed tion of a block copolymeric material, although the DSC data
that reagents should be changed every 5 min. Mantzaris was also consistent with a blend material and no NMR data
et al. [294] developed kinetic models and a theoretical was presented.
underpinning of synthetic conditions for the formation of Li et al. [298] used the PHA synthesis genes phaPCJAC
block copolymers. In terms of P(3HB-co-3HV)-P(3HB) and cloned from Aeromonas caviae transformed into Pseu-
P(3HB-co-3HV)-P(3HB)-P(3HB-co-3HV) di- and tri-block domonas putida KTOY06C (phaPCJA,c ) to produce a
copolymer formation, the optimum conditions were pre- material consisting of P(3HB) as one theoretical block
dicted to require 39 carbon source switches (to give 50% (nominally at 70 mol% of the total polymer) and the ran-
of total polymer as a block copolymer). This translated to dom copolymer P(3HV-co-3HHp) as another block (at
almost 30 min for each P(3HB-co-3HV) producing stage and 30 mol%), through sequential substrate feeding. Manu-
a little more than 1 h for each P(3HB) producing stage. Kel- factured blends as well as random copolymers and the
ley and Srienc [295] expanded the theoretical concept that homopolymers were compared with the blocky material
either core–shell or block copolymeric structures for PHA using NMR, DSC, and cone plate rheology, and the latter
could be produced by controlling timing of feed sequences. was found to have distinctly different characteristics.
Chanprateep et al. [251] synthesized possible block McChalicher and Srienc [288] also prepared block
copolymers of HV and HB (with 7 and 10 mol% HV) by copolymer (P(3HB)-block-P(3HB-co-3HV)) using C. necator
switching ethanol and pentanol as carbon sources; samples and compared it with random copolymers of P(3HB-
were fractionated using chloroform/heptane. The kinetics co-3HV). 30 min of fructose/valeric acid feeding was
of nonisothermal crystallization was studied using DSC and alternated with 1 h of fructose-only feed to give 50/50
analyzed using Ozawa’s equation via three different meth- block copolymers. Likewise, Pereira et al. [299] reported
ods; the crystallization rate of the nominal block polymers the probable formation of a P(3HB)-block-P(3HB-co-3HV)
was found to be faster than that of the random copoly- when R. eutropha was fed fructose in the first and propionic
mers. The n values from the Avrami equation for P(3HB) acid in the second stage using “alternating feeding” – the
homopolymer and for the block copolymer were both 1, timing and sequence of this was not described. The prod-
in contrast to the random copolymer (n ∼ 3). A model pre- uct, even after fractionation, did not appear to be a blend
dictive controller (MPC) coupled with a metabolic reaction of P(3HB) homopolymer and P(3HB-co-3HV) copolymer
model controller was used for controlling ethanol and based on the NMR, crystallinity and DSC data (see Section
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 415

9.2.1). Finally Ivanova et al. [81] explored the use of alter-


nating feeding of acetate and propionate to an enriched
mixed culture, and found that 13 C NMR data indicative of
blocky copolymers resulted, although it was not possible
to clearly determine if blends of random copolymers were
formed instead.
In all of the work above, the structural and composi-
tional similarity between the P(3HB) and P(3HB-co-3HV)
blocks means that many quantitative assessment tech-
niques (such as X-ray scattering) are complicated to apply
[288]. Qualitative assessment techniques using a com-
bination of fractionation, calorimetry, rheological, and
nuclear magnetic resonance techniques have been the
tools used to date [297]. Additional assessment tech-
niques may be of benefit. High-resolution solid state NMR
(via 13 C magic angle spinning with dipolar decoupling
and cross-polarization) is a useful technique for pro- Fig. 8. Effect of HV content on solvent fractionation of P(3HB-co-3HV)
using chloroform/hexane.
bing the molecular structure, conformation and dynamics Reprinted (adapted) from Wang et al. [201], Copyright (2001), with per-
of polymer chains at the interphase between phases in mission from American Chemical Society.
multicomponent polymer systems, and has been used to
distinguish PE–PP copolymer blends from block copoly- at the same composition fractionation based on molecu-
mers [25,26]. MALDI-TOF analysis of partially hydrolyzed lar weight can occur. The driving force for the separation
PHA could also be used to distinguish blocky from random of differing comonomer compositions in mixed solvents is
copolymers [407]. In addition, the molar Kerr constant, as believed to be the different polarities of the side chains,
determined in an electrical birefringence experiment, is a with molecular weight being a secondary factor; however,
macroscopic property of the polymer chain that is sensi- there is some inconsistency in this mechanism in that as
tive to local polymer microstructures, including tacticity the mol% 3HV increases above 70%, the solubility in hex-
and comonomer sequences [408]. Infrared and Raman ane decreases (Fig. 8). It has been proposed that this is due
spectroscopy also has the potential to provide useful infor- to the increased relative crystallizability of the materials of
mation, given that a characteristic distinguishing feature higher 3HV content.
of blocky copolymers is microphase separation which can The chemical compositional distribution (CCD) of
lead to more ordered molecular environments. These and P(3HB-co-3HV) is usually determined by 13 C NMR spec-
other characterization techniques could provide insight troscopy. Diad and triad sequence analysis is used to assess
into these as-produced materials. the extent of deviation of the copolymer composition from
The effect of these alternative polymerization strate- the statistically random (Bernoullian) compositional distri-
gies on the mechanical and rheological properties of PHA bution. The parameter D is given by:
is discussed in Section 11.3.
FBB FVV
D= (3)
9. Comonomer sequence compositional FBV FVB
distribution where FBB , FVV , FBV , and FVB are the fractions of the BB,
VV, BV and VB diad sequences, respectively [303]. The
Given that blends of P(3HB) and P(3HB-co-3HV) can D value for statistically random copolymers is 1.0. Non-
form different crystalline phases depending on the HV random copolymers will have D values greater or less
content of the copolymer, it is important to know the than 1, with “blocky” and alternating copolymers having
actual compositional distribution of microbial PHA. To date D and R values 1 and very close to 0, respectively. How-
the discussion of PHA polymer properties has been based ever, while polymers with a high D value could be true
on the presumption that the bacterial copolymer being block copolymers, they may also be a mixture of copoly-
analyzed is a random or tailored copolymer of narrow poly- mers with different compositions or of P(3HB) and P(3HV)
dispersity, which can be treated as a pure sample with homopolymers. Therefore D values much higher than 1
well-defined and reproducible properties. However, this clearly indicate bimodal (or multimodal) chemical compo-
assumption is not necessarily valid. sitional distribution but give little information on the width
of CCD. The parameter R, which is an indicator of the degree
9.1. Characterization of comonomer compositional of randomness obtained from analysis of the triad distri-
distribution bution, is more sensitive to the broadness of the chemical
compositional distribution. This parameter has a value of
Many as-produced P(3HB)-based copolymers have 1 for a completely random distribution of HB and HV units
been fractionated using a solvent/nonsolvent fractionation in the copolymer chain and 0 for a diblock copolymer.
technique to give a series of fractions with narrow composi- Yoshie and Inoue [302] noted that even if the same
tional distributions (as judged by 13 C NMR) [197,300–302]. author(s) synthesize two samples from the same carbon
Comonomer composition appears to be the determining source by the same organism under the same fermen-
characteristic by which fractions are separated, although tation conditions, one sample may be a homogeneous
416 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

random copolymer and another may be a mixture of ran- present as acetate in the feed increased. When propionate
dom copolymers. They concluded that the origin of the alone was used as a feed, only a random copolymer was
complex CCD in bacterial copolyesters is unclear, although formed, while other PHAs were found to be a mixture of
Tanadchangsaeng et al. [304] proposed that varying CCD random copolymers.
can result from a range of factors such as changing bacterial Lau et al. [311] prepared P(3HB-co-19 mol% 3H4MV)
strains, carbon substrates, pH in the fermentation medium, using a Burkholderia sp. (wild type) fed with fructose,
physiological state and so on during the cultivation. Even crude palm kernel oil, and sodium valerate along with 4-
materials with D values close to 1 (D = 0.994–1.5) could methylvaleric acid. A D value of 4 was obtained for this
be fractionated into several fractions [301]. It was shown polymer, indicating it was most likely a blend of copoly-
that samples with very large D values bear multimodal or mers.
bimodal CCD, while samples with D ≈ 1.5 had a broad CCD.
As an additional technique, some authors use a statistical 9.2. Solvent fractionation
approach looking at the relative fit of the data to a Bernoul-
lian model, a first-order Markovian model, a Newton model There have been many studies looking at in particular
and a mixture Markovian model [305–307]. the chemical but also the molecular weight fractionation of
Several papers have also reported on other meth- pure culture materials, and one examining mixed culture
ods for determining the comonomer composition of PHA. The thermal property and crystallinity data from the
P(3HB-co-3HV), including FFT-IR spectroscopy, electro- finely fractionated materials is more consistent than for as-
spray ionization multistage mass spectrometry (ESI-MSn ) produced materials and should by preference be used as a
[308] and fast atom bombardment mass spectroscopy reference for comparison (see Fig. 5 and references given
(FAB-MS) [309], etc. Much of the early research (prior to below).
1999) characterizing CCD by means of D values has been
summarized by Yoshie and Inoue [302]. It is beyond the 9.2.1. Solvent fractionation of P(3HB-co-3HV)
scope of this review to capture all of the subsequent NMR A summary of results for the fractionation of as-
studies; the following sections therefore firstly highlight produced P(3HB-co-3HV) across a broad range of initial
more recent studies of particular relevance where CCD compositions is given in Table 1. In most cases, a very broad
only is characterized, and then summarize the results from CCD was evidenced, with the as-produced material being
actual fractionation studies on as-produced PHA. shown to be a blend of random copolymers of differing 3HV
Some recent analyses of sequence distributions in content. It was consistently found that the polymers were
P(3HB-co-3HV) and related copolymers are of note. Žagar fractionated firstly according to 3HV unit content; how-
et al. [310] analyzed the monomer distribution in bacterial ever, when the 3HV-unit contents of fractions were similar,
P(3HB-co-3HV) containing 12.9 mol% HV (from Ralstonia then they were fractionated according to differences in
eutropha) and 12.0 and 27.1 mol% HV (produced using molecular weight. For example, Don et al. [312] produced
an osmophilic bacterium). Both 13 C NMR (diad and triad P(3HB-co-3HV) of 10.8 mol% HV content using Haloferax
analysis) and multistage electrospray ionization mass mediterranei. Fractionation using chloroform/acetone pro-
spectrometry (ESI-MS) were used to characterize the com- duced two compositionally different co-polymers. The
positional distribution, with the results being in good major fraction (at 93.4 wt%, with 10.7 mol% of 3HV) had a
agreement for all samples. The 12.0 mol% HV sample had high molecular weight of 5.7 × 105 Da. The second fraction
a much higher molecular weight and narrower polydisper- (ca. 12.3 mol% HV) had a much lower molecular weight of
sity than the other samples, with a sharp single melting 0.8 × 105 Da.
peak (Tm = 130 ◦ C) and narrow CCD. The other samples Four commercial samples were also shown to have
were mixtures of random copolymers of widely different broad compositional distribution in the bulk [313]
comonomer-unit composition. A 1:1 mixture of two P(3HB- (Table 1); for three of these materials, the melting temper-
co-3HV) copolyesters (containing 12.0 and 27.1 mol% HV ature and spherulitic growth rate corresponded to that of
respectively) was also prepared and had D and R values the average composition in spite of their complex composi-
that were between those of the constituent materials, with tion distribution. However, one as received P(3HB-co-3HV)
a single melting peak. with 21.8 mol% 3HV had a higher melting temperature and
Dai et al. [73] used glycogen accumulating orga- faster growth rate than might be expected from its overall
nisms (GAOs), found in mixed culture anaerobic–aerobic composition, which was attributed to the more rapid crys-
wastewater treatment processes, to produce PHA polymers tallization of the component chains of relatively low 3HV
consisting of 3HB, 3HV, 3H2MV and 3H2MB monomer units content.
with a non-HB fraction of up to 37%. The D value was found PHA produced using mixed cultures has also been frac-
to be between 2.4 and 4.7 for different runs and the prod- tionated. Inoue et al. [314] used an enriched mixed culture
uct was more likely to be a mixture of random copolymers sludge (prepared using a continuous anaerobic/aerobic cul-
(from the anaerobic production period) with an additional ture process) to accumulate PHA using propionate as the
segment of pure HB blocks or P(3HB) homopolymer (from sole carbon source. The resulting as-produced polymer
the aerobic production stage); it is possible that a poly- was composed of four types of monomer units (3HB, 3HV,
mer of blocky character was formed due to the sequential 3H2MB and 3H2MV) and could be fractionated into at least
uptake of reagents. In a related study [74], similar PHA four fractions of random copolymers of very different com-
copolymers were prepared by manipulating the ratio of position and molecular weights using water/acetone mixed
propionate to acetate: less 3H2MV and 3H2MB units were solvent.
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 417

Table 1
Solvent fractionation results for as-produced bacterial P(3HB-co-3HV).

Organism 3HV content of initial # of Range of 3HV Solvent system Character of resulting Ref.
material fractions content in fractions fractions
(mol% 3HV) (mol% 3HV)

46 4 23–72 Random copolymers by


Ralstonia eutropha Acetone/water [193]
64 6 45–85 NMR

24 9 21–31
Random copolymers by
Ralstonia eutropha 36 7 19–55 Acetone/water [194]
NMR
61 8 28–78

Burkholderia cepacia D1 36 5 10–81 Acetone/water Random copolymers by [203]


NMR

45 9 26–60
49 7 27–68
Assumed random
Ralstonia eutropha 70 5 30–74 Chloroform/n-heptane [201]
copolymers by DSC
80 12 54–94
96 8 74–98

45 10 26–60 Assumed random


Ralstonia eutropha Chloroform/n-heptane [202]
47 8 26–56 copolymers by DSC

Sourced from ICI 6.5 3 6.1–15.3


Sourced from ICI 15.3 5 10.9–25.3 Assumed random
Chloroform/n-heptane [313]
Sourced from Aldrich 19.4 5 15.7–24.9 copolymers by DSC
Sourced from Aldrich 21.8 7 10.2–34.4

The fractionation of nominally blocky or core–shell 9.2.2. Solvent fractionation of other PHAs
copolymers of P(3HB-co-3HV) has also been investigated. A number of studies have looked at the fractionation
Madden et al. [293] produced mixtures of homopolymer of other PHA copolymers (Table 2). In a series of investiga-
P(3HB) and random copolymer P(3HB-co-3HV) with Ral- tions, as-produced P(3HB-co-3HP) microbially synthesized
stonia eutropha by alternating long feeding periods of 5 h or by Alcaligenes latus [217,315–320] was found to have a
more of glucose and propionic acid, a result that was sup- broad CCD consisting of blends of copolymers. The HP con-
ported by solvent fractionation studies. This material had tent of the fractions decreased with increasing n-heptane
thermal properties that were markedly different from ran- content in the chloroform/n-heptane solvent mix. A min-
dom copolymers of similar monomer content yet exhibited imum melting temperature was found at 60 mol% HP
a single glass transition and a single melting peak that was content for the fractions produced; only the P(3HB) type
significantly higher than expected for a random copolymer. of lattice structure was found at low 3HP unit content,
The P(3HB-co-3HV) fraction that was produced had prop- with the 3HP lattice found at high 3HP contents and an
erties that were identical to those of the normal random amorphous polymer at intermediate 3HP contents. The
copolymer of the same HV content. presence of the minor 3HP comonomer units significantly
Pederson, McChalicher and Srienc (in their key paper on suppressed the crystallization of the 3HB-rich copolymers
block copolymer formation [297]) fractionated the blocky [317]. It was found that an average block length of more
P(3HB-co-3HV) material produced using a co-feeding time than two HB units was needed to form P(3HB)-type crys-
of 14 min, using a chloroform/heptane fractionation sys- tallites [217]. In a further extension [320], the effect of
tem. Analysis of the four resulting fractions showed a very pH was evaluated (Table 2). The more basic the fermen-
broad compositional distribution (from 0% to 44% 3HV). tation medium, the less 3HP units were introduced into
In addition, the third fraction (which was 30% of the total the copolyester, with the narrowest CCD being obtained
weight) had a D value of 9.1, compared to only 3.1 for F4 at pH 7.0. The distribution became broader with increas-
and 1.9 for F2, indicating a potentially blocky character for ing fermentation time. Manufactured blends of P(3HB) and
this fraction. Multiple melting peaks were present in the P(3HB-co-3HP) fractions have also been prepared: it was
DSC scan of this fraction indicating that complex blends of found that at 11.3 and 14.9 mol% 3HP, blends with P(3HB)
blocky and random copolymers were likely present. Žagar were miscible in the melt, with crystallization of the P(3HB)
and Kržan [303] also fractionated two P(3HB)-block-P(3HB- component dominating and intermolecular interactions
co-3HV) polymers according to molecular weight using between P(3HB) and P(3HB-co-3HP)s found to be present
preparative size exclusion chromatography. These materi- [318,319].
als were formed using alternating feeding in pure cultures. The solvent/non-solvent fractionation of P(3HB-co-
For the lower HV content (12.9 mol% HV) material there 3HHx) also gave evidence of a broad CCD for most of these
was almost no difference in D or R parameters for the dif- copolymers as produced (Table 2). Again, composition was
ferent molecular weight fractions, while for the 27 mol% HV the primary driver for fractionation, followed by molecu-
material there was a slight increase in D value (from 6.89 lar weight, and the thermal and crystallization behavior as
to 8.43) as the molecular weight decreased; however the well as the crystalline morphology was strongly affected by
R value remained relatively unchanged. Multiple melting the compositional distribution [321–324]. Asrar et al. [324]
peaks could be seen for these complex mixtures. found that P(3HB-co-3HHx) copolymers produced using
418
Table 2
Solvent fractionation results for as-produced bacterial 3HB-containing copolymers (PHAs) with comonomers other than 3HV (3HP, 3HHx, 4HB, 3H4MB).

Copolymer Organism Comonomer # of fractions Range of comonomer Solvent system Notes Ref.
content of initial content of fractions
material (mol% (mol% comonomer)
comonomer)

P(3HB-co-3HP) 26.4 4 14.8–54.2


Non-random
Alcaligenes latus 30.6 11 8.5–60.5 Chloroform/n-heptane [315]
copolymers or blends
62 6 41.3–85.3
36.5 10 13.2–73.6 Blend of random
Alcaligenes latus Chloroform/n-heptane [317]
68.1 12 31.4–95.9 copolymers
Alcaligenes latus 19.4 9 11.3–85.2 Chloroform/n-heptane [319]
A. latus pH 6 33.4 5 20.5–62.3

B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442


A. latus pH 6.5 23 7 15.1–31.4
A. latus pH 7 16.1 5 11.6–17.7 Chloroform/n-heptane [320]
A. latus pH 7.5 20.9 6 13.3–27.1
A. latus pH 8 13.2 4 11.7–37.5

P(3HB-co-3HHx) 7.5 12 5.6–12.9


11.2 21 8.4–18.1 Note refractionation of
Ralstonia eutropha Chloroform/n-heptane [321]
12.3 18 8.1–20.2 some fractions
13.7 21 8.1–21.5
7.4 6 4.4–10.6
Ralstonia eutropha 18 15 7.4–35.6 Chloroform/n-heptane Some fractions still [322]
a blend by DSC
22 13 14.8–41.9 Some fractions still
a blend by DSC
Aeromonas hydrophila 13.8 9 9.2–17.1
Ralstonia eutropha 35.2 2 29.8 and 69.8 Fractionated at
Chloroform/n-heptane [323]
70 ◦ C
Ralstonia eutropha 24.8 2 19.5 and 80.7 Fractionated at
70 ◦ C
Ralstonia eutropha 54 4 50.5–66.3
Aeromonas hydrophila 25.7 2 6.9 and 33.6 Chloroform/n-hexane Blend of random [324]
copolymers

P(3HB-co-4HB) Ralstonia eutropha 24 2 7 mol% and 86 mol% Acetone Blend of random [325]
copolymers
Ralstonia eutropha 28 3 1, 28 and 82 mol% Chloroform/acetone Blend of random [326]
then methanol/acetone copolymers
16.7 7 6.3–46.4
Ralstonia eutropha 50.9 10 7.1–95 Chloroform/n-hexane [327]
69.8 9 11–99.4

P(3HB-co-3H4MV) Ralstonia eutropha 7 4 2–39 Blend of random


(modified) Chloroform/n-hexane copolymers [304]
11 5 3–28 Blend of random
copolymers
35 3 20–47 Blend of random
copolymers
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 419

Aeromonas hydrophila had a Bernoullian (random) compo- enthalpy of fusion. Two different crystal structures were
sitional distribution with 3HHx content up to 9.5 mol%. By identified, which were temperature dependent.
contrast, copolymers produced using a recombinant strain Blends of P(3HB) and mcl-PHA rich PHA were also pro-
of Pseudomonas putida GPp104 had >9.5 mol% 3HHx con- duced when Pseudomonas species was fed fructose and
tent and a bimodal polymer composition and were found mannose, sodium glutamate, glucose, octanoic acid or oleic
to be blends of random copolymers. acid [88,139,330,331]; these were fractionated using boil-
Three studies have evaluated the fractionation of ing acetone to give in general an acetone-insoluble P(3HB)
P(3HB-co-4HB) copolymers. Doi et al. [325] found that fraction and an acetone-soluble fraction composed of mcl-
while many of the as-produced materials had D values PHA.
close to 1, indicating that they were random copolymers, Yu et al. [332] used chloroform/heptane fraction-
others had much higher D values (up to D = 66). These ation to fractionate novel sulfur-containing biopolymers,
were fractionated into acetone-soluble (with high 4HB poly[(3-hydroxybutyrate)-co-(3-mercaptopropionate)]s
content, ∼86 mol%) and acetone-insoluble (<7 mol% 4HB) [P(3HB-co-3MP)s]; a sample with 3MP unit content of
fractions, which each also had D values close to 1, indicating 16.3 mol% was fractionated into eight fractions with 3MP
the original polymers were bimodal mixtures of random unit content ranging from 10.3 to 37.2 mol%. Both the 3MP
copolymers that had very different compositions. Shi et al. and the 3HB crystal units were present in some fractions,
[326] used a similar approach but further fractionated the either because the material was naturally blocky in char-
acetone soluble material with methanol/acetone into solu- acter or because there were 3MP and 3HB rich phases.
ble and insoluble components. The three fractions resulting The earlier fractions were highly blocky in character (as
were found to have narrow compositional distribution and assessed by D value) while others were more random;
were composed of P(3HB), P(3HB-co-28 mol% 4HB), and however, there was no regular progression in terms of
P(3HB-co-82 mol% 4HB), respectively, again indicating the increase in mol% 3MP units with increasing heptane con-
broad compositional distribution in the as-produced poly- tent, nor was there a regular pattern in terms of molecular
mer. In a more detailed study, Ishida et al. [327] very weight distribution.
finely fractionated three as-produced P(3HB-co-4HB) sam- In conclusion, solvent fractionation has proved to be an
ples using chloroform/n-hexane mixed solvent. Again, they effective method for production of copolymer PHA with
found that the as-produced material had an extremely narrow CCD. Many as-produced copolymers have been
broad compositional distribution. In addition, the 4HB-rich shown to be composed of a broad range of compositions,
P(3HB-co-4HB) polymer fractions were immiscible with so the properties of the as-produced materials are often
the 3HB-rich ones. Thus, as-produced original P(3HB-co- different to the component copolymers – it is important to
4HB) samples should be considered as immiscible polymer establish the CCD of any PHA produced using bacterial PHA.
blend systems. Whichever method is used for solvent fractionation, care
As-produced P(3HB-co-3H4MV) was also found to have must be taken to ensure that the polymer is not degraded
a broad CCD (Table 2) [304]. Both the melting tempera- during the sample preparation and fractionation stages
ture and glass-transition temperature of the copolymers since it has been shown that dissolution of the polymer
decreased with an increase in the 3H4MV content. The Tm followed by precipitation can cause a severe decrease in
dropped from 172 to 99 ◦ C and the enthalpy of fusion fell molecular weight [22], particularly for chloroform/alcohol
from 92 to 9 J g−1 , while Tg decreased with an increase in combinations over time. In addition chloroform often con-
3H4MV content. The rates of enzymatic erosion initially tains ethanol as a stabilizer; refluxing for extended periods
increased markedly with an increase in the 3H4MV content in this can result in significant esterolysis particularly if acid
to 7 mol%, then decreased. The crystallinity of fraction- impurities are present to catalyze the reaction.
ated P(3HB-co-3H4MV) films decreased from 60 to 13% (by
WAXS) as the 3H4MV fraction increased from 0 to 39 mol%.
Crystalline structures were those of P(3HB) and the d spac- 10. PHA–PHA blends
ing remained constant with mol% 3H4MV; 3H4MV was
excluded from the P(3HB) crystal [304]. Given that many as-produced microbial PHA polymers
Valentin et al. [328] prepared 3HB and 3-hydroxy-4- consist of blends of copolymers of differing composition,
pentenoic acid (3HPE) containing PHAs using Burkholderia it is obviously important to understand the effect of such
sp. DSMZ # 9243. From solvent fractionation of the result- polymer blend systems on polymer properties.
ing polymers it was shown that the resulting material was Chain isomorphism can occur with blends of different
a blend of two homopolymers poly(3HB) and poly(3HPE), P(3HB-co-3HV) copolymers or with P(3HB)/P(3HB-co-
rather than a co-polyester with random monomer distri- 3HV) blends [14] whereby two or more different
bution. semicrystalline polymers are miscible in the crystalline as
Chen et al. [329] fractionated mcl-PHA containing well as the molten state. In such cases, the blend will exhibit
high proportions of 3-hydroxydodecanoate (3HDD) and 3- single composition-dependent values of Tg and Tm . How-
hydroxytetradecanoate (3HTD), using mixed solvents of ever, this is not the general case for PHA: there have been a
chloroform/ethanol. The first fractions had both higher number of studies exploring the effect of deliberate blend-
3HTD composition and higher molecular weight, although ing of different random copolymers of PHA on polymer
fractions with relatively low Mn but extremely high 3HTD properties, and it has been shown that there is only a nar-
content were also precipitated earlier. The samples with row window of composition in which chain isomorphism
higher 3HTD composition had in general higher Tm and can occur.
420 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

Fig. 9. Variation of phase structures in PHB/PHB-HV blends.


Reprinted (adapted) from Yoshie and Inoue [337], Copyright (2005), with permission from John Wiley and Sons.

Kamiya et al. [305] prepared three blends of bacterially thicker amorphous layers. When the difference exceeds
synthesized random copolymeric P(3HB-co-3HV) (contain- 30–40%, the blends become immiscible [336]. The co-
ing 23.4, 40.7 and 92.5 mol% 3HV respectively) with P(3HB), crystallizable blends by contrast had the same properties
as well as a blend of the 40.7 and 92.5 mol% 3HV copoly- as the true copolymer of same overall 3HV content. It was
mers, using solvent casting. Two melting peaks were found found that the melting temperatures reflected the blend
for each sample, except for the mix of P(3HB) and 23.4 mol% type; blends with more than two crystalline phases had two
P(3HB-co-3HV) that had a melting point corresponding or more melting peaks, while only one melting peak (at the
to pure P(3HB) homopolymer. Melting peak temperatures temperature expected from the average HV content) was
were independent of the relative amount of each polymer, observed when blends formed complete co-crystals. Yoshie
indicating that there were different phases present. Kuma- and Inoue [337] found that PHB/PHBV blends exhibited a
gai et al. [333] prepared P(3HB)/P(3HB-co-76 mol% 3HV) spectrum of properties ranging from complete cocrystal-
blends and found that all blends were immiscible based on lization (with rapid crystallization rates) through partial
Tg data. The rate of enzymatic degradation was found to be segregation with some cocrystallization to complete phase
higher for the blends than for the individual components, segregation depending on the HV content of the copolymer
perhaps due to the large holes that formed. Barham, Organ (Fig. 9). If the crystalline phase is rich in the faster-
and coworkers [211,240,241,334] also showed that blends crystallizing structures, then the melting temperature will
containing above 60–70% P(3HB-co-3HV) (18.4 mol% HV) be higher than expected, and at least two melting temper-
with P(3HB) were phase separated in the melt, with two- atures will be observed in phase separated systems [223].
phase crystallization occurring. There was some evidence Other blend mixes have also been explored. Pearce
of partial miscibility between the blend components as et al. [223] blended P(3HB-co-3HV) (77 mol% 3HV) and
judged by a region of liquid–liquid phase separation, and P(3HV) at a 50/50 ratio; a single melting peak was observed
there was a kinetic effect associated with the characteris- plus a linear Hoffman–Weeks plot was obtained, implying
tics of the phases. Saito et al. [335] reported the almost that co-crystallization occurred. The temperature of maxi-
perfect co-crystallization of P(3HB) and P(3HB-co-3HV) mum crystallization rate decreased as mol% HV decreased.
when the HV content was less than 9 mol% 3HV, whereas Blends of P(3HV) and P(3HB) were also prepared (at 77%
phase segregation preceded the crystallization in blends HV load): individual glass transition temperatures were
when the 3HV content was >15 mol%. It was proposed obtained showing that the homopolymers phase separated
that the component of slower crystallization is partially in the melt. This suggests that, for the copolymer, the co-
excluded from the crystals (since P(3HB-co-3HV) with a crystallization may be partially kinetic in origin. Gassner
lower crystallization rate moves away from the growing and Owen [225] also investigated the blends of P(3HB)
front) resulting in a P(3HB)-rich crystalline phase with and P(3HV). Their results show that melt-pressed blends
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 421

contain phase-separated domains in the melt that then 12 and 19 mol% 3HV) and P(3HB-co-3HP) (10 and 14 mol%
crystallize as P(3HB) and P(3HV) spherulites respectively. 3HP), the resulting polymers were generally phase segre-
The two melting regions were detected by DSC, and were gated with partial co-crystallization, although some blends
almost unaffected by blend composition. A summary of co-crystallized. The copolymer with the fastest crystal-
these outcomes was prepared by Yoshie and Inoue, and is lization rate forms the lamellae and the other copolymer
given in Fig. 9 [337]. is involved into the lamellae. The same was observed
The effect of molecular weight has also been explored. by Yoshie et al. [336] who found that the blends of
High molecular weight P(3HB) (Mw = 0.35 × 106 Da) and P(3HB-co-3HV) and P(3HB-co-3HP) adopt various phase
low molecular weight P(3HB) (Mw = 3900 Da) were blended structures depending on the extent of phase segrega-
with P(3HB-co-3HV) (12 mol% 3HV, Mw = 0.24 × 106 Da). tion, from complete co-crystallization to complete phase
After isothermal crystallization, a lowering of the Tm of segregation. The degree of phase separation was found
P(3HB) was observed, indicating miscibility in the liquid to depend on the relative isothermal spherulite growth
and the solid state. Only one melting peak was observed, rate, G, for the two copolymers in the blend. Feng et al.
indicating co-crystallization [227]. Saad et al. [338] [342] prepared blends of finely fractionated P(3HB-co-
prepared solvent cast blends of P(3HB) and oligo[(R,S)- 3HHx) copolymers, including blends with P(3HB), using
3-hydroxybutyrate]-diol and found that up to 60 wt% solvent casting. It was found that when the difference in
oligo-HB, the blends behaved as single-phase amorphous the 3HHx content between the two different copolymers
glasses with a composition dependent glass transition, Tg , was less than 20 mol%, then they were miscible in the
and a depression in the equilibrium melting temperature amorphous phase; when the difference was larger than
of P(3HB), indicating that the components were miscible. 30 mol%, then they were immiscible. In addition, a con-
The glass transition of melt-crystallized blends contain- tinuous compositional distribution was found to favor a
ing 0–20 wt% oligo-HB were dielectrically investigated and homogenous amorphous phase, while bimodal composi-
revealed the existence of a single alpha-relaxation process tional distributions lead to a phase-separated morphology.
for blends, indicating the miscibility between amorphous Zhao et al. [250] also blended PHA and Nodax (13 mol%
fractions of P(3HB) and oligo-HB. HHx); the blends were miscible, and had good mechani-
Tacticity also influences compatibility between PHA cal properties, e.g. elongation to break increased from 15%
components. Pearce et al. [339] prepared blends of bacte- to 106% as % Nodax went from 40% to 60%. Abe et al. [88]
rial and partially isotactic P(3HB); both solvent cast films sought to compatibilize blends of P(3HB)/P(3HHx) through
and rapidly co-precipitated powders were used. In the the use of a synthetic block copolymer of atactic P(3HB)
latter case a single melting point was observed (i.e. a sin- and P(3HHx). The addition of small amounts of this block
gle phase material was produced), while slow evaporation copolymer to blends of the two homopolymers led to a
led to two endotherms, one for each phase. Totally atac- decrease in the size of the P(3HHx) dispersed domains and
tic P(3HB) was found to be incompatible with bacterial improved their mechanical properties. It was found that the
P(3HB) but did cause a significant drop in the melting point. rate order for enzymatic hydrolysis of these samples was
Abe et al. [88,89] also prepared binary blends of synthetic P(3HB)/P(3HHx) blend > ternary blend > P(3HB) > P(3HHx).
atactic or syndiotactic and bacterial isotactic P(3HB) and In summary, phase separation is a common character-
found that the equilibrium Tm decreased with the addi- istic of blends of PHA copolymers of sufficiently different
tion of synthetic P(3HB) from 191 to 174 ◦ C, suggesting that composition, so in interpreting the properties of any PHA
the blends are miscible in the melt and in the amorphous polymer, it is important to understand the compositional
state. Crystallinity (based on XRD) also decreased with an make-up and distribution of that polymer.
increase in synthetic P(3HB) and the mechanical proper-
ties changed remarkably. For example, Young’s modulus 11. Mechanical properties of PHA
drops from 1560 MPa to around 180–340 MPa with the
inclusion of 50% atactic or syndiotactic P(3HB) (of vary- All of the preceding methods for controlling compo-
ing % (R)). Scandola et al. [340] also looked at blends of sitional distribution, crystallinity, microstructure, blend
synthetic atactic P(3HB) with a natural bacterial P(3HB-co- composition and molecular weight were techniques for
3HV) containing 10 mol% of 3HV units. These blends were manipulating polymer processing/mechanical properties.
also miscible in the melt and solidified to give a spherulitic Three core mechanical properties should be considered
morphology. The degree of crystallinity decreased as the for commodity applications [288]: (1) the elongation at
content of atactic P(3HB) increased, and the elongation at break, which is a measure of toughness or total deformation
break for a sample containing 50% of atactic P(3HB) was before fracture; (2) the Young’s modulus which is a mea-
30-fold that of pure P(3HB-co-3HV). The rate of enzymatic sure of stiffness; and (3) the ultimate tensile strength which
degradation of the blends was higher than that of P(3HB- is a measure of strength prior to the onset of permanent
co-3HV) and increased with the atactic P(3HB) content in plastic deformation.
the blends, whereas pure atactic P(3HB) did not biodegrade
under these conditions. It was concluded that a partially 11.1. Properties of pure culture PHA
crystalline P(3HB) was essential for enzymatic degradation.
It is not just the P(3HB-co-3HV) copolymers that display It is well known that P(3HB) homopolymer and to a
phase separation; this phenomenon has also been demon- lesser extent the P(3HB-co-3HV) copolymers of low HV
strated for a range of other PHA copolymers. Hirota et al. content are stiff and brittle, with poor impact strength,
[341] found that on solvent blending P(3HB-co-3HV) (7, because of the relatively high crystallinity of the materials
422 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

Fig. 10. Effect of aging time on mechanical properties of P(3HB-co-3HV):


+: tensile modulus; : Izod notched impact strength.
Reprinted from de Koning and Lemstra [381], Copyright (1993), with per-
mission from Elsevier Limited.

and in particular the significant secondary crystallization


that occurs post processing with age (Fig. 10). Based on
the data available (Appendix B) the average properties
for P(3HB) are: Tg ≈ 5 ◦ C (by DSC) or 12 ◦ C (by DMTA);
Tm ≈ 176 ◦ C; tensile modulus 2.9 GPa; tensile strength
37 MPa; maximum elongation to break ∼4%. Addition of
a nucleating agent can increase the elongation to break
slightly [343], typically up to 5%.
This brittleness has been a reported obstacle to the
practical applications of these materials, and significant Fig. 11. Effect of HV content (mol%) of P(3HB-co-3HV) on Young’s modu-
research effort has been devoted to manipulating these lus and tensile strength (see Appendix B for references).
mechanical properties. They are known to be dependent
on polymer composition and processing conditions, and in although there is a lot of scatter in the reported data. There
particular on whether or not the polymers were produced is even more variation in the data for elongation to break
using solvent casting or melt mixing and casting. It was (Appendix B), which is a property that is very depend-
noted by Holmes [22], for example, that films of P(3HB) and ent on processing and testing conditions. Considering the
its copolymers prepared by solvent casting have a very fine information presented in Section 9, it is also likely that
spherulitic morphology because of low crystallization tem- some of the materials analyzed are in fact blends, or of
perature and resulting high nucleation density, which can a blocky nature, rather than being true copolymers; only
result in higher elongation to break and impact strength. data derived from random copolymers of very narrow com-
The same was observed for P(3HB-co-3HHx) copolymers positional distributions that have been allowed to fully
(Appendix B) where the solvent cast films had much higher crystallize provides reliable information on correlations
elongation to break than melt pressed films, although between mechanical properties and measures of crys-
tensile strength and Young’s modulus were similar. In tallinity, microstructure, blend composition and molecular
assessing mechanical properties, therefore, it is important weight.
to note how samples are prepared. Annealing should improve the polymer properties, with
One of the most direct solutions for improving mechan- some suggestion that maximum elongation to break up
ical properties of PHA is to use copolymeric materials, to 130% could be found for 12% HV even 3 months after
such as P(3HB-co-3HV) with higher HV contents, or mcl- heat treatment [343]. Noohom et al. [344] also noted
PHA copolymer. Compared to P(3HB), P(3HB-co-3HV) has that the effects of annealing on mechanical properties for
decreased stiffness and brittleness, increased flexibility melt pressed discs of commercial P(3HB-co-3HV) (Fig. 12)
(higher elongation to break), and increased tensile strength – Young’s modulus increased significantly where tensile
and toughness. These improved properties have been strength was little affected apart from an increase in vari-
attributed to the presence of dislocations, crystal strain ability. It was noted that there was <20% loss in molecular
and smaller crystallites due to the insertion of the 3- weight under these conditions. Bloembergen et al. [345]
hydroxyvalerate unit into the P(3HB) lattice [179], with created synthetic analogs of P(3HB-co-3HV), in which long
HV units acting as defects in the HB crystals. This results blocks of the copolymer chain had either the (R) or the
in significant depression in the melting point, the heat (S) configuration, by ring-opening polymerization of mix-
of fusion of the HB component crystal and the fold sur- tures of racemic ␤-butyrolactone and ␤-valerolactone with
face free energy [172]. Both Young’s modulus and tensile a stereoselective alumoxane catalyst. It was found that
strength decrease with an increase in HV content (Fig. 11) by blending this synthetic P(3HB-co-3HV) with bacterial
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 423

P(3HB-co-3HHx)

Foam mold

0.8–1.8

120–125
10–100
5–10

10–20
Kaneka

1.2
Extrusion, injection

1.2–1.8
and fibre
Biocycle

15–25

25–30
20–30
2400-5

1.2
P(3HB-co-3HV)

Extrusion and
Fig. 12. Effect of annealing temperature on mechanical properties of
P(3HB-co-3HV): 䊉: Young’s modulus and : compressive strength.

injection

170–175
Biocycle

2.5–6
2.5–3
10–12

50–60
30–40
1.22
Reprinted from Noohom et al. [344], Copyright (2009), with permission

1000
from IOP Publishing.

P(3HB), the lamellae crystals were disrupted somewhat

P(3HB-co-3HV)

Injection mold
and as a result the brittleness was reduced and the tough-
ness increased.

5–10
ENMAT

1.25
The degree of molecular segmental motion the amor-

1.4

1.4
36

61

147
143
phous portion of PHA can achieve is strongly influenced by
the difference between the Tg and the end-use tempera-

Extrusion and injection


ture. Therefore, even at the same crystallinity, a material
with a lower Tg (i.e. a longer side chain) should be more
flexible [12]. The mechanical properties for the mcl-PHA
copolymers P(3HB-co-3HHx), P(3HB-co-4HB) and P(3HB-
co-3H4MV) have also been characterized (Appendix B) P1002
and as expected these materials are more flexible than
Mirel

1.3

1.9
the P(3HB-co-3HV) copolymers. As for P(3HB-co-3HV),

26
13

35

137
P(3HB) copolymers

the modulus and tensile strength for these copolymers


decrease with an increase in comonomer content, while
Injection mold

elongation to break increases. Again, there is very signifi-


Properties of commercial PHA (data reproduced with permission from Shen et al. [406]).

cant variability in elongation to break. It should be noted


P1001

1.39
Mirel

that the mechanical properties of a number of terpolymers


3.2
28
6

46

148
and above have also been characterized, and that the elon-
gation to break for these materials can often be higher than
for the comparable copolymers containing the equivalent
Injection mold

monomer units [346].


Shimamura et al. [347] examined the dynamic mechan-
Biomer

9–13

24–27
60–70
1.25

6–9
P226

ical properties of P(3HB-co-3HP) but to date the static


35

96

mechanical properties of this copolymer have not been


characterized. Gagnon et al. [247] found that poly(3-
Injection mold

hydroxyoctanoate) (PHO) crystallized very slowly, taking


approximately 7 weeks at 20 ◦ C and 16 weeks at 5 ◦ C to
Biomer

reach a constant value. The thermal history had a large


P(3HB)

10–17
60–70
18–20
1.17
5–7
240

effect on the mechanical properties, with Young’s mod-


17

53

ulus varying from 2.5 to 9 MPa, tensile strength at break


from 6 to 10 MPa, and ultimate elongation from 450% to
VICAT softening temperature (◦ C)

300% depending on the crystallization temperature. The


equilibrium melting point of PHO was approximately 68 ◦ C.
Melting temperature (◦ C)
Melt flow rate (g/10 min)

A summary of the properties of some commercial grades


Flexural strength (MPa)
Flexural modulus (GPa)
Tensile strength (MPa)

Tensile modulus (GPa)

of PHA is also given in Table 3, illustrating the current range


Application grade

of properties available from processed resins [11].


Density (g/cm3 )
Crystallinity (%)

Elongation (%)

11.2. Properties of mixed culture PHA


Table 3

The physicochemical properties of PHA produced using


mixed culture production have not been widely reported
424 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

until recently. While in principle the properties of PHA expected. In addition Serafim et al. [41] noted that for a
obtained from mixed cultures should be a combination P(3HB-co-3HV) copolymer (30 mol% HV) obtained under
of those obtained from pure cultures, the complexities of ADF conditions, the as-produced material had an impurity
these diverse biological systems mean that, as discussed in of molecular weight of 2000; a 25% increase in the enthalpy
the previous sections, the final properties may be hard to of melting was observed when it was removed, indicating
predict and so it is worth considering separately the data its presence was possibly disrupting the lamellae forma-
available to date. tion.
The first reports on characterization of the molecular Patel et al. [349] produced low polydispersity, ultrahigh
weight properties of mixed culture PHA were published molecular weight P(3HB) and P(3HB-co-3HV) at 18 wt% and
in 1998 [348] and were of PHA synthesized using an 30 wt% respectively using a mixed culture under conditions
EBPR process. In this case P(3HB-co-3HV) copolymers were of extreme nitrogen limitation (such that nitrogen fixa-
produced with an HV content determined by the feed com- tion was necessary). Only 3.3% of the carbon substrate was
position. A summary of the studies for which thermal data recovered as PHA, but the properties as reported were very
have been obtained is given in Appendix A. comparable with pure culture PHA (Appendices A and B).
As Bengtsson et al. [62] noted, there is often a higher Tm Consistent with other literature, the Tg seems to be mainly
and lower enthalpy of melting for mixed culture materials determined from the HB content.
than is found in the literature for PHA from pure culture. The thermal stability of mixed culture P(3HB-co-3HV)
One hypothesis is that micro-block copolymers could be materials was assessed by Carrasco et al. [350] using ther-
forming based on the feeding strategies adopted [62]. It mogravimetric analysis. They found that these materials
was observed that with pulse feeding, there was often vari- (from biomass enriched using different procedures, but
ability in the feed concentration of the VFAs – for example, both containing 20 mol% HV) had a similar thermal stability
Bengtsson et al. [62] found that in the early stages of each (with decomposition temperatures of 266.7 and 273.3 ◦ C)
pulse, propionate was present and PHA rich in 3HV was to the commercially available P(3HB) and P(3HB-co-3HV)
produced. At the end of the pulse, propionate had all been (10 mol% HV). They also found that an increase in organic
consumed and PHA dominated by 3HB was produced. This load during the mixed culture accumulation was linked to
alternating feeding reflects the deliberate strategy adopted an increase in thermal stability.
for the formation of block copolymers in PHA [81]. It is In the only study to date that has reported on the
likely that other PHA production systems experience a sim- mechanical properties of mixed culture PHA, Dobroth
ilar dynamic in VFA feeding [54], leading also to similar et al. [351] enriched a mixed microbial consortium from
types of block or blend type polymers. Dai et al. [44,73,74] a wastewater treatment plant using crude glycerol as car-
studied the production of PHA copolymers using enriched bon feed. The P(3HB) that accumulated was produced from
GAO cultures and also found there was a blocky character to the methanol fraction. The thermal properties and molec-
the materials produced. It is believed that the P(3HB) gen- ular weight are reported in Appendix A; the molecular
erated during the aerobic period formed as separate P(3HB) weight was slightly low, and there was some variability
homopolymer “blocks”, which led to a high melting tem- in the thermal properties, which may be in part related to
perature. The small fraction of P(3HB-co-3HV) was likely the molecular weight variability, but in general the data
present in the form of separate copolymeric chains. It was is comparable with that of pure culture P(3HB). The tensile
also found that when propionate was used as the sole car- strength and Young’s modulus of the solvent cast films after
bon source, random copolymers of 3HB, 3HV, 3H3MB and aging were 14 MPa and 1.8 GPa, respectively, which were
3H2MV were produced, while co-feeding propionate and lower than generally reported for P(3HB). This could be
acetate resulted in heterogeneous copolymers with non- because of the lower molecular weight compared to com-
random sequence distributions. This was consistent with mercial materials (308 kDa c.f. 437 kDa) or because there
sequential uptake of the carbon source, forming HV- and was some minor contaminant or copolymeric component.
HMV-rich copolymers from propionate and HB- and HV- Tensile strength at 20 MPa was comparable with pure cul-
rich copolymers from acetate. By contrast, Inoue et al. [314] ture materials as were the Young’s modulus (2.3 GPa) and
examined the microstructure of the PHAs produced using elongation at break (1.3%).
EBPR sludge with propionate feed as the carbon source In the studies that have produced the highest PHA
using 13 C NMR. Four types of monomer units were iden- contents to date, Johnson et al. [77,78] found that as HV
tified (3HB, 3HV, 3H3MB and 3H2MV) and in this case the content increased from 15 to 39 mol%, Mw decreased from
data suggests they were statistically random copolymers. 6.5 to 2.2 × 105 Da. The crystallinity by XRD was 40 ± 5%
There was no true evidence for four-monomer-component (which is the first reference in the mixed culture literature
copolymers. to analysis of crystallinity using this method, and which
For blocky materials, the crystal unit will likely have is in line with literature expectations). The crystallization
higher HB content than the bulk. For blends, the melting rate was very slow; immediately after melting there were
temperatures will reflect the faster crystallizing compo- no crystals formed. After one day a large number of very
nents in the blend that will be higher in the HB or HV small crystals was observed. The crystallization rate slowed
comonomer unit than the bulk unless they are completely as 3HV content increased.
miscible blends. Either way, the melting temperature is In summary, it is important to demonstrate that the
likely to be higher than expected from the nominal 3HV quality of the biopolymers produced by mixed microbial
content. The lamellae are also likely to be small and cultures can consistently meet the standards required for
imperfect, leading to lower heats of fusion than would be use in commercially interesting plastic applications. More
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 425

Table 4
Effect of block copolymer structure on thermomechanical properties of PHA.

Sample 3HB 3HV 3HHp 3HHx Tm (◦ C) Tc (◦ C) Tg (◦ C) Mn (×105 ) Mw /Mn Young’s Tensile Elongation
mol% mol% mol% mol% modulus strength at break
(GPa) (MPa) (%)

P(3HB) homopolymer 100 0 0 0 171.8 41.2 3.1 – – 1.47 18 3


Random copolymer 4.4 46.5 49.1 0 – – −26.4 – – – – –
Random copolymer 72.9 13.1 14 0 – – −7.3 1.27 2.9 0.13 7.0 462
Block copolymer 71.5 10.2 16.5 0 170.6 50.6 −23.6,3.5 0.52 8.7 0.37 7.5 63
Blend 71.5 10.2 16.5 0 166.5 54.3 −20.0,4.4 0.81 4.4 0.26 5.3 24
PHBHHx 88 0 0 12 – – – – – 0.28 7.0 400
P(3HV) homopolymer 0 100 0 0 106.2 57.9 −15.7 – – 0.39 6.6 3.5

Data abstracted with permission from Li et al. [298].

fundamental studies on thermomechanical properties of storage modulus with cooling. The properties are summa-
these materials are needed, and with these one would hope rized in Table 4. Young’s modulus was significantly higher
to gain deeper appreciation of the importance of controlled than for the random or blend copolymers of equivalent
or uncontrolled compositional distribution variation on composition, and tensile strength was higher than for the
microstructure and relevant mechanical properties. blend, although elongation to break is significantly lower
than for the random material.
Overall, while materials obtained as block copolymeric
11.3. Properties of “block” copolymeric PHA compounds do exhibit unusual properties, there is incon-
sistency in results to date and some detailed structure-
A number of studies have explored the effect of nomi- property relationships that correlate comonomer sequence
nally blocky coplymeric compositions on mechanical and distributions with end-use properties need to be developed
rheological properties of PHA. Pederson et al. [297] in their and detailed characterization undertaken.
study of HB/HV di- and triblocks used solvent fraction-
ation to separate homopolymer and random copolymer
from the 30% fraction with block copolymer character- 11.4. Physical techniques for the modification of PHA
istics that was present. The resulting blocky materials mechanical properties
showed significantly increased elongation to break – for
example, the polymer formed using an 8 min cycle had One common method for introducing modifications to
250% elongation to break while having the largest Young’s mechanical properties is to induce strain into the struc-
modulus and tensile strength of the materials produced. ture through stretching or melt spinning of PHA fibers
Additional mesophase transitions were observed during [352]. Miller and Williams [353] used monofilament fibers
rheological testing, which are only found in block copoly- of P(3HB) and P(3HB-co-3HV) (from ICI) for implantation
mers, and secondary transitions in the storage modulus studies in rats. The mechanical properties of the origi-
(not found in random copolymers) were also observed. nal materials were significantly improved as fibers (tensile
When McChalicher et al. [288] prepared block copoly- strength of P(3HB) of ∼226 MPa at 60 ◦ C for example) –
mer (P(3HB)-block-P(3HB-co-3HV)) it was found that the there was little change in these properties with time in
random P(3HB-co-3HV) copolymers had less than 20% vivo. Schmack and coworkers [354] obtained P(3HB) fibers
elongation at break within a few days after annealing. How- with tensile strengths of 330 MPa and a tensile modulus of
ever, the block copolymer films retained higher than 100% 7.7 GPa using high-speed melt spinning and drawing fol-
elongation at break a full 3 months after annealing. It was lowed by annealing (with a draw ratio of 6.9). Both the ␣-
also noted that “because block copolymers covalently link and ␤-form crystals were present (while at a draw ratio of
polymers that would otherwise form thermodynamically 4, only the ␣-form was found). Iwata et al. [352,355] pro-
separate phases, the rates and degrees of crystallization of duced high strength ultrahigh molecular weight (UHMW)
the block copolymers are less than the random copolymer P(3HB) films with a strength of 287 MPa or ∼400 MPa using
samples”. cold or hot two-step drawing respectively; these proper-
Finally Li et al. [298] used ethanol fractionation to ties were stable for 4 months, indicating that secondary
remove random copolymer and retain the block copolymer crystallization did not occur for these materials. Yamane
in a mixture of materials comprising a blocky copolymeric and coworkers [180] obtained P(3HB) fibers with a ten-
component containing 70% P(3HB) as one block and the sile strength of 310 MPa by a two-step annealing process
random copolymer P(3HV-co-3HHp) as another block. A of melt-spun fibers; in an extension of this work [185] they
comparison was made with the random copolymer of the obtained P(3HB) fibers with a tensile strength of 416 MPa
same composition, as well as the homopolymer of P(3HB) by a combination of cold drawing and two-step draw-
and the random copolymer of 3HV and 3HHp; the blend of ing methods followed by annealing under tension. Strong
P(3HB), P(3HV) and P(3HHp) was also prepared to provide fibers with high tensile strength were produced by Iwata
a composition similar to that of the block copolymer. et al. [355,356] using UHMW P(3HB), by quenching in
The 13 C NMR signal of the carbonyl region was used to ice-water followed by two-step drawing. Tensile strength
distinguish the block copolymer from the blend, as was increased from 38 MPa to 121 MPa after cold drawing,
rheological analysis, which showed a rapid drop in the and then reached 1.32 GPa after total draw-down reached
426 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

60-times, with an elongation to break of 35%. Using micro- when the completely melted polymer is cooled from the
beam X-ray diffraction, it was shown that the central core melt; self-seeding, which is nucleation by unmelted crys-
of the fiber contained both the ␣- and ␤-crystal forms, tallites when the polymer is not completely melted (the
with an ␣-form outer sheath. It is believed that this is due maximum temperature is kept to less than 15 ◦ C above the
to the rotation of the ␣-crystal form lamellar crystals in melting temperature); enhanced self-nucleation, which is
the outer sheath during the drawing process, which pre- like self-seeding in that when the maximum tempera-
vents the molecular chains between the lamellar regions ture is limited to only 1 ◦ C above the melting temperature,
from elongating sufficiently. A similar process was adopted and the original spherulites simply reappear on cooling;
for commercial P(3HB) fibers (Mw = 7 × 105 ); a four-times nucleation by impurities such as talc which are coated
drawn fiber crystallized for 72 h had a tensile strength of with bound crystallites having a higher-than-usual melt-
740 MPa, elongation to break of 26% and Young’s modulus ing point; and finally nucleation through the addition of
of 10.7 GPa [355]. In this drawing method, small crys- nucleating agents which can initiate nucleation from their
tal nuclei grow initially during the isothermal nucleation surface. The nucleation rate of PHA copolymer is extremely
process. The molecular chains between the small crystal low so in general nucleating agents need to be added to
nuclei then act as entanglement points that are oriented facilitate processing. Spherulites of P(3HB-co-3HV) con-
by stretching. taining from 0 to 25 mol% 3HV are formed only after
Tanaka et al. [357] stretched P(3HB) fibers after isother- melting above 184 ◦ C. This suggests that seeding of the
mal crystallization near the glass transition temperature. copolymer spherulites by P(3HB) nuclei is occurring. There
Increasing the time for crystallization resulted in a decrease are a large number of nucleating agents that have been
in the maximum draw ratio. However, if the fibers were tried with PHA materials such as boron nitride [255,361],
allowed to crystallize for 24 h then the tensile strength saccharin [361], lignin [362], terbium oxide (Tb2 O3 ), lan-
increased markedly. They also used 8% HV P(3HB-co-3HV) thanum oxide (La2 O3 ) [255], ␣-cyclodextrin [363], orotic
– in a similar way, the ten times drawn fiber had a tensile acid [364], and phthalimide [365]. Clarifying agents may
strength of 90 MPa. However, ten times drawn P(3HB-co- also be required to prevent clouding.
3HV) after an isothermal nucleation treatment had a tensile The use of impact modifiers, whereby a rubbery additive
strength of 1.06 GPa, elongation to break of 40% and Young’s such as hyperbranched polymers, chlorinated polyethyl-
modulus of 8.0 GPa. In contrast to the P(3HB) fibers, those ene, polychloroprene and poly(butyl acrylate) is added to
made from P(3HB-co-3HV) did not exhibit a core/shell improve toughness is also common; however, this can be
arrangement. Instead the ␤-crystal form was uniformly dis- a result of a decrease in yield stress rather than a genuine
tributed throughout the fiber along with the ␣-form, and increase in the material’s toughness as judged by the crit-
many voids were seen in the cross-sectional analysis. ical stress intensity factor [22]. Plasticizers can also have
The spherulites of P(3HB) and P(3HB-co-3HV) of low a dramatic effect on Tg with concomitant improvements
3HV content usually contain cracks, which cause brit- in ductility and crystallization kinetics at the expense of
tleness. Radial cracks [358–360] occur even at room lowering softening points [366]. Interactions with many
temperature; circumferential cracks appear if films are blends are dominated by the tendency for P(3HB) to self-
crystallized at higher temperatures and then cooled. It has crystallize with exclusion of the additive to the crystalline
been suggested that this is due to differences in the thermal phase. In general, the more polar the plasticizer, the greater
expansion coefficients along the radius and circumference, its impact on the polymer and its compounds, since plas-
which causes large internal stresses. It is possible to make ticizers work by interacting with the polymer chains,
spherulites without cracks either by using a copolymer increasing the free volume and thus significantly lowering
of more than 12 mol% 3HV content, or by increasing the the glass transition temperature.
nucleation density until the spherulites are small enough The use of fillers such as calcium carbonate, china clay
to avoid cracking (i.e. <20 ␮m) [300]. Cracks can also be and nano-sized fillers has been extensively explored [367];
sealed by rolling the films [300]. the fillers need to be well dispersed into the matrix. For
clay particles, for example, it is preferred if the particles
11.5. Effect of additives are fully exfoliated, although an intercalated structure is
also beneficial. With good integration between filler and
The use of additives is also a common approach for prop- polymer, the mechanical reinforcement, thermal stabil-
erty modification, and a thorough review is beyond the ity, biodegradability and barrier properties were generally
scope of this report; the following is a brief summary of improved. Galego et al. [368] suggested that P(3HB-co-
some of the approaches that can be adopted. It should be 3HV) containing less than 8% HV content was best for a load
noted that additives can also be grouped according to their bearing application, because the elastic modulus reduced
environmental performance, e.g. traditional additives that dramatically above this level.
have no impact on health or the environment and do not De Koning et al. [248] recognized two major prob-
limit biodegradability; “renewable” additives derived from lems with the higher HV content and mcl-PHAs: their low
natural sources, which are not necessarily biodegradable; crystallization rates and their tendency to soften and lose
and additives that are both renewable and biodegradable. their coherence at temperatures as low as 40 ◦ C makes
Nucleating agents accelerate the rate of crystallization processing difficult. Crosslinking (such as with peroxide
and can help limit secondary crystallization. Five types addition during processing) has been tested as one solu-
of nucleation processes have been identified for P(3HB) tion [248]. Other additives for processing would include
[155]. They are: homogeneous nucleation, which happens slip and antiblock agents, heat stabilizers to increase the
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 427

processing window, antioxidants, impact modifiers, lubri- heterogeneities and processing conditions. Care therefore
cants to enhance flow properties for injection molding, and needs to be taken that the PHA material being used in a
melt strength enhancers for improved draw in films and particular study is well-characterized before extrapolating
thermoforming [369]. conclusions to other systems.

11.6. PHA blends 12. Outlook

The mechanical properties of blended PHA–PHA phase Polyhydroxyalkanoates have been intensively studied
separated materials have been assessed [370]. Blends of for some decades and are very attractive candidates as
P(3HB) with random copolymeric P(3HB-co-3HV) (contain- supplementary or replacement materials for petroleum-
ing 18.4 mol% HV and 19.7 mol% HV, with Mw of 732 and derived polyolefins, having good general properties and
42 kDa respectively) were prepared using a single screw being truly biodegradable. Many of the limitations that
extruder followed by hot melt pressing. The 19.7 mol% were identified early on have been addressed through the
3HV copolymer blend also contained nucleating agent. use of different copolymeric compositions and combina-
Samples crystallized from such phase separated melts tions to improve mechanical properties and also through
showed a higher fracture toughness than reference copoly- the addition of appropriate additives and advanced
mer. In fresh samples, there was a pronounced impact on processing and modeling. Mixed culture PHA production is
yield and fracture of the blends, with the blends being emerging as an attractive approach that has the potential
more ductile. However, this improvement was found to be to use cheap waste materials and low-capital installations
largely lost during the ageing process. Satoh et al. [371] for the production of cost-effective PHA. However, in spite
studied the hydrolysis of P(3HB)/(3HB-co-3HV) (22.3 mol% of the wealth of experience embodied in the reviewed sci-
3HV) blends. DSC results showed that P(3HB) was the entific literature, core fundamental principles still need
main component of the crystalline phase, and the crys- to be developed for broader practical implementation of
tallinity decreased with increasing wt% P(3HB-co-3HV) in PHA as bio-based engineering materials. Evidently a more
the blends. There was a correlation between an accelerated mechanistic understanding of structure function relation-
enzymatic degradation rate and a decrease of crystallinity ships is required for PHA, particularly at higher comonomer
in the blends. contents. Therefore it is suggested that efforts be placed on
The blending of PHA with other polymeric materials three main fronts:
is a technique for modifying end-use properties that is
very cost effective and relatively easy to adopt. P(3HB- 12.1. Modeling of PHA accumulation process and chain
co-3HV) is miscible with many polymers such as poly- growth kinetics
ethylene oxide, polyvinyl alcohol, poly (p-vinylphenol),
poly (epichlorohydrin-co-ethylene oxide) and poly(vinyl There is a need to develop predictive tools that can
acetate) (displaying a single glass transition temperature). model the PHA accumulation process, and in particular
P(3HB-co-3HV) is immiscible with a great many other provide a detailed polymer kinetic model for P(3HB-co-
materials, including (as examples) polybutylene adipate, 3HV) chain growth, and thereby assess copolymer type and
polylactic acid > 20 kDa, ethylene-vinyl acetate copoly- distribution in the matrix. One of the limiting factors with
mer (containing 28 mol% vinyl acetate), poly(ethylene respect to developing such robust models of PHA molecular
succinate), poly(propylene carbonate), poly(butylene suc- dynamics in vivo is that there have been too few studies on
cinate), and poly(vinylidene fluoride) (see [19,372,373] the mechanistic prediction of PHA accumulation and chain
and references therein). The use of multilayer and fiber- growth under varying culture conditions. It is important
reinforced composites is also a valuable approach that can for the development of good process control that further
lead to property improvement while lowering overall cost fundamental research into effect of process conditions,
[373]. feeding strategy and culture type on the kinetics of molec-
ular weight and compositional distribution be undertaken,
11.7. PHA processing particularly in light of the evidence for progress around
block copolymer formation.
The processing of PHA polymers is also an area that
requires careful management, particularly considering the 12.2. Compositional distribution and mechanical
widely reported (although not necessarily intrinsic) [374] properties
thermal instability of such polymers. The use of reverse
temperature profile processing [375] is an important tech- The effect of compositional distribution on macro-
nique that is commonly used to minimize degradation with molecular structure and the subsequent effect on mechan-
melt extrusion. The unique rheological performance of PHA ical properties, secondary crystallization and other key
materials (which have very low viscosity at melt and thus polymer properties, particularly for high HV content mate-
can be used to produce articles with fine detail) is another rials, is not well understood, and it is clear that even for pure
important aspect that cannot be covered herein [365]. culture materials this is an important and largely unex-
Overall, it can be seen that a specific PHB or PHBV used plored parameter. Specific findings to date reveal that:
for one investigation can exhibit quite different properties
than that from another study, due to differences in ther- - Compositional distribution can be very broad in both pure
mal stability, impurities, blend composition or copolymer and mixed culture PHAs.
428 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

- Blends of PHBV copolymers of distinctly different com- feeding strategies and final mechanical properties for
position can be poorly miscible or immiscible; phase mixed culture PHA. It is important to demonstrate that the
separated domains can result and can lead to unexpected chemomechanical properties are predictable and may be
and/or uncontrolled changes in mechanical properties. controlled and within desirable limits for specific practical
- For phase separated blends the domain that crystallizes applications.
most rapidly apparently controls the final polymer prop- In many ways these questions are all interlinked and
erties. feed in to the full integration of process modeling with
product performance for reliable and cost effective PHA
Questions still arise in terms of the following effects: polymer production. Clearly there is great potential and
much growing anticipation for the broad application of PHA
- What impact does a broad compositional distribution materials.
have on end use properties and crystallization rates?
- How predictable are the properties once compositional Acknowledgments
distribution is characterized, and can properties be tailo-
red by appropriate blending post extraction? This research was supported under Australian Research
- How reproducible is the raw material? Council’s Linkage Projects funding scheme (project number
LP099091). The authors thank Paul Luckman for prepara-
12.3. Full characterization of mixed culture PHA tion of Figs. 1, 3 and 9, and Monica Arcos-Hernandez for
valuable discussions relating to PHA in general.
To date, there has been limited research into the link
between compositional distribution, process parameters,
Appendix A. Physical properties of mixed culture PHA

Culture enrichment PHA accumulation PHA composition PHA content Thermal properties Molecular Ref.
weight data

Substrate System Substrate VFA (g%) mol% (%) (Max) Tm(1) (◦ C) Tm(2) (◦ C) Total Tg1 (◦ C) Tg2 (◦ C) PDI Mw (×105 )
HAc:HPr:HBut: 3HB:3H2MB:3HV: Hm
HVal:other 3H2MV:3HHx (J g−1 )

Municipal wastewater AE EBPR Acetate 100:0:0:0:0 75:0:25:0:0 19.2 – – – – – 1.9 6.5 [348]
Municipal wastewater AE EBPR Propionate 0:100:0:0:0 28:0:72:0:0 14.9 – – – – – 2.1 6 [348]
Municipal wastewater AE EBPR Butyrate 0:0:100:0:0 60:0:40:0:0 11.4 – – – – – 1.6 4 [348]
Municipal wastewater AE SBR Butyrate 0:0:100:0:0 100:0:0:0:0 37 178 – – – – – – [376]

B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442


Municipal wastewater AE SBR Butyrate/valerate 0:0:80:20:0 88:0:12:0:0 40 144 – – – – – – [376]
Municipal wastewater AE SBR Butyrate/valerate 0:0:60:40:0 70:0:30:0:0 35 133 – – – – – – [376]
Municipal wastewater AE SBR Butyrate/valerate 0:0:40:60:0 65:0:35:0:0 24 127 – – – – – – [376]
Municipal wastewater AE SBR Butyrate/valerate 0:0:20:80:0 49:0:51:0:0 18 109 – – – – – – [376]
Municipal wastewater AE SBR Valerate 0:0:0:100:0 46:0:54:0:0 22 99 – – – – – – [376]
Acetate/propionate AE Acetate/propionate Unspecified 90:0:10:0:0 – 168 – 44.8 1.6 – 3.3 4 [46]
Acetate/propionate AE Acetate/propionate Unspecified 75:0:25:0:0 – 139 – 4.7 56 – 2.5 17 [46]
Acetate/propionate AE Acetate/propionate Unspecified 70:0:30:0:0 – 141 – 4.7 42 – 2.1 18 [46]
Acetate AE Acetate 100:0:0:0:0 100:0:0:0:0 – 145 – 40 −19 – 1.2 3.5 [46]
Acetate/propionate MAA/AE Acetate 100:0:0:0:0 58:0:42:0:0 35.6 75 – – – – – 3.23 [377]
(P
limited)
Acetate AN/AE Acetate (Ae) 100:0:0:0:0 93:0:7:0:0 41 – – – – – 2.7 1.9 [44]
Acetate AN/AE Acetate (Ae) 100:0:0:0:0 89:0:11:0:0 – – – – – – 2.6 1.1 [44]
Acetate AN/AE Acetate (Ae) 100:0:0:0:0 80:0:16:4:0 – 170 – 25.1 −3 – 2.2 3.1 [44]
Acetate AN/AE Acetate (Ae) 100:0:0:0:0 77:0:23:0:0 – – – – – – 3.7 3.2 [44]
Acetate AN/AE Acetate (An) 100:0:0:0:0 63:0:25:12:0 – 126 – 14.5 −4.8 – 1.6 3.9 [44]
Acetate AN/AE Acetate (An) 100:0:0:0:0 71:0:29:0:0 – 112 – 6.6 −3.7 – 1.4 4.1 [44]
Acetate AN/AE Acetate (An) 100:0:0:0:0 68:0:32:0:0 – 99 – 11.9 −6.6 – 1.4 5.5 [44]
Acetate AN/AE Acetate (An) 100:0:0:0:0 65:0:35:0:0 28 105 – 44.9 −9.7 – 5.6 2.9 [44]
Acetate AN/AE Acetate (An) 100:0:0:0:0 65:0:23:12:0 – 96 – 19.8 0.3 – 1.9 1.4 [44]
Acetate AN/AE Acetate (An) 100:0:0:0:0 70:0:30:0:0 31 106 – 37.0 −3.9 – 2.9 2.5 [44]
Acetate AN/AE Propionate 0:100:0:0:0 11:13:35:41:0 – 83.8 20 – −0.2 – 1.5 5.6 [73]
Acetate AN/AE Acetate/propionate 12.5:87.5:0:0:0 20:9:39:32:0 – 119 154.9 9.5 −6.1 – 3.2 4.1 [73]
Acetate AN/AE Acetate/propionate 36.4:63.6:0:0:0 46:8:32:15:0 – 70.3 96.5 18.5 −1.6 – 1.7 5 [73]
Acetate AN/AE Acetate/propionate 69.6:30.4:0:0:0 55:5:30:9:0 – 124.3 156.4 31.5 −8.1 – 3.1 3.9 [73]
Acetate AN/AE Acetate 100:0:0:0:0 66:4:28:2:0 – 96.4/109.1 145.5/161 9.1 −6.1 – 1.4 5.5 [73]
Acetate EA SBR Acetate/propionate 55:45:0:0:0 54:0:33:13:0 26 – – – – – 2.98 29.8 [378]
Acetate EA SBR Propionate 0:100:0:0:0 31:0:47:22:0 13.5 – – – – – 2.34 23.4 [378]
Acetate EA SBR Propionate 0:100:0:0:0 12:0:67:21:0 23.7 – – – – – 2.27 22.7 [378]
Acetate EA SBR Butyrate 0:0:100:0:0 100:0:0:0:0 14.8 – – – – – 2.79 27.9 [378]
Acetate EA SBR Valerate 0:0:0:100:0 32:0:52:16:0 14.6 – – – – – 2.32 23.2 [378]
Propionate EA SBR Propionate 0:100:0:0:0 28:0:72:0:0 20.8 99.2 – 8.9 −29.6 – 2.87 28.7 [378]

429
430
Appendix A (Continued)

Culture enrichment PHA accumulation PHA composition PHA content Thermal properties Molecular Ref.
weight data

Substrate System Substrate VFA (g%) mol% (%) (Max) Tm(1) (◦ C) Tm(2) (◦ C) Total Tg1 (◦ C) Tg2 (◦ C) PDI Mw (×105 )
HAc:HPr:HBut: 3HB:3H2MB:3HV: Hm
HVal:other 3H2MV:3HHx (J g−1 )

Acetate AE SBR Acetate 100:0:0:0:0 100:0:0:0:0 21 171 – 63.6 – – 1.3 31.6 [349]
(N
limited)
Acetate/propionate AE SBR Acetate/propionate 59:41:0:0:0 94:0:6:0:0 25 157 – 53.6 1 – 1.3 32.7 [349]
(N
limited)
Acetate AE SBR Mixed VFA – 31:18:29:22:0 61:0:39:0:0 77 113 138 – −16 – 2.3 2.15 [61]
simulated feed

B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442


Acetate AE SBR Mixed VFA – 60:16:20:4:0 79:0:21:0:0 68 121 137 – −10 – 2.7 3.86 [61]
simulated feed
Acetate AE SBR Fermented 60:9:25:6:0 85:0:15:0:0 56 134 147 – −1 – 2.3 6.46 [61]
molasses
Fermented molasses AN/AE Fermented 75:8:13:5:0 70:0:25:0:5 37 129.2 144.5 22.8 −7.2 0.5 2.1 4 [62]
molasses
Fermented molasses AN/AE Fermented 69:21:9:1:1 56:0:43:0:1 37.3 155 – 21.1 −5.5 1.2 1.8 3.5 [62]
molasses
Fermented molasses AN/AE Fermented 72:6:20:3:0 60:2:13:1:23 31.6 – – – −3.3 – 1.8 4.3 [62]
molasses
Fermented molasses AN/AE Fermented 71:7:17:3:2 47:0:53:0:0 – 153.7 165.3 10.2 −10.3 2.4 – – [62]
molasses
Fermented molasses AN/AE Acetate 100:0:0:0:0 90:4:4:1:1 29 170.9 – 77 4.6 – 2 8.1 [62]
Fermented molasses AN/AE Acetate 100:0:0:0:0 96:0:3:1:1 – 174 – 82.1 – – – – [62]
Fermented molasses AN/AE Acetate 100:0:0:0:0 89:0:11:0:0 – 164.6 171.9 77.5 4.8 – – – [62]
Fermented molasses AN/AE Propionate 0:100:0:0:0 12:6:63:14:6 15.8 89.2 – 17.6 −14.0 – 2 4.5 [62]
Fermented molasses AN/AE Butyrate 0:0:100:0:0 83:5:7:2:2 21.7 137.4 149.9 50.4 2.8 – 1.7 9 [62]
Fermented molasses AN/AE Valerate 0:0:0:100:0 12:5:78:4:1 24.8 98.1 – 1.5 −12.3 – 3.9 6.2 [62]
Fermented WAS 12 AE SBR Fermented WAS 42:18:17:3:20 82:0:18:0:0 8 131 143 42 0.7 – 2.6 6.1 [379]
gCOD/L/day
Fermented WAS 12 AE SBR Mixed VFA – 43:16:14:6:21 87:0:13:0:0 29 131 144 46 0.1 – 2.5 8.4 [379]
gCOD/L/day simulated feed
Fermented WAS 12 AE SBR Mixed VFA – 43:16:14:6:21 85:0:15:0:0 13 136 148 36 1.9 – 2.1 9.2 [379]
gCOD/L/day simulated feed
Fermented WAS 6 AE SBR Fermented WAS 42:18:17:3:20 74:0:26:0:0 19 138 147 25 −0.9 – 1.8 8.2 [379]
gCOD/L/day
Fermented WAS 6 AE SBR Mixed VFA – 43:16:14:6:21 71:0:29:0:0 23 140 155 16 −0.9 – 1.5 10.4 [379]
gCOD/L/day simulated feed
Fermented WAS 6 AE SBR Mixed VFA – 43:16:14:6:21 56:0:44:0:0 23 139 151 32 −13.2 2.2 2.5 8.8 [379]
gCOD/L/day simulated feed
Fermented WAS 6 AE SBR Acetate/propionate 77:23:0:0:0 72:0:28:0:0 23 146 – 17 −1.2 – 2.2 7.3 [379]
gCOD/L/day
Fermented WAS 6 AE SBR Acetate/propionate 77:23:0:0:0 73:0:27:0:0 25 142 – 14 −1.4 – 2.4 8.7 [379]
gCOD/L/day
Glycerol AE Glycerol Glycerol 100:0:0:0:0 10–35 158–175 – 85–95 −5 to 5 – – 2.0–3.8 [351]
AE = aerobic; AN = anaerobic; COD = chemical oxygen demand; EBPR = enhanced biological phosphorus removal; HAc = acetate; HPr = propionate; HBut = butyrate; HVal = valerate; MAA = microaerophilic;
SBR = sequencing batch reactor.
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 431

Appendix B. Effect of polymer composition on mechanical properties of PHA – data sampled from the literature

Sample type Aging time mol% Tm (◦ C) Tg (◦ C) Young’s Stress at Tensile Elongation Ref.
comonomer modulus yield (MPa) strength at break (%)
unit (GPa) (MPa)

P(3HB)
Melt pressed sheet 1 week 0 1.5–1.8 8–20 0.8 [358]
Solvent cast film Unspecified 0 30 4 [221]
Solvent cast film 4 weeks 0 178.2 2.4 1.5 1.93 [380]
Injection molded 150 days (1% boron nitride) 0 3.5 7 [381]
Solvent cast film Unspecified 0 177 2 43 5 [24]
Solvent cast film Unspecified 0 177 4 43 5 [382]
Solvent cast film 6 days dried, hot pressed 140 ◦ C 0 172 ∼0 1.2 25 [383]
Melt pressed sheet 3 weeks 0 181 17.7 2.99 36.4 2.1 [384]
Melt pressed sheet Small amount of talc, 60 ◦ C 24 h 0 179 10 3.5 40 [22]
Unspecified Unspecified 0 175 −4 4 40 6 [385]
Unspecified Unspecified 0 175 15 40 6 [386]
Unspecified Unspecified 0 177 3.5 40 8 [264]
Solvent cast film 60 ◦ C 20 h 0 175 9 3.6 44 3 [387]
Solvent cast film 60 days 0 32.7 3.3 [388]
Solvent cast film Unspecified 0 162 −1.2 1.51 18.5 4.45 [346]
Solvent cast film “Several days” 0 2.1 28 3.6 [389]
P(3HB-co-3HV)
Film, unspecified Unspecified 0.05 172 3.0 6.0 [199]
preparation
Melt pressed sheet 24 ◦ C 1 day 1.1 158 3.65 19.7 0.17 [390]
Melt sheet Small amount of talc, 60 ◦ C 24 h 3 170 8 2.9 38 [22]
Solvent cast film 40 ◦ C, >1 day, high vacuum 3.6 169.6 3.18 50.5 1.91 [391]
Solvent cast film 25 ◦ C 2 weeks; 60 ◦ C 2 weeks 5 166 2 [392]
under vacuum
Melt pressed sheet 25 ◦ C 2 weeks; 60 ◦ C 2 weeks 5 171 −0.4 13 22.1 1.39 [392]
under vacuum
Melt pressed sheet Unspecified 5 166 1 1.4 19 2.4 [393]
Melt pressed sheet 50 ◦ C 2 days; 25 ◦ C 2 weeks 5 171 0.95 34.5 4 [394]
Solvent cast film Unspecified 5 170.1 2.3 1.47 18.09 2.31 [395]
Solvent cast film Unspecified 5 156 −1.7 2.4 123 [346]
Solvent cast film 40 ◦ C, 2 days 6.6 1.37 26.9 4.1 [396]
Solvent cast film Unspecified 7 0.124 3.0 [397]
Injection molded 90 ◦ C 12 h 8 153 3.5 27.1 3.6 [398]
dogbones
Melt pressed sheet 145 ◦ C 1 h then slow cooled 8 1.06 64 [344]
Solvent cast film 224 days 8 164 −0.6 21.5 16 [388]
Solvent cast film Unspecified 8 151 10 0.187 19 15 [297]
Melt pressed sheet Small amount of talc, 60 ◦ C 24 h 9 162 6 1.9 37 [22]
Unspecified Unspecified 10 140 1.2 25 20 [264]
Solvent cast film 224 days 10 163.2 1.7 19.3 25 [388]
Solvent cast film Unspecified 10 145 1.5 27 1 [340]
Solvent cast film 60 ◦ C 20 h 11.3 157 2 3.2 38 5 [387]
Melt pressed sheet 25 ◦ C 2 weeks; 60 ◦ C 2 weeks 12 143 −2.7 8.67 20.9 4.49 [392]
under vacuum
Melt pressed sheet Unspecified 13 157 2.9 1.07 12.7 [399]
Injection molded Unspecified 13 157.3 0.3 1 26 [400]
Solvent cast film 224 days 13 161 −0.9 18.8 48 [388]
Solvent cast film Unspecified 14 0.082 1.03 [397]
Melt pressed sheet Small amount of talc, 60 ◦ C 24 h 14 150 4 1.5 35 [22]
Solvent cast film 60 ◦ C 20 h 15.5 140 −1 2.5 34 30 12 [387]
Solvent cast film Unspecified 17 21 203 [221]
Solvent cast film 224 days 17 124.4 −4.4 18.2 43 [388]
Spin cast films 40 ◦ C, 4 days 18 158/172 13 1.12 25.1 4.03 [401]
Solvent cast film 224 days 18 123 −1.9 20.2 164 [388]
Melt pressed sheet 25 ◦ C 2 weeks; 60 ◦ C 2 weeks 20 132 −0.05 8.37 15.2 1.26 [392]
under vacuum
Solvent cast film 3 weeks 20 145 −1 2.9 50 [24]
Melt pressed sheet Small amount of talc, 60 ◦ C 24 h 20 145 −1 1.2 32 [22]
Unspecified Unspecified 20 130 0.8 20 50 [264]
Solvent cast film 224 days 20 116.4 −6.3 15.6 182 [388]
Solvent cast film 60 ◦ C 20 h 20.1 114 −5 1.9 26 23 27 [387]
Solvent cast film Unspecified 22 0.09 2.29 [397]
Film, unspecified Unspecified 22.6 97 −6.4 4.0 [199]
preparation

Melt pressed sheet Small amount of talc, 60 C 24 h 25 137 −6 0.7 30 [22]
Solvent cast film 60 ◦ C 20 h 28.4 102 −8 1.4 22 30 700 [387]
432 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

Sample type Aging time mol% Tm (◦ C) Tg (◦ C) Young’s Stress at Tensile Elongation Ref.
comonomer modulus yield (MPa) strength at break (%)
unit (GPa) (MPa)

Solvent cast film Unspecified 29 87 −10 0.144 15 900 [297]


Solvent cast film Unspecified 34 97 −8 13 18 970 [24]
Melt pressed sheet Unspecified 36 166 ∼0 0.5 16.6 70 [383]
Film, unspecified Unspecified 39.2 70 −5.8 3.2 [199]
preparation
Melt pressed sheet Unspecified 50 162 ∼0 0.41 13.4 230 [383]
Film, unspecified Unspecified 54.1 64 −6.4 2.6 [199]
preparation
Solvent cast film Unspecified 55 77 −10 12 16 >1200 [24]
Film, unspecified Unspecified 69.1 70 −3.9 1.8 [199]
preparation
Solvent cast film Unspecified 71 83 −13 11 5 [24]
Solvent cast film Unspecified 25–30 18 120 [221]
P(3HB-co-4HB)
Films (solvent cast) 3 weeks 3 166 28 45 [325]
Films (solvent cast) 4 weeks 5 168.5 −2 1.23 10.7 [380]
Films (solvent cast) 3 weeks 10 159 24 242 [325]
Films (solvent cast) 3 weeks 16 150 −7 26 444 [325]
Films (solvent cast) 4 weeks 24 161.1 −5 0.79 22.2 [380]
Films (solvent cast) 4 weeks 38 152.2 −10 0.66 48 [380]
Films (solvent cast) 3 weeks 44 10 511 [325]
Films (solvent cast) 3 weeks 64 50 −35 0.03 17 591 [402]
Films (solvent cast) 3 weeks 78 49 −37 0.024 42 1120 [402]
Films (solvent cast) 3 weeks 82 52 −39 0.045 58 1320 [402]
Films (solvent cast) 3 weeks 90 50 −42 0.1 65 1080 [402]
Films (solvent cast) Unspecified 94 51 −46 0.055 39 500 [403]
Films (solvent cast) 3 weeks 100 53 −48 0.149 104 1000 [402]
P(3HB-co-3H4MV)
Films (solvent cast) 180 days 7 149, 162 −2 18 19 [404]
Films (solvent cast) 180 days 9 146, 159 −2 25 22 [404]
Films (solvent cast) 180 days 14 136, 148 −2 19 110 [404]
Films (solvent cast) 180 days 16 133, 143 −2 17 260 [404]
Films (solvent cast) 180 days 22 126 −2 14 440 [404]
P(3HB-co-3HHx)
Melt pressed sheet Unspecified 2.5 142 0.631 25.7 6.7 [324]
Melt pressed sheet 3 months 2.5 142 5 [324]
Melt pressed sheet 3 months 4.5 22 [324]
Melt pressed sheet Unspecified 4.6 130 0.6 22.9 6.5 [324]
Melt pressed sheet 3 months 4.7 3 [324]
Melt pressed sheet Unspecified 5.4 0.494 23.9 17.6 [324]
Melt pressed sheet 3 months 5.4 130 10 [324]
Melt pressed sheet Unspecified 5.9 130 0.412 17.7 163 [324]
Melt pressed sheet Unspecified 6.3 0.343 17.9 24 [324]
Melt pressed sheet 3 months 6.4 128 6 [324]
Melt pressed sheet Unspecified 7 0.289 17.3 23.6 [324]
Melt pressed sheet 3 months 7.6 119 12 [324]
Melt pressed sheet 3 months 8.1 119 12 [324]
Melt pressed sheet Unspecified 8.4 0.989 14.5 11.9 [324]
Melt pressed sheet Unspecified 8.5 0.249 17.8 25.1 [324]
Melt pressed sheet Unspecified 8.5 0.232 15.6 34.3 [324]
Melt pressed sheet Unspecified 9.5 0.155 8.8 43 [324]
Melt pressed sheet Unspecified 10 ∼0.52 [38]
Films (solvent cast) Unspecified 10 127 −1 21 400 [382]
Films (solvent cast) Unspecified 12 97 −1.8 0.135 4.45 107.7 [346]
Films (solvent cast) Unspecified 12 96.2 −1.2 0.498 9.36 112.9 [395]
Films (solvent cast) Unspecified 15 115 0 23 760 [382]
Films (solvent cast) Unspecified 17 120 −2 20 850 [382]
Melt pressed sheet Unspecified ∼12 106.3 −0.2 0.267 5.99 11.5 [405]
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 433

References [26] Anderson AJ, Haywood GW, Dawes EA. Biosynthesis and composi-
tion of bacterial poly(hydroxyalkanoates). International Journal of
Biological Macromolecules 1990;12:102–5.
[1] Rehm BHA. Polyester synthases: natural catalysts for plastics. Bio-
[27] Lemoigne M. Production d’acide ␤-oxybutyrique par certaines bac-
chemical Journal 2003;376:15–33.
téries du groupe du B. subtilis. Comptes Rendus de l’Académie des
[2] Anderson AJ, Dawes EA. Occurrence, metabolism, metabolic role,
Sciences 1923;176:1761–3.
and industrial uses of bacterial polyhydroxyalkanoates. Microbio-
[28] Lemoigne M. Sur le mécanisme de la production de l’acide ␤-
logical Reviews 1990;54:450–72.
oxybutyrique par voie biochimique. Comptes Rendus de l’Académie
[3] Pötter M, Steinbüchel A. Poly(3-hydroxybutyrate) granule-
des Sciences 1924;178:1093–5.
associated proteins: impacts on poly(3-hydroxybutyrate)
[29] Lemoigne M. Sur l’origine de l’acide ␤-oxybutyrique obtenu par
synthesis and degradation. Biomacromolecules 2005;6:
processus microbien. Comptes Rendus de l’Académie des Sciences
552–60.
1925;180:1539–41.
[4] Steinbüchel A, Valentin HE. Diversity of bacterial polyhydroxyalka-
[30] Lemoigne M. Études sur I’autolyse microbienne. Acidification par
noic acids. FEMS Microbiology Letters 1995;128:219–28.
formation d’acide ␤-oxybutyrique. Annales de l’Institut Pasteur
[5] Steinbüchel A, Lutke-Eversloh T. Metabolic engineering and path-
Paris 1925;39:l44–173.
way construction for biotechnological production of relevant
[31] Lemoigne M. Origine de l’acide 3-oxo-butyrique obtenue par autol-
polyhydroxyalkanoates in microorganisms. Biochemical Engineer-
yse microbienne. Comptes Rendus des Seances de la Societe de
ing Journal 2003;16:81–96.
Biologie et de Ses Filiales 1926;95:1359–60.
[6] Kim YB, Lenz RW. Polyesters from microorganisms. Advances in
[32] Lemoigne M. Produit de déshydratation et de polymérisation de
Biochemical Engineering/Biotechnology 2001;71:51–79.
l’acide ␤-oxybutyrique. Bulletin de la Societe de Chimie Biologique
[7] Lenz RW, Marchessault RH. Bacterial polyesters: biosynthesis,
1926;8:770–82.
biodegradable plastics and biotechnology. Biomacromolecules
[33] Lemoigne M. Études sur l’autolyse microbienne. Origine de l’acide
2005;6:1–8.
␤-oxybutyrique formé par autolyse. Annales de l’Institut Pasteur
[8] Hazer B, Steinbüchel A. Increased diversification of polyhy-
Paris 1927;41:148–65.
droxyalkanoates by modification reactions for industrial and
[34] Williamson DH, Wilkinson JF. The isolation and estimation of the
medical applications. Applied Microbiology and Biotechnology
poly-␤-hydroxybutyrate inclusions of bacillus species. Journal of
2007;74:1–12.
General Microbiology 1958;19:198–209.
[9] Sudesh K, Abe H. Practical guide to microbial polyhydroxyalka-
[35] Doudoroff M, Stanier RY. Role of poly-␤-hydroxybutyric acid in
noates. Shrewsbury, UK: Smithers Rapra Technology; 2010, 160
the assimilation of organic carbon by bacteria. Nature 1959;183:
pp.
1440–2.
[10] Rehm BHA. Bacterial polymers: biosynthesis, modifications and
[36] Baptist JN, Ziegler JB. Method of making absorbable surgical sutures
applications. Nature Reviews Microbiology 2010;8:578–92.
from poly ␤ hydroxy acids. US Patent 3,225,766 (A), W R Grace &
[11] Chanprateep S. Current trends in biodegradable polyhy-
Co.; 1965.
droxyalkanoates. Journal of Bioscience and Bioengineering
[37] Holmes PA. Applications of PHB – a microbially produced
2010;110:621–32.
biodegradable thermoplastic. Physics in Technology 1985;16:32–6.
[12] Chen G. Plastics completely synthesized by bacteria: polyhydrox-
[38] Noda I, Green PR, Satkowski MM, Schechtman LA. Preparation and
yalkanoates. In: Chen G, editor. Plastics from bacteria: natural
properties of a novel class of polyhydroxyalkanoate copolymers.
functions and applications. Berlin; Heidelberg: Springer-Verlag;
Biomacromolecules 2005;6:580–6.
2010. p. 17–37.
[39] Byrom D. Polyhydroxyalkanoates. In: Mobley DP, editor. Plastics
[13] Verlinden RAJ, Hill DJ, Kenward MA, Williams CD, Radecka I. Bacte-
from microbes: microbial synthesis of polymers and polymer pre-
rial synthesis of biodegradable polyhydroxyalkanoates. Journal of
cursors. Munich: Hanser; 1994. p. 5–33.
Applied Microbiology 2007;102:1437–49.
[40] Dias JML, Lemos PC, Serafim LS, Oliveira C, Eiroa M, Albuquerque
[14] Pan P, Inoue Y. Polymorphism and isomorphism in biodegradable
MGE, Ramos AM, Oliveira R, Reis MA. Recent advances in poly-
polyesters. Progress in Polymer Science 2009;34:605–40.
hydroxyalkanoate production by mixed aerobic cultures: from
[15] Marchessault RH. Polyhydroxyalkanoate (PHA) history at Syracuse
the substrate to the final product. Macromolecular Bioscience
University and beyond. Cellulose 2009;16:357–9.
2006;6:885–906.
[16] Orts WJ, Nobes GAR, Kawada J, Nguyen S, Yu G-E, Raveneile F.
[41] Serafim LS, Lemos PC, Albuquerque MGE, Reis MAM. Strategies for
Poly(hydroxyalkanoates): biorefinery polymers with a whole range
PHA production by mixed cultures and renewable waste materials.
of applications. The work of Robert H. Marchessault. Canadian Jour-
Applied Microbiology and Biotechnology 2008;81:615–28.
nal of Chemistry 2008;86:628–40.
[42] Albuquerque MGE, Eiroa M, Torres C, Nunes BR, Reis MAM.
[17] Castilho LR, Mitchel DA, Freire DMG. Production of polyhydrox-
Strategies for the development of a side stream process for poly-
yalkanoates (PHAs) from waste materials and by-products by
hydroxyalkanoate (PHA) production from sugar cane molasses.
submerged and solid-state fermentation. Bioresource Technology
Journal of Biotechnology 2007;130:411–21.
2009;100:5996–6009.
[43] Bengtsson S, Werker A, Christensson M, Welander T. Production of
[18] Khanna S, Srivastava AK. Recent advances in microbial polyhydrox-
polyhydroxyalkanoates by activated sludge treating a paper mill
yalkanoates. Process Biochemistry 2005;40:607–19.
wastewater. Bioresource Technology 2008;99:509–16.
[19] Sudesh K, Abe H, Doi Y. Synthesis, structure and properties of
[44] Dai Y, Yuan ZG, Jack K, Keller J. Production of targeted
polyhydroxyalkanoates: biological polyesters. Progress in Polymer
poly(3-hydroxyalkanoates) copolymers by glycogen accumulating
Science 2000;25:1503–55.
organisms using acetate as sole carbon source. Journal of Biotech-
[20] Braunegg G, Lefebvre G, Genser KF. Polyhydroxyalkanoates,
nology 2007;129:489–97.
biopolyesters from renewable resources: physiological and engi-
[45] Kleerebezem R, van Loosdrecht MCM. Mixed culture biotechnol-
neering aspects. Journal of Biotechnology 1998;65:127–61.
ogy for bioenergy production. Current Opinion in Biotechnology
[21] Amass W, Amass A, Tighe B. A review of biodegradable polymers:
2007;18:207–12.
uses, current developments in the synthesis and characterization
[46] Reis MAM, Serafim LS, Lemos PC, Ramos AM, Aguiar FR,
of biodegradable polyesters, blends of biodegradable polymers and
van Loosdrecht MCM. Production of polyhydroxyalkanoates by
recent advances in biodegradation studies. Polymer International
mixed microbial cultures. Bioprocess and Biosystems Engineering
1998;47:89–144.
2003;25:377–85.
[22] Holmes PA. Biologically produced (R)-3-hydroxyalkanoate poly-
[47] Gurieff N, Lant P. Comparative life cycle assessment and finan-
mers and copolymers. In: Bassett OC, editor. Developments in
cial analysis of mixed culture polyhydroxyalkanoate production.
crystalline polymers – 2. Essex, UK: Elsevier Applied Science Pub-
Bioresource Technology 2007;98:3393–403.
lishers Ltd.; 1988. p. 1–65.
[48] Wallen LL, Rohwedder WK. Poly-␤-hydroxyalkanoate from
[23] Kessler B, Witholt B. Synthesis, recovery and possible application of
activated sludge. Environmental Science and Technology
medium-chain-length polyhydroxyalkanoates: a short overview.
1974;8:576–9.
Macromolecular Symposium 1998;130:245–60.
[49] Daigger GT, Grady CPL. The dynamics of microbial growth
[24] Doi Y. Microbial polyesters. New York: VCH Publishers; 1990, 166
on soluble substrates – a unifying theory. Water Research
pp.
1982;16:365–82.
[25] Sudesh K, Doi A. Polyhydroxyalkanoates. In: Bastioli C, editor.
[50] Salehizadeh H, Van Loosdrecht MCM. Production of polyhydrox-
Handbook of biodegradable polymers. United Kingdom: RAPRA
yalkanoates by mixed culture: recent trends and biotechnological
Technology Ltd.; 2005. p. 219–56.
importance. Biotechnology Advances 2004;22:261–79.
434 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

[51] Beccari M, Majone M, Massanisso P, Ramadori R. A bulking sludge [72] Beccari M, Bertin L, Dionisi D, Fava F, Lampis S, Majone M, Valentino
with high storage response selected under intermittent feeding. F, Vallini G, Villano M. Exploiting olive oil mill effluents as a renew-
Water Research 1998;32:3403–13. able resource for production of biodegradable polymers through a
[52] Beun JJ, Dircks K, Heijnen JJ, van Loosdrecht MCM. Poly-␤- combined anaerobic–aerobic process. Journal of Chemical Technol-
hydroxybutyrate metabolism in dynamically fed mixed microbial ogy and Biotechnology 2009;84:901–8.
cultures. Water Research 2002;36:1167–80. [73] Dai Y, Lambert L, Yuan Z, Keller J. Characterisation of poly-
[53] Bengtsson S, Pisco AR, Reis MAM, Lemos PC. Production of polyhy- hydroxyalkanoate copolymers with controllable four-monomer
droxyalkanoates from fermented sugar cane molasses by a mixed composition. Journal of Biotechnology 2008;134:137–45.
culture enriched in glycogen accumulating organisms. Journal of [74] Dai Y, Lambert L, Yuan ZG, Keller J. Microstructure of copolymers of
Biotechnology 2010;145:253–63. polyhydroxyalkanoates produced by glycogen accumulating orga-
[54] Pisco AR, Bengtsson S, Werker A, Reis MAM, Lemos PC. nisms with acetate as the sole carbon source. Process Biochemistry
Community structure evolution and enrichment of glycogen- 2008;43:968–77.
accumulating organisms producing polyhydroxyalkanoates from [75] Takabatake H, Satoh H, Mino T, Matsuo T. Recovery of biodegradable
fermented molasses. Applied and Environmental Microbiology plastics from activated sludge process. Water Science and Technol-
2009;75:4676–86. ogy 2000;42:351–6.
[55] Dionisi D, Majone M, Papa V, Beccari M. Biodegradable poly- [76] Lee WH, Loo CY, Nomura CT, Sudesh K. Biosynthesis of poly-
mers from organic acids by using activated sludge enriched hydroxyalkanoate copolymers from mixtures of plant oils
by aerobic periodic feeding. Biotechnology and Bioengineering and 3-hydroxyvalerate precursors. Bioresource Technology
2004;85:569–79. 2008;99:6844–51.
[56] Beccari M, Dionisi D, Giuliani A, Majone M, Ramadori R. Effect of [77] Johnson K, Jiang Y, Kleerebezem R, Muyzer G, van Loosdrecht MCM.
different carbon sources on aerobic storage by activated sludge. Enrichment of a mixed bacterial culture with a high polyhydrox-
Water Science and Technology 2002;45:157–68. yalkanoate storage capacity. Biomacromolecules 2009;10:670–6.
[57] Serafim LS, Lemos PC, Oliveira R, Reis MAM. Optimization of polyhy- [78] Johnson K, van Loosdrecht MCM, Kleerebezem R. Influence
droxybutyrate production by mixed cultures submitted to aerobic of ammonium on the accumulation of polyhydroxybutyrate
dynamic feeding conditions. Biotechnology and Bioengineering (P(3HB)) in aerobic open mixed cultures. Journal of Biotechnology
2004;87:145–60. 2010;147:73–9.
[58] Lemos PC, Serafim LS, Reis MAM. Synthesis of polyhydroxyalka- [79] Albuquerque MGE, Martino V, Pollet E, Averous L, Reis MAM. Mixed
noates from different short-chain fatty acids by mixed cultures culture polyhydroxyalkanoate (PHA) production from volatile fatty
submitted to aerobic dynamic feeding. Journal of Biotechnology acid (VFA)-rich streams: effect of substrate composition and feed-
2006;122:226–38. ing regime on PHA productivity, composition and properties.
[59] Solaiman DKY, Ashby RD, Foglia TA, Marmer WM. Con- Journal of Biotechnology 2011;151:66–76.
version of agricultural feedstock and coproducts into [80] Villano M, Beccari M, Dionisi D, Lampis S, Miccheli A, Vallini G.
poly(hydroxyalkanoates). Applied Microbiology and Biotechnology Effect of pH on the production of bacterial polyhydroxyalkanoates
2006;71:783–9. by mixed cultures enriched under periodic feeding. Process Bio-
[60] Albuquerque MGE, Bengtsson S, Martino V, Pollet E, Reis MAM. Eco- chemistry 2010;45:714–23.
engineering of mixed microbial cultures to develop a cost-effective [81] Ivanova G, Serafim LS, Lemos PC, Ramos AM, Reis MAM, Cabrita
bioplastic (PHA) production process from a surplus feedstock – EJ. Influence of feeding strategies of mixed microbial cultures
sugar molasses. Journal of Biotechnology 2010;150:S70. on the chemical composition and microstructure of copolyesters
[61] Albuquerque MGE, Martino V, Pollet E, Avérous L, Reis MAM. Mixed P(3HB-co-3HV) analyzed by NMR and statistical analysis. Magnetic
culture polyhydroxyalkanoate (PHA) production from volatile fatty Resonance in Chemistry 2009;47:497–504.
acid (VFA)-rich streams: effect of substrate composition and feed- [82] van Beilen JB, Poirier Y. Production of renewable polymers from
ing regime on PHA productivity, composition and properties. crop plants. Plant Journal 2008;54:684–701.
Journal of Biotechnology 2011;151:66–76. [83] Snell KD, Peoples OP. PHA bioplastic: a value-added coproduct
[62] Bengtsson S, Pisco AR, Johansson P, Lemos PC, Reis MAM. Molecular for biomass biorefineries. Biofuels, Bioproducts and Biorefining
weight and thermal properties of polyhydroxyalkanoates produced 2009;3:456–67.
from fermented sugar molasses by open mixed cultures. Journal of [84] Poirier Y, Dennis DE, Klomparens K, Somerville C. Polyhydroxy-
Biotechnology 2010;147:172–9. butyrate, a biodegradable thermoplastic, produced in transgenic
[63] Md Din MF, Ujang Z, Van Loosdrecht M, Ahmad M. Polyhy- plants. Science 1992;256:520–3.
droxyalkanoates (PHAs) production from saponified sunflower [85] Bohmert-Tatarev K, McAvoy S, Daughtry S, Peoples OP, Snell
oil in mixed cultures under aerobic condition. Jurnal Teknologi KD. High levels of bioplastic are produced in fertile transplas-
2008;48:1–19. tomic tobacco plants engineered with a synthetic operon
[64] Gurieff N. Production of biodegradable polyhydroxyalkanoate poly- for the production of polyhydroxybutyrate. Plant Physiology
mers using advanced biological wastewater treatment process 2011;155:1690–708.
technology. PhD Dissertation. School of Engineering, Brisbane: The [86] Somleva MN, Snell KD, Beaulieu JJ, Peoples OP, Garrison BR, Pat-
University of Queensland; 2007. 160 pp. terson NA. Production of polyhydroxybutyrate in switchgrass, a
[65] Liu HY, Hall PV, Darby JL, Coats ER, Green PG, Thompson DE, value-added co-product in an important lignocellulosic biomass
Loge FJ. Production of polyhydroxyalkanoate during treatment crop. Plant Biotechnology Journal 2008;6:663–78.
of tomato cannery wastewater. Water Environment Research [87] Hempel F, Bozarth AF, Lindenkamp N, Klingl A, Zauner S, Linne U,
2008;80:367–72. Steinbüchel A, Maier UG. Microalgae as bioreactors for bioplastic
[66] Cai MM, Chua H, Zhao QL, Shirley SN, Ren J. Optimal production production. Microbial Cell Factories 2011;10:81–6.
of polyhydroxyalkanoates (PHA) in activated sludge fed by volatile [88] Abe H, Doi Y, Satkowski MM, Noda I. Miscibility and morphology
fatty acids (VFAs) generated from alkaline excess sludge fermenta- of blends of isotactic and atactic poly(3-hydroxybutyrate). Macro-
tion. Bioresource Technology 2009;100:1399–405. molecules 1994;27:50–4.
[67] Coats ER, Loge FJ, Wolcott MP, Englund K, McDonald AG. Synthe- [89] Abe H, Matsubara I, Doi Y. Physical properties
sis of polyhydroxyalkanoates in municipal wastewater treatment. and enzymatic degradability of polymer blends of
Water Environment Research 2007;79:2396–403. bacterial poly[(R)-3-hydroxybutyrate] and poly[(R,S)-
[68] Dionisi D, Majone M, Levantesi C, Di Gregorio S, Beccari M. Effect 3-hydroxybutyrate]stereoisomers. Macromolecules
of feed length on settleability, substrate uptake and storage in a 1995;28:844–53.
sequencing batch reactor treating an industrial wastewater. Envi- [90] Madison LL, Huisman GW. Metabolic engineering of poly(3-
ronmental Technology 2006;27:901–8. hydroxyalkanoates): from DNA to plastic. Microbiology and
[69] Carucci A, Dionisi D, Majone M, Rolle E, Smurra P. Aerobic stor- Molecular Biology Reviews 1999;63:21–53.
age by activated sludge on real wastewater. Water Research [91] Peoples OP, Sinskey AJ. Poly-␤-hydroxybutyrate biosynthesis in
2001;35:3833–44. Alcaligenes-Eutrophus H16 – characterization of the genes encoding
[70] Moita R, Lemos PC. Biopolymers production from mixed cul- ␤-ketothiolase and acetoacetyl-CoA reductase. Journal of Biological
tures and pyrolysis by-products. Journal of Biotechnology Chemistry 1989;264:15293–7.
2012;157:578–83. [92] Lawrence AG, Choi JH, Rha CK, Stubbe J, Sinskey AJ. In vitro analysis
[71] Dionisi D, Carucci G, Papini MP, Riccardi C, Majone M, Carrasco F. of the chain termination reaction in the synthesis of poly-(R)-␤-
Olive oil mill effluents as a feedstock for production of biodegrad- hydroxybutyrate by the class III synthase from Allochromatium
able polymers. Water Research 2005;39:2076–84. vinosum. Biomacromolecules 2005;6:2113–9.
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 435

[93] Rehm BHA, Steinbüchel A. Biochemical and genetic analysis of PHA [113] McCool GJ, Fernandez T, Li N, Cannon MC. Polyhydroxyalkanoate
synthases and other proteins required for PHA synthesis. Interna- inclusion-body growth and proliferation in Bacillus megaterium.
tional Journal of Biological Macromolecules 1999;25:3–19. FEMS Microbiology Letters 1996;138:41–8.
[94] Grage K, Jahns AC, Parlane N, Palanisamy R, Rasiah IA, Atwood [114] Jendrossek D. Polyhydroxyalkanoate granules are complex sub-
JA, Rehm BHA. Bacterial polyhydroxyalkanoate granules: bio- cellular organelles (carbonosomes). Journal of Bacteriology
genesis, structure, and potential use as nano-/micro-beads in 2009;191:3195–202.
biotechnological and biomedical applications. Biomacromolecules [115] Stubbe J, Tian J. Polyhydroxyalkanoate (PHA) homeostasis: the role
2009;10:660–9. of the PHA synthase. Natural Products Reports 2003;20:445–57.
[95] Ratledge C, Kristiansen B. Basic biotechnology. Cambridge, UK: [116] Cho M, Brigham CJ, Sinskey AJ, Stubbe J. Purification of polyhydrox-
Cambridge University Press; 2001, 584 pp. ybutyrate synthase from its native organism, Ralstonia eutropha:
[96] Henderson RA, Jones CW. Poly-3-hydroxybutyrate production by implications for the initiation and elongation of polymer formation
washed cells of Alcaligenes eutrophus; purification, characterisa- in vivo. Biochemistry 2012;51:2276–88.
tion and potential regulatory role of citrate synthase. Archives of [117] Rehm BHA, Qi QS, Beermann BB, Hinz HJ, Steinbüchel A. Matrix-
Microbiology 1997;168:486–92. assisted in vitro refolding of Pseudomonas aeruginosa class II
[97] Chowdhury R. Poly-␤-hydroxybuttersaüre abbauende bakterien polyhydroxyalkanoate synthase from inclusion bodies produced in
und exoenzyme. Archiv fur Mikrobiologie 1963;47:167–200. recombinant Escherichia coli. Biochemical Journal 2001;358:263–8.
[98] Nuti MP, Bertoldi Md, Lepidi AA. Influence of phenylacetic acid on [118] Jendrossek D. Fluorescence microscopical investigation of poly(3-
poly-␤-hydroxybutyrate (PHB) polymerization and cell elongation hydroxybutyrate) granule formation in bacteria. Biomacro-
in Azotobacter chroococcum Beij. Canadian Journal of Microbiology molecules 2005;6:598–603.
1972;18:1257–61. [119] Gerngross TU, Reilly P, Stubbe J, Peoples OP. Immunocytochemical
[99] Pötter M, Madkour MH, Mayer F, Steinbüchel A. Regulation analysis of poly-␤-hydroxybutyrate (PHB) synthase in Alcaligenes-
of phasin expression and polyhydroxyalkanoate (PHA) gran- Eutrophus H16 – localization of the synthase enzyme at the surface
ule formation in Ralstonia eutropha H16. Microbiology-Sgm of PHB granules. Journal of Bacteriology 1993;175:5289–93.
2002;148:2413–26. [120] Tian JM, Sinskey AJ, Stubbe J. Kinetic studies of polyhy-
[100] Ballard DGH, Holmes PA, Senior PJ. Formation of polymers of droxybutyrate granule formation in Wautersia eutropha H16
␤-hydroxybutyric acid in bacterial cells and a comparison of by transmission electron microscopy. Journal of Bacteriology
the morphology of growth with the formation of polyethylene 2005;187:3814–24.
in the solid state. In: Fontanille M, Guyot A, editors. Recent [121] Mumtaz T, Abd-Aziz S, Rahman NA, Yee PL, Shirai Y, Hassan MA.
advances in mechanistic and synthetic aspects of polymeriza- Visualization of core–shell P(3HB)V granules of wild type Coma-
tion, vol. 215. Bandol, France: NATO Asi Science Series C; 1987. monas sp. EB172 in vivo under transmission electron microscope.
p. 293–314. International Journal of Polymer Analysis and Characterization
[101] Steinbüchel A, Aerts K, Babel W, Follner C, Liebergesell M, Madk- 2011;16:228–38.
our MH, Wieczorek R, Mayer F, Pieper-Fürst U, Pries A, Valentin HE. [122] Muh U, Sinskey AJ, Kirby DP, Lane WS, Stubbe JA. PHA synthase from
Considerations on the structure and biochemistry of bacterial poly- Chromatium vinosum: cysteine 149 is involved in covalent catalysis.
hydroxyalkanoic acid inclusions. Canadian Journal of Microbiology Biochemistry 1999;38:826–37.
1995;41(13):94–105. [123] Zhang SM, Yasuo T, Lenz RW, Goodwin S. Kinetic and mechanistic
[102] Jendrossek D, Selchow O, Hoppert M. PHB granules at the characterization of the polyhydroxybutyrate synthase from Ralsto-
early stages of formation are localized close to the cytoplas- nia eutropha. Biomacromolecules 2000;1:244–51.
mic membrane in Caryophanon latum. Applied and Environmental [124] Kawaguchi Y, Doi Y. Kinetics and mechanism of synthesis and
Microbiology 2007;73:586–93. degradation of poly(3-hydroxybutyrate) in Alcaligenes-Eutrophus.
[103] Mayer F, Hoppert M. Determination of the thickness of the Macromolecules 1992;25:2324–9.
boundary layer surrounding bacterial PHA inclusion bodies, and [125] Jia Y, Yuan W, Wodzinska J, Park C, Sinskey AJ, Stubbe J. Mecha-
implications for models describing the molecular architecture of nistic studies on class I polyhydroxybutyrate (PHB) synthase from
this layer. Journal of Basic Microbiology 1997;37:45–52. Ralstonia eutropha: class I and III synthases share a similar catalytic
[104] Dennis D, Sein V, Martinez E, Augustine B. PhaP is involved in mechanism. Biochemistry 2001;40:1011–9.
the formation of a network on the surface of polyhydroxyalka- [126] Wodzinska J, Snell KD, Rhomberg A, Sinskey AJ, Biemann K,
noate inclusions in Cupriavidus necator H16. Journal of Bacteriology Stubbe J. Polyhydroxybutyrate synthase: evidence for covalent
2008;190:555–63. catalysis. Journal of the American Chemical Society 1996;118:
[105] Beeby M, Cho M, Stubbe J, Jensen GJ. Growth and localization of 6319–20.
polyhydroxybutyrate granules in Ralstonia eutropha. Journal of Bac- [127] Gerngross TU, Snell KD, Peoples OP, Sinskey AJ. Overexpression and
teriology 2012;194:1092–9. purification of the soluble polyhydroxyalkanoate synthase from
[106] York GM, Stubbe J, Sinskey AJ. The Ralstonia eutropha PhaR protein Alcaligenes-eutrophus – evidence for a required posttranslational
couples synthesis of the PhaP phasin to the presence of polyhydrox- modification for catalytic activity. Biochemistry 1994;33:9311–20.
ybutyrate in cells and promotes polyhydroxybutyrate production. [128] Tian JM, Sinskey AJ, Stubbe J. Class III polyhydroxybutyrate
Journal of Bacteriology 2002;184:59–66. synthase: involvement in chain termination and reinitiation. Bio-
[107] Neumann L, Spinozzi F, Sinibaldi R, Rustichelli F, Pötter M, Stein- chemistry 2005;44:8369–77.
büchel A. Binding of the major phasin, PhaP1, from Ralstonia [129] Cai L, Tan D, Aibaidula G, Chen J-C, Tian W-D, Chen G-Q. Compar-
eutropha H16 to poly(3-hydroxybutyrate) granules. Journal of Bac- ative genomics study of polyhydroxyalkanoates (PHA) and ectoine
teriology 2008;190:2911–9. relevant genes from Halomonas sp TD01 revealed extensive hor-
[108] York GM, Stubbe J, Sinskey AJ. New insight into the role of the PhaP izontal gene transfer events and co-evolutionary relationships.
phasin of Ralstonia eutropha in promoting synthesis of polyhydrox- Microbial Cell Factories 2011;10:88.
ybutyrate. Journal of Bacteriology 2001;183:2394–7. [130] Li P, Chakraborty S, Stubbe J. Detection of covalent and non-
[109] Tian JM, He AM, Lawrence AG, Liu P, Watson N, Sinskey AJ, covalent intermediates in the polymerization reaction catalyzed
Stubbe J. Analysis of transient polyhydroxybutyrate production in by a C149S class III polyhydroxybutyrate synthase. Biochemistry
Wautersia eutropha H16 by quantitative western analysis and trans- 2009;48:9202–11.
mission electron microscopy. Journal of Bacteriology 2005;187: [131] Yamanaka K, Aoki T, Kudo T, Kimura Y. End-group analysis of bac-
3825–32. terially produced poly(3-hydroxybutyrate): discovery of succinate
[110] Hanley SZ, Pappin DJC, Rahman D, White AJ, Elborough KM, Slabas as the polymerization starter. Macromolecules 2009;42:4038–46.
AR. Re-evaluation of the primary structure of Ralstonia eutropha [132] Doi Y. Microbial synthesis, physical properties, and biodegrad-
phasin and implications for polyhydroxyalkanoic acid granule ability of polyhydroxyalkanoates. Macromolecular Symposium
binding. FEBS Letters 1999;447:99–105. 1995;98:585–99.
[111] Wieczorek R, Pries A, Steinbüchel A, Mayer F. Analysis of a [133] Kawaguchi Y, Doi Y. Kinetics and mechanism of synthesis and
24-Kilodalton protein associated with the polyhydroxyalkanoic degradation of poly(3-hydroxybutyrate) in Alcaligenes eutrophus.
acid granules in Alcaligenes-Eutrophus. Journal of Bacteriology Macromolecules 1992;25:2324–9.
1995;177:2425–35. [134] Kusaka S, Abe H, Lee SY, Doi Y. Molecular mass of poly[(R)-3-
[112] Horowitz DM, Sanders JKM. Amorphous, biomimetic granules hydroxybutyric acid] produced in a recombinant Escherichia coli.
of polyhydroxybutyrate – preparation, characterization, and bio- Applied Microbiology and Biotechnology 1997;47:140–3.
logical implications. Journal of the American Chemical Society [135] Madden LA, Anderson AJ, Shah DT, Asrar J. Chain termination
1994;116:2695–702. in polyhydroxyalkanoate synthesis: involvement of exogenous
436 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

hydroxy-compounds as chain transfer agents. International Journal [159] Harrison STL, Chase HA, Amor SR, Bonthrone KM, Saunders JKM.
of Biological Macromolecules 1999;25:43–53. Plasticization of poly(hydroxybutyrate) in vivo. International Jour-
[136] Kichise T, Fukui T, Yoshida Y, Doi Y. Biosynthesis of polyhydrox- nal of Biological Macromolecules 1992;14:50–6.
yalkanoates (PHA) by recombinant Ralstonia eutropha and effects [160] Barham PJ, Keller A, Otun EL, Holmes PA. Crystallization and mor-
of PHA synthase activity on in vivo PHA biosynthesis. International phology of a bacterial thermoplastic – poly-3-hydroxybutyrate.
Journal of Biological Macromolecules 1999;25:69–77. Journal of Materials Science 1984;19:2781–94.
[137] Shah DT, Berger PA, Tran M, Asrar J, Madden LA, Anderson [161] Tanaka T, Fujita M, Takeuchi A, Suzuki Y, Uesugi K, Doi Y. Structure
AJ. Synthesis of poly(3-hydroxybutyrate) by Ralstonia eutropha investigation of narrow banded spherulites in polyhydroxyalka-
in the presence of 13C-labeled ethylene glycol. Macromolecules noates by microbeam X-ray diffraction with synchrotron radiation.
2000;33:6624–6. Polymer 2005;46:5673–9.
[138] Shah DT, Tran M, Berger PA, Aggarwal P, Asrar J. Synthesis [162] Fujita M, Sawayanagi T, Tanaka T, Iwata T, Abe H, Doi Y. Synchrotron
and properties of hydroxy-terminated poly(hydroxyalkanoate)s. SAXS and WAXS studies on changes in structural and thermal prop-
Macromolecules 2000;33:2875–80. erties of poly[(R)3-hydroxybutyrate] single crystals during heating.
[139] Ashby RD, Solaiman DKY, Foglia TA. Synthesis of short-/medium- Macromolecular Rapid Communications 2005;26:678–83.
chain-length poly(hydroxyalkanoate) blends by mixed culture [163] Yokouchi M, Chatani Y, Tadokoro H, Teranish K, Tani H. Struc-
fermentation of glycerol. Biomacromolecules 2005;6:2106–12. tural studies of polyesters. 5. Molecular and crystal structures
[140] Ashby RD, Shi FY, Gross RA. Use of poly(ethylene glycol) of optically-active and racemic poly(␤-hydroxybutyrate). Polymer
to control the end group structure and molecular weight of 1973;14:267–72.
poly(3-hydroxybutyrate) formed by Alcaligenes latus DSM 1122. [164] Cornibert J, Marchessault RH. Physical properties of poly-␤-
Tetrahedron 1997;53:15209–23. hydroxybutyrate. IV. Conformational analysis and crystalline
[141] Shi FY, Ashby R, Gross RA. Use of poly(ethylene glycol)s to regu- structure. Journal of Molecular Biology 1972;71:735–56.
late poly(3-hydroxybutyrate) molecular weight during Alcaligenes [165] Scandola M, Ceccorulli G, Pizzoli M, Gazzano M. Study of the crystal
eutrophus cultivations. Macromolecules 1996;29:7753–8. phase and crystallization rate of bacterial poly(3-hydroxybutyrate-
[142] Sanguanchaipaiwong V, Gabelish CL, Hook J, Scholz C, Fos- co-3-hydroxyvalerate). Macromolecules 1992;25:1405–10.
ter LJR. Biosynthesis of natural-synthetic hybrid copolymers: [166] Marchessault RH, Morikawa H, Revol JF, Bluhm TL. Physical
polyhydroxyoctanoate-diethylene glycol. Biomacromolecules properties of a naturally-occurring polyester: poly(␤-
2004;5:643–9. hydroxyvalerate)/poly(␤-hydroxybutyrate). Macromolecules
[143] Saha SP, Patra A, Paul AK. Incorporation of polyethylene gly- 1984;17:1882–4.
col in polyhydroxyalkanoic acids accumulated by Azotobacter [167] Iwata T, Doi Y. Crystal structure and biodegradation of aliphatic
chroococcum MAL-201. Journal of Industrial Microbiology and polyester crystals. Macromolecular Chemistry and Physics
Biotechnology 2006;33:377–83. 1999;200:2429–42.
[144] Zanzig J, Scholz C. Effects of poly(ethylene glycol) on the production [168] Marchessault RH, Debzi EM, Revol JF, Steinbüchel A. Single-crystals
of poly(␤-hydroxybutyrate) by Azotobacter vinelandii UWD. Journal of bacterial and synthetic poly(3-hydroxyvalerate). Canadian Jour-
of Polymers and the Environment 2003;11:145–54. nal of Microbiology 1995;41(13):297–302.
[145] Suzuki T, Deguchi H, Yamane T, Shimizu S, Gekko K. Control of [169] Marchessault RH, Coulombe S, Morikawa H, Okamura K, Revol
molecular-weight of poly-␤-hydroxybutyric acid produced in fet- JF. Solid-state properties of poly-␤-hydroxybutyrate and of its
batch culture of Protomonas-Extorquens. Applied Microbiology oligomers. Canadian Journal of Chemistry 1981;59:38–44.
and Biotechnology 1988;27:487–91. [170] Sato H, Murakami R, Mori K, Ando Y, Takahashi I, Noda I. Spe-
[146] Doi Y, Segawa A, Kawaguchi Y, Kunioka M. Cyclic nature of cific crystal structure of poly(3-hydroxybutyrate) thin films studied
poly(3-hydroxyalkanoate) metabolism in Alcaligenes-Eutrophus. by infrared reflection-absorption spectroscopy. Vibrational Spec-
FEMS Microbiology Letters 1990;67:165–9. troscopy 2009;51:132–5.
[147] Taidi B, Mansfield DA, Anderson AJ. Turnover of poly(3- [171] Lundgren DG, Alper R, Schnaitman C, Marchessault RH. Char-
hydroxybutyrate) (P(3HB)) and its influence on the molecular-mass acterization of poly-␤-hydroxybutyrate extracted from different
of the polymer accumulated by Alcaligenes-Eutrophus during batch bacteria. Journal of Bacteriology 1965;89:245–51.
culture. FEMS Microbiology Letters 1995;129:201–5. [172] Mitomo H, Barham PJ, Keller A. Crystallization and morphology
[148] Penloglou G, Roussos A, Chatzidoukas C, Kiparissides C. A of poly(␤-hydroxybutyrate) and its copolymer. Polymer Journal
combined metabolic/polymerization kinetic model on the micro- 1987;19:1241–53.
bial production of poly(3-hydroxybutyrate). New Biotechnology [173] Welland EL, Stejny J, Halter A, Keller A. Selective degradation of
2010;27:358–67. chain folded single-crystals of poly(␤-hydroxybutyrate). Polymer
[149] Barnard GN, Sanders JKM. Observation of mobile poly(␤- Communications 1989;30:302–4.
hydroxybutyrate) in the storage granules of Methylobacterium Am1 [174] Birley C, Briddon J, Sykes KE, Barker PA, Organ SJ, Barham
by in vivo C-13-Nmr spectroscopy. FEBS Letters 1988;231:16–8. PJ. Morphology of single-crystals of poly(hydroxybutyrate) and
[150] Barnard GN, Sanders JKM. The poly-␤-hydroxybutyrate granule copolymers of hydroxybutyrate and hydroxyvalerate. Journal of
in vivo – a new insight based on NMR-spectroscopy of whole cells. Materials Science 1995;30:633–8.
Journal of Biological Chemistry 1989;264:3286–91. [175] Revol JF, Chanzy HD, Deslandes Y, Marchessault RH. High-
[151] Kawaguchi Y, Doi Y. Structure of native poly(3-hydroxybutyrate) resolution electron-microscopy of poly(␤-hydroxybutyrate). Poly-
granules characterized by X-ray-diffraction. FEMS Microbiology mer 1989;30:1973–6.
Letters 1990;70:151–6. [176] Nobes GAR, Briese BH, Jendrossek D, Marchessault RH. Microscopic
[152] Sanders JKM. NMR of nature’s plastics and spider’s webs – chem- visualization of the enzymatic degradation of poly(3HB-co-3HV)
istry, physics, or biology. Chemical Society Reviews 1993;22:1–7. and poly(3HV) single crystals by PHA depolymerases from Pseu-
[153] de Koning GJM, Lemstra PJ. The amorphous state of bac- domonas lemoignei. Journal of Environmental Polymer Degradation
terial poly[(R)-3-hydroxyalkanoate] in vivo. Polymer 1992;33: 1998;6:99–107.
3292–4. [177] Marchessault RH, Dou HY, Ramsay J. Microbial medium chainlength
[154] Bonthrone KM, Clauss J, Horowitz DM, Hunter BK, Sanders poly[(R)-3-hydroxyalkanoate] shows liquid crystal behaviour.
JKM. The biological and physical chemistry of polyhydroxyalka- International Journal of Biological Macromolecules 2011;48:
noates as seen by NMR-spectroscopy. FEMS Microbiology Review 271–5.
1992;103:269–77. [178] Lambeek G, Vorenkamp EJ, Schouten AJ. Structural study
[155] Barham PJ. Nucleation behavior of poly-3-hydroxy-butyrate. Jour- of Langmuir–Blodgett monolayers and multilayers of poly(␤-
nal of Materials Science 1984;19:3826–34. hydroxybutyrate). Macromolecules 1995;28:2023–32.
[156] Organ SJ, Barham PJ. Nucleation, growth and morphology of [179] Orts WJ, Marchessault RH, Bluhm TL, Hamer GK. Observation
poly(hydroxybutyrate) and its copolymers. Journal of Materials Sci- of strain-induced ␤-form in poly(␤-hydroxyalkanoates). Macro-
ence 1991;26:1368–74. molecules 1990;23:5368–70.
[157] Horowitz DM, Sanders JKM. Biomimetic, amorphous granules [180] Yamane H, Terao K, Hiki S, Kimura Y. Mechanical properties and
of polyhydroxyalkanoates – composition, mobility, and stabi- higher order structure of bacterial homo poly(3-hydroxybutyrate)
lization in vitro by proteins. Canadian Journal of Microbiology melt spun fibers. Polymer 2001;42:3241–8.
1995;41(13):115–23. [181] Fischer JJ, Aoyagi Y, Enoki M, Doi Y, Iwata T. Mechanical properties
[158] Lauzier C, Revol JF, Marchessault RH. Topotactic crystallization and enzymatic degradation of poly([R]-3-hydroxybutyrate-co-
of isolated poly(␤-hydroxybutyrate) granules from Alcaligenes [R]-3-hydroxyhexanoate) uniaxially cold-drawn films. Polymer
eutropha. FEMS Microbiology Review 1992;103:299–310. Degradation and Stability 2004;83:453–60.
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 437

[182] Tanaka T, Fujita M, Takeuchi A, Suzuki Y, Uesugi K, Ito K. Forma- [203] Mitomo H, Takahashi T, Ito H, Saito T. Biosynthesis and charac-
tion of highly ordered structure in poly[(R)-3-hydroxybutyrate- terization of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) pro-
co-(R)-3-hydroxyvalerate] high-strength fibers. Macromolecules duced by Burkholderia cepacia D1. International Journal of Biological
2006;39:2940–6. Macromolecules 1999;24:311–8.
[183] Iwata T, Sato S, Park JW. ␤ structure and unique crystalline ori- [204] Iwata T, Doi Y, Nakayama S, Sasatsuki H, Teramachi S. Structure
entation analysis of P(3HB) fibers and films. In: 11th international and enzymatic degradation of poly(3-hydroxybutyrate) copolymer
symposium on biological polyesters. 2008. p. 101. single crystals with an extracellular P(3HB) depolymerase from
[184] Kusaka S, Iwata T, Doi Y. Properties and biodegradability of ultra- Alcaligenes faecalis T1. International Journal of Biological Macro-
high-molecular-weight poly[(R)-3-hydroxybutyrate] produced by molecules 1999;25:169–76.
a recombinant Escherichia coli. International Journal of Biological [205] Kamiya N, Sakurai M, Inoue Y, Chujo R, Doi Y. Studies of cocrys-
Macromolecules 1999;25:87–94. tallization of poly(3-hydroxybutyrate-co-3-hydroxyvalerate)
[185] Furuhashi Y, Imamura Y, Jikihara Y, Yamane H. Higher order by solid-state high-resolution C-13 NMR-spectroscopy and
structures and mechanical properties of bacterial homo poly(3- differential scanning calorimetry. Macromolecules 1991;24:
hydroxybutyrate) fibers prepared by cold-drawing and annealing 2178–82.
processes. Polymer 2004;45:5703–12. [206] Yoshie N, Fujiwara M, Kasuya K, Abe H, Doi Y, Inoue Y. Effect of
[186] Nishiyama Y, Tanaka T, Yamazaki T, Iwata T. 2D NMR observation monomer composition and composition distribution on enzymatic
of strain-induced ␤-form in poly[(R)-3-hydroxybutyrate]. Macro- degradation of poly(3-hydroxybutyrate-co-3-hydroxyvalerate).
molecules 2006;39:4086–92. Macromolecular Chemistry and Physics 1999;200:977–82.
[187] Murakami R, Dybal J, Iwata T, Ozaki Y, Sato H. Formation and sta- [207] Barker PA, Mason F, Barham PJ. Density and crystallinity of
bility of ␤-structure in biodegradable ultra-high-molecular-weight poly(3-hydroxybutyrate 3-hydroxyvalerate) copolymers. Journal
poly(3-hydroxybutyrate) by infrared, Raman, and quantum chem- of Materials Science 1990;25:1952–6.
ical calculation studies. Polymer 2007;48:2672–80. [208] Yoshie N, Saito M, Inoue Y. Structural transition of lamella
[188] Ublekov F, Baldrian J, Nedkov E. Crystalline ␤-structure of P(3HB)V crystals in a isomorphous copolymer, poly(3-hydroxybutyrate-co-
grown epitaxially on silicate layers of MMT. Journal of Polymer 3-hydroxyvalerate). Macromolecules 2001;34:8953–60.
Science Part B: Polymer Physics 2009;47:751–5. [209] Barker PA, Barham PJ, Martinez-Salazar J. Effect of crystal-
[189] Antipov EM, Dubinsky VA, Rebrov AV, Nekrasov YP, Gordeev lization temperature on the cocrystallization of hydroxybu-
SA, Ungar G. Strain-induced mesophase and hard-elastic tyrate/hydroxyvalerate copolymers. Polymer 1997;38:913–9.
behaviour of biodegradable polyhydroxyalkanoates fibers. [210] Yoshie N, Sakurai M, Inoue Y, Chujo R. Cocrystallization of isother-
Polymer 2006;47:5678–90. mally crystallized poly(3-hydroxybutyrate-co-3-hydroxyvalerate).
[190] Bluhm TL, Hamer GK, Marchessault RH, Fyfe CA, Veregin Macromolecules 1992;25:2046–8.
RP. Isodimorphism in bacterial poly(␤-hydroxybutyrate-co-␤- [211] Barham PJ, Barker P, Organ SJ. Physical properties of
hydroxyvalerate). Macromolecules 1986;19:2871–6. poly(hydroxybutyrate) and copolymers of hydroxybutyrate and
[191] Inoue Y. Solid-state structure and properties of bacterial hydroxyvalerate. FEMS Microbiology Review 1992;103:289–98.
copolyesters. Journal of Molecular Structure 1998;441:119–27. [212] Yoshie N, Fujiwara M, Ohmori M, Inoue Y. Temperature depend-
[192] Kunioka M, Tamaki A, Doi Y. Crystalline and ence of cocrystallization and phase segregation in blends
thermal properties of bacterial copolyesters – of poly(3-hydroxybutyrate) and poly(3-hydroxybutyrate-co-3-
poly(3-hydroxybutyrate-co-3-hydroxyvalerate) and poly(3- hydroxyvalerate). Polymer 2001;42:8557–63.
hydroxybutyrate-co-4-hydroxybutyrate). Macromolecules [213] Guo LH, Sato H, Hashimoto T, Ozaki Y. Thermally induced exchanges
1989;22:694–7. of hydrogen bonding interactions and their effects on phase struc-
[193] Mitomo H, Morishita N, Doi Y. Composition range of crystal tures of poly(3-hydroxybutyrate) and poly(4-vinylphenol) blends.
phase-transition of isodimorphism in poly(3-hydroxybutyrate-co- Macromolecules 2011;44:2229–39.
3-hydroxyvalerate). Macromolecules 1993;26:5809–11. [214] Chen Y, Yang G, Chen Q. Solid-state NMR study on the structure and
[194] Mitomo H, Morishita N, Doi Y. Structural changes of poly(3- mobility of the noncrystalline region of poly(3-hydroxybutyrate)
hydroxybutyrate-co-3-hydroxyvalerate) fractionated with and poly(3-hydroxybutyrate-co-3-hydroxyvalerate). Polymer
acetone–water solution. Polymer 1995;36:2573–8. 2002;43:2095–9.
[195] Mitomo H, Doi Y. Lamellar thickening and cocrystallization of [215] Orts WJ, Vanderhart DL, Bluhm TL, Marchessault RH. Cocrys-
poly(hydroxyalkanoate)s on annealing. International Journal of tallization in random copolymers of poly(␤-hydroxybutyrate-
Biological Macromolecules 1999;25:201–5. co-␤-hydroxyvalerate) and its effect on crystalline morphology.
[196] Ishida K, Asakawa N, Inoue Y. Structure, properties and biodegrada- Canadian Journal of Chemistry 1995;73:2094–100.
tion of some bacterial copoly(hydroxyalkanoate)s. Macromolecular [216] Xie YP, Noda I, Akpaiu YA. Influence of cooling rate on the thermal
Symposium 2005;224:47–57. behavior and solid-state morphologies of polyhydroxyalkanoates.
[197] Feng LD, Yoshie Y, Asakawa N, Inoue Y. Comonomer-unit composi- Journal of Applied Polymer Science 2008;109:2259–68.
tions, physical properties and biodegradability of bacterial copoly- [217] Cao A, Asakawa N, Yoshie N, Inoue Y. High-resolution solid-state
hydroxyalkanoates. Macromolecular Bioscience 2004;4:186–98. C-13 n. m. r. study on phase structure of the compositionally
[198] Feng LD, Wang Y, Inagawa Y, Kasuya K, Saito T, Doi Y. Enzymatic fractionated bacterial copolyester poly(3-hydroxybutyric acid-co-
degradation behavior of comonomer compositionally fraction- 3-hydroxypropionic acid)s. Polymer 1999;40:3309–22.
ated bacterial poly(3-hydroxybutyrate-co-3-hydroxyvalerate)s by [218] Runt J, Kanchanasopa M. Crystallinity determination. Encyclopedia
poly(3-hydroxyalkanoate) depolymerases isolated from Ralstonia of polymer science and technology, vol. 9, 4th ed. New York: John
pickettii T1 and Acidovorax sp TP4. Polymer Degradation and Sta- Wiley & Sons Inc.; 2004. p. 446–64.
bility 2004;84:95–104. [219] Zhang LM, Tang HR, Hou GJ, Shen YD, Deng F. The domain struc-
[199] Tsz-Chun M, Chan PL, Lawford H, Chua H, Lo WH, Yu PH. Micro- ture and mobility of semi-crystalline poly(3-hydroxybutyrate) and
bial synthesis and characterization of physiochemical properties poly(3-hydroxybutyrate-co-3-hydroxyvalerate): a solid-state NMR
of polyhydroxyalkanoates (PHAs) produced by bacteria isolated study. Polymer 2007;48:2928–38.
from activated sludge obtained from the municipal wastewater [220] Kansiz M, Dominguez-Vidal A, McNaughton D, Lendl B. Fourier-
works in Hong Kong. Applied Biochemistry and Biotechnology transform infrared (FTIR) spectroscopy for monitoring and
2005;121:731–9. determining the degree of crystallisation of polyhydroxyalka-
[200] Thellen C, Coyne M, Froio D, Auerbach M, Wirsen C, Ratto JA. A noates (PHAs). Analytical and Bioanalytical Chemistry 2007;388:
processing, characterization and marine biodegradation study of 1207–13.
melt-extruded polyhydroxyalkanoate (PHA) films. Journal of Poly- [221] Bauer H, Owen AJ. Some structural and mechanical prop-
mers and the Environment 2008;16:1–11. erties of bacterially produced poly-␤-hydroxybutyrate-co-␤-
[201] Wang Y, Yamada S, Asakawa N, Yamane T, Yoshie N, Inoue Y. hydroxyvalerate. Colloid and Polymer Science 1988;266:241–7.
Comonomer compositional distribution and thermal and mor- [222] Steinbüchel A, Schmack G. Large-scale production of poly(3-
phological characteristics of bacterial poly(3-hydroxybutyrate- hydroxyvaleric acid) by fermentation of Chromobacterium vio-
co-3-hydroxyvalerate)s with high 3-hydroxyvalerate content. laceleum, processing, and characterization of the homopolyester.
Biomacromolecules 2001;2:1315–23. Journal of Environmental Polymer Degradation 1995;3:243–58.
[202] Yamada S, Wang Y, Asakawa N, Yoshie N, Inoue Y. Crys- [223] Pearce RP, Marchessault RH. Melting and crystallization in
talline structural change of bacterial poly(3-hydroxybutyrate- bacterial poly(␤-hydroxyvalerate), PHV, and blends with
co-3-hydroxyvalerate) with narrow compositional distribution. poly(␤-vhydroxybutyrate-co-hydroxyvalerate). Macromolecules
Macromolecules 2001;34:4659–61. 1994;27:3869–74.
438 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

[224] Bloembergen S, Holden DA, Marchessault RH. Non-biochemical [246] Wen X, Lu X, Peng Q, Zhu F, Zheng N. Crystallization behaviors
synthesis and characterization of poly(␤-hydroxybutyrate-co-␤- and morphology of biodegradable poly(3-hydroxybutyrate-co-4-
hydroxyvalerate). Polymer Preprints (American Chemical Society, hydroxybutyrate). Journal of Thermal Analysis and Calorimetry
Division of Polymer Chemistry) 1988;29(1):594–5. 2011, http://dx.doi.org/10.1007/s10973-011-1768-2.
[225] Gassner F, Owen AJ. Some properties of poly(3- [247] Gagnon KD, Lenz RW, Farris RJ, Fuller RC. Crystallization behavior
hydroxybutyrate)–poly(3-hydroxyvalerate) blends. Polymer and its influence on the mechanical properties of a thermoplastic
International 1996;39:215–9. elastomer produced by Pseudomonas-Oleovorans. Macromolecules
[226] Cheng ML, Sun YM. Relationship between free volume properties 1992;25:3723–8.
and structure of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) [248] de Koning GJM, van Bilsen HMM, Lemstra PJ, Hazenberg W, With-
membranes via various crystallization conditions. Polymer olt B, Preusting H, van der Galiën JG, Schirmer A, Jendrossek D.
2009;50:5298–307. A biodegradable rubber by cross-linking poly(hydroxyalkanoate)
[227] Chan CH, Kummerlowe C, Kammer HW. Crystallization and melting from Pseudomonas oleovorans. Polymer 1994;35:2090–7.
behavior of poly(3-hydroxybutyrate)-based blends. Macromolecu- [249] Marchessault RH, Monasterios CJ, Morin FG, Sundararajan PR. Chi-
lar Chemistry and Physics 2004;205:664–75. ral poly(␤-hydroxyalkanoates) – an adaptable helix influenced by
[228] Orts WJ, Romansky M, Guillet JE. Measurement of the crystallinity the alkane side-chain. International Journal of Biological Macro-
of poly(␤-hydroxybutyrate-co-␤-hydroxyvalerate) copolymers molecules 1990;12:158–65.
by inverse gas-chromatography. Macromolecules 1992;25: [250] Zhao K, Deng Y, Chen GQ. Effects of surface morphology on the bio-
949–53. compatibility of polyhydroxyalkanoates. Biochemical Engineering
[229] Kamiya N, Sakurai M, Inoue Y, Chujo R. Isomorphic behavior of ran- Journal 2003;16:115–23.
dom copolymers – thermodynamic analysis of co-crystallization of [251] Chanprateep S, Kikuya K, Shimizu H, Seki S, Tagawa S,
poly(3-hydroxybutyrate-co-3-hydroxyvalerate). Macromolecules Shioya S. Nonisothermal crystallization kinetics of biodegrad-
1991;24:3888–92. able random poly(3-hydroxybutyrate-co-3-hydroxyvalerate) and
[230] Chen HL, Hwang JC. Some comments on the degree of crys- block one. Journal of Chemical Engineering of Japan 2003;36:
tallinity defined by the enthalpy of melting. Polymer 1995;36: 639–46.
4355–7. [252] Chuah KP, Gan SN, Chee KK. Determination of Avrami exponent by
[231] Sawayanagi T, Tanaka T, Iwata T, Abe H, Doi Y, Ito K, Fujisawa differential scanning calorimetry for non-isothermal crystallization
T, Fujita M. Real-time synchrotron SAXS and WAXD studies on of polymers. Polymer 1999;40:253–9.
annealing behavior of poly[(R)-3-hydroxybutyrate] single crystals. [253] Shan GF, Gong X, Chen WP, Chen L, Zhu MF. Effect of
Macromolecules 2006;39:2201–8. multi-walled carbon nanotubes on crystallization behavior of
[232] Poley LH, Siqueira APL, da Silva MG, Sanchez R, Prioli R, Mansanares poly(3-hydroxybutyrate-co-3-hydroxyvalerate). Colloid and Poly-
AM, Vargas H. Photothermal methods and atomic force microscopy mer Science 2011;289:1005–14.
images applied to the study of poly(3-hydroxybutyrate) [254] Peng SW, An YX, Chen C, Fei B, Zhuang YG, Dong LS. Isothermal
and poly(3-hydroxybutyrate-co-3-hydroxyvalerate) dense crystallization of poly(3-hydroxybutyrate-co-3-hydroxyvalerate).
membranes. Journal of Applied Polymer Science 2005;97: European Polymer Journal 2003;39:1475–80.
1491–7. [255] Liu WJ, Yang HL, Wang Z, Dong LS, Liu JJ. Effect of nucleating
[233] Lotti N, Pizzoli M, Ceccorulli G, Scandola M. Binary blends of agents on the crystallization of poly(3-hydroxybutyrate-
microbial poly(3-hydroxybutyrate) with polymethacrylates. Poly- co-3-hydroxyvalerate). Journal of Applied Polymer Science
mer 1993;34:4935–40. 2002;86:2145–52.
[234] Gunaratne LMWK, Shanks RA. Multiple melting behaviour of [256] Saad GR, Mansour AA, Hamed AH. Dielectric investigation
poly(3-hydroxybutyrate-co-hydroxyvalerate) using step-scan DSC. of cold crystallization of poly(3-hydroxybutyrate). Polymer
European Polymer Journal 2005;41:2980–8. 1997;38:4091–6.
[235] Liu T, Petermann J. Multiple melting behavior in isothermally cold- [257] Xu J, Guo BH, Yang R, Wu Q, Chen GQ, Zhang ZM. In situ FTIR study
crystallized isotactic polystyrene. Polymer 2001;42:6453–61. on melting and crystallization of polyhydroxyalkanoates. Polymer
[236] Gan ZH, Abe H, Doi Y. Biodegradable poly(ethylene succinate) (PES). 2002;43:6893–9.
1. Crystal growth kinetics and morphology. Biomacromolecules [258] Ziaee Z, Supaphol P. Non-isothermal melt- and cold-
2000;1:704–12. crystallization kinetics of poly(3-hydroxybutyrate). Polymer
[237] Scandola M, Pizzoli M, Ceccorulli G, Cesaro A, Paoletti S, Testing 2006;25:807–18.
Navarini L. Viscoelastic and thermal properties of bacterial [259] Kammer HW, Kressler J, Kummerloewe C. Phase-behavior of poly-
poly(d-(−)-␤-hydroxybutyrate). International Journal of Biological mer blends – effects of thermodynamics and rheology. Advances in
Macromolecules 1988;10:373–7. Polymer Science 1993;106:31–85.
[238] Hoffman JD, Miller RL. Kinetic of crystallization from the melt and [260] Hill DJ, Markotsis M, Whittaker AK, Wong KW. NMR characteriza-
chain folding in polyethylene fractions revisited: theory and exper- tion of blends of poly(hydroxybutyrate-co-hydroxyvalerate) with
iment. Polymer 1997;38:3151–212. poly(vinyl acetate). Polymer International 2003;52:1780–9.
[239] Marand H, Xu JN, Srinivas S. Determination of the equilib- [261] Olabisi O, Robeson LM, Shaw MT. Polymer–polymer miscibility.
rium melting temperature of polymer crystals: linear and New York: Academic Press; 1979, 370 pp.
nonlinear Hoffman–Weeks extrapolations. Macromolecules [262] An YX, Li G, Mo ZS, Feng ZL, Dong LS. Miscibility, crystalliza-
1998;31:8219–29. tion kinetics, and morphology of poly(␤-hydroxybutyrate) and
[240] Organ SJ. Variation in melting point with molecular weight poly(methyl acrylate) blends. Journal of Polymer Science Part B:
for hydroxybutyrate hydroxyvalerate copolymers. Polymer Polymer Physics 2000;38:1860–7.
1993;34:2175–9. [263] Cox MK. Recycling biopol – composting and material recycling.
[241] Organ SJ, Barham PJ. On the equilibrium melting temperature of Journal of Macromolecular Science: Pure and Applied Chemistry
polyhydroxybutyrate. Polymer 1993;34:2169–74. 1995;A32:607–12.
[242] Kemnitzer JE, Gross RA, Mccarthy SP, Liggat J, Blundell DJ, Cox [264] Luzier WD. Materials derived from biomass biodegradable materi-
M. Crystallization behavior of predominantly syndiotactic poly(␤- als. Proceedings of the National Academy of Sciences of the United
hydroxybutyrate). Journal of Environmental Polymer Degradation States of America 1992;89:839–42.
1995;3:37–47. [265] Anderson AJ, Williams DR, Taidi B, Dawes EA, Ewing DF. Studies
[243] Bloembergen S, Holden DA, Hamer GK, Bluhm TL, Marches- on copolyester synthesis by Rhodococcus ruber and factors influ-
sault RH. Studies of composition and crystallinity of bacterial encing the molecular mass of polyhydroxybutyrate accumulated
poly(␤-hydroxybutyrate-co-␤-hydroxyvalerate). Macromolecules by Methylobacterium extorquens and Alcaligenes-eutrophus. FEMS
1986;19:2865–71. Microbiology Review 1992;103:93–101.
[244] Furukawa T, Sato H, Murakami R, Zhang JM, Duan YX, Noda I, [266] Van der Walle GAM, de Koning GJM, Weusthuis RA, Eggink G. Prop-
Ochiai S, Ozaki Y. Structure, dispersibility, and crystallinity of erties, modifications and applications of biotechnology. Advances
poly(hydroxybutyrate)/poly(l-lactic acid) blends studied by FT-IR in Biochemical Engineering/Biotechnology 2001;71:264–91.
microspectroscopy and differential scanning calorimetry. Macro- [267] van Wegen RJ, Lee SY, Middelberg APJ. Metabolic and kinetic
molecules 2005;38:6445–54. analysis of poly(3-hydroxybutyrate) production by recombinant
[245] Sato H, Murakami R, Padermshoke A, Hirose F, Senda K, Noda I, Escherichia coli. Biotechnology and Bioengineering 2001;74:70–80.
Ozaki Y. Infrared spectroscopy studies of CH••O hydrogen bondings [268] Reusch RN. Low-molecular-weight complexed poly(3-
and thermal behavior of biodegradable poly(hydroxyalkanoate). hydroxybutyrate) – a dynamic and versatile molecule in vivo.
Macromolecules 2004;37:7203–13. Canadian Journal of Microbiology 1995;41(13):50–4.
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 439

[269] Fiorese ML, Freitas F, Pais J, Ramos AM, de Araga GMF, Reis MAM. [290] Kelley AS, Srienc F. Production of two phase polyhydroxyalkanoic
Recovery of polyhydroxybutyrate (P(3HB)) from Cupriavidus neca- acid granules in Ralstonia eutropha. International Journal of Biolog-
tor biomass by solvent extraction with 1,2-propylene carbonate. ical Macromolecules 1999;25:61–7.
Engineering in Life Sciences 2009;9:454–61. [291] Yoon SC, Choi MH. Local sequence dependence of polyhydroxyalka-
[270] Akita S, Einaga Y, Miyaki Y, Fujita H. Solution properties of noic acid degradation in Hydrogenophaga pseudoflava. Journal of
poly(d-␤-hydroxybutyrate). 1. Biosynthesis and characterization. Biological Chemistry 1999;274:37800–8.
Macromolecules 1976;9:774–80. [292] Pederson EN, Srienc F. Mass spectrometry feedback control for
[271] Dias JML, Serafim LS, Lemos PC, Reis MAM, Oliveira R. Mathematical synthesis of polyhydroxyalkanoate granule microstructures in Ral-
modelling of a mixed culture cultivation process for the produc- stonia eutropha. Macromolecular Bioscience 2004;4:243–54.
tion of polyhydroxybutyrate. Biotechnology and Bioengineering [293] Madden LA, Asrar J, Anderson AJ. Synthesis and characterization
2005;92:209–22. of poly(3-hydroxybutyrate) and poly(3-hydroxybutyrate-co-3-
[272] Dias JML, Oehmen A, Serafim LS, Lemos PC, Reis MAM, Oliveira hydroxyvalerate) polymer mixtures produced in high-density
R. Metabolic modelling of polyhydroxyalkanoate copolymers pro- fed-batch cultures of Ralstonia eutropha (Alcaligenes eutrophus).
duction by mixed microbial cultures. BMC Systems Biology Macromolecules 1998;31:5660–7.
2008;2:59–80. [294] Mantzaris NV, Kelley AS, Srienc F, Daoutidis P. Optimal carbon
[273] Bradel R, Reichert KH. Modeling of molar-mass distribution source switching strategy for the production of PHA copolymers.
of poly(d-(−)-3-hydroxybutyrate) during bacterial synthesis. Die AIChE Journal 2001;47:727–43.
Makromolekulare Chemie 1993;194:1983–90. [295] Kelley AS, Srienc F. Controlling the polymer microstructure of
[274] Jurasek L, Marchessault RH. Polyhydroxyalkanoate (PHA) gran- biodegradable polyhydroxyalkanoates. In: Gross RA, Cheng HN,
ule formation in Ralstonia eutropha cells: a computer simulation. editors. Biocatalysis in polymer science. ACS Symposium Series, vol.
Applied Microbiology and Biotechnology 2004;64:611–7. 840. Washington, DC: American Chemical Society; 2002. p. 124–7.
[275] Mantzaris NV, Kelley AS, Daoutidis P, Srienc F. A population balance [296] Chanprateep S, Shimizu H, Shioya S. Characterization and enzy-
model describing the dynamics of molecular weight distributions matic degradation of microbial copolyester P(3HB-co-3HV)s
and the structure of PHA copolymer chains. Chemical Engineering produced by metabolic reaction model-based system. Polymer
Science 2002;57:4643–63. Degradation and Stability 2006;91:2941–50.
[276] Iadevaia S, Mantzaris NV. Genetic network driven control [297] Pederson EN, McChalicher CWJ, Srienc F. Bacterial synthesis of PHA
of P(3HB)V copolymer composition. Journal of Biotechnology block copolymers. Biomacromolecules 2006;7:1904–11.
2006;122:99–121. [298] Li SY, Dong CL, Wang SY, Ye HM, Chen GQ. Microbial produc-
[277] Yamanaka K, Aoki T, Kudo T, Kimura Y. Effect of ethylene glycol tion of polyhydroxyalkanoate block copolymer by recombinant
on the end group structure of poly(3-hydroxybutyrate). Polymer Pseudomonas putida. Applied Microbiology and Biotechnology
Degradation and Stability 2010;95:1284–91. 2011;90:659–69.
[278] Koizumi F, Abe H, Doi Y. Molecular-weight of poly(3- [299] Pereira SMF, Sanchez RJ, Rieumont J, Cabrera JG. Synthesis of
hydroxybutyrate) during biological polymerization in biodegradable polyhydroxyalcanoate copolymer from a renew-
Alcaligenes-eutrophus. Journal of Macromolecular Science: able source by alternate feeding. Polymer Engineering and Science
Pure and Applied Chemistry 1995;A32:759–74. 2008;48:2051–9.
[279] Shimizu H, Tamura S, Shioya S, Suga K. Kinetic study of poly-d(−)-3- [300] Inoue Y, Yoshie N. Structure and physical properties of bac-
hydroxybutyric acid (P(3HB)) production and its molecular-weight terially synthesized polyesters. Progress in Polymer Science
distribution control in a fed-batch culture of Alcaligenes eutrophus. 1992;17:571–610.
Journal of Fermentation and Bioengineering 1993;76:465–9. [301] Yoshie N, Saito M, Inoue Y. Effect of chemical composi-
[280] Hiraishi T, Kikkawa Y, Fujita M, Normi YM, Kanesato M, Tsuge tional distribution on solid-state structures and properties
T, Sudesh K, Maeda M, Doi Y. Atomic force microscopic observa- of poly(3-hydroxybutyrate-co-3-hydroxyvalerate). Polymer
tion of in vitro polymerized poly[(R)-3-hydroxybutyrate]: insight 2004;45:1903–11.
into possible mechanism of granule formation. Biomacromolecules [302] Yoshie N, Inoue Y. Chemical composition distribution of bacterial
2005;6:2671–7. copolyesters. International Journal of Biological Macromolecules
[281] Kurja J, Zirkzee HF, de Koning GM, Maxwell IA. A new 1999;25:193–200.
kinetic-model for the accumulation of poly-3-hydroxybutyrate in [303] Žagar E, Kržan A. Sequence distribution dependence on molar
Alcaligenes-eutrophus, 1. Granule growth. Macromolecular Theory mass in microbial poly(3-hydroxybutyrate-co-3-hydroxyvalerate)
and Simulations 1995;4:839–55. copolyesters. Acta Chimica Slovenica 2009;56:386–91.
[282] Sim SJ, Snell KD, Hogan SA, Stubbe J, Rha CK, Sinskey AJ. PHA syn- [304] Tanadchangsaeng N, Tsuge T, Abe H. Comonomer compo-
thase activity controls the molecular weight and polydispersity of sitional distribution, physical properties, and enzymatic
polyhydroxybutyrate in vivo. Nature Biotechnology 1997;15:63–7. degradability of bacterial poly(3-hydroxybutyrate-co-3-
[283] Huisman GW, Wonink E, de Koning G, Preusting H, Witholt B. hydroxy-4-methylvalerate) copolyesters. Biomacromolecules
Synthesis of poly(3-hydroxyalkanoates) by mutant and recombi- 2010;11:1615–22.
nant Pseudomonas strains. Applied Microbiology and Biotechnology [305] Kamiya N, Yamamoto Y, Inoue Y, Chujo R, Doi Y. Micro-
1992;38:1–5. structure of bacterially synthesized poly(3-hydroxybutyrate-co-3-
[284] Harris JM, editor. Poly(ethylene glycol) chemistry – biotechnical hydroxyvalerate). Macromolecules 1989;22:1676–82.
and biomedical applications. New York: Plenum Press; 1992, 408 [306] Doi Y, Kunioka M, Nakamura Y, Soga K. Nuclear-magnetic-
pp. resonance studies on poly(␤-hydroxybutyrate) and a copolyester
[285] Yamazaki M, Ito T. Deformation and instability in membrane- of ␤-hydroxybutyrate and ␤-hydroxyvalerate isolated from Alca-
structure of phospholipid-vesicles caused by osmophobic asso- ligenes eutrophus H16. Macromolecules 1986;19:2860–4.
ciation – mechanical-stress model for the mechanism of [307] Inoue Y, Kamiya N, Yamamoto Y, Chujo R, Doi Y. Micro-
poly(ethylene glycol)-induced membrane-fusion. Biochemistry structures of commercially available poly(3-hydroxybutyrate-co-
1990;29:1309–14. 3-hydroxyvalerate)s. Macromolecules 1989;22:3800–2.
[286] Helm J, Wendlandt KD, Jechorek M, Stottmeister U. Potassium [308] Adamus G, Sikorska W, Kowalczuk M, Montaudo M, Scandola
deficiency results in accumulation of ultra-high molecular weight M. Sequence distribution and fragmentation studies of bacte-
poly-␤-hydroxybutyrate in a methane-utilizing mixed culture. rial copolyester macromolecules: characterization of P(3HB)V
Journal of Applied Microbiology 2008;105:1054–61. macroinitiator by electrospray ion-trap multistage mass spectrom-
[287] Taidi B, Anderson AJ, Dawes EA, Byrom D. Effect of carbon etry. Macromolecules 2000;33:5797–802.
source and concentration on the molecular-mass of poly(3- [309] Ballistreri A, Montaudo G, Garozzo D, Chujo R, Doi Y. Microstruc-
hydroxybutyrate) produced by Methylobacterium-extorquens and ture of bacterial poly(␤-hydroxybutyrate-co-␤-hydroxyvalerate)
Alcaligenes-eutrophus. Applied Microbiology and Biotechnology by fast-atom-bombardment mass-spectrometry: analysis of the
1994;40:786–90. partial pyrolysis products. Macromolecules 1991;24:1231–6.
[288] McChalicher CWJ, Srienc F. Investigating the structure–property [310] Žagar E, Kržan A, Adamus G, Kowalczuk M. Sequence distribu-
relationship of bacterial PHA block copolymers. Journal of Biotech- tion in microbial poly(3-hydroxybutyrate-co-3-hydroxyvalerate)
nology 2007;132:296–302. co-polyesters determined by NMR and MS. Biomacromolecules
[289] Curley JM, Lenz RW, Fuller RC. Sequential production of two 2006;7:2210–6.
different polyesters in the inclusion bodies of Pseudomonas oleovo- [311] Lau NS, Tsuge T, Sudesh K. Formation of new polyhydroxyalkanoate
rans. International Journal of Biological Macromolecules 1996;19: containing 3-hydroxy-4-methylvalerate monomer in Burkholderia
29–34. sp. Applied Microbiology and Biotechnology 2011;89:1599–609.
440 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

[312] Don TM, Chen CW, Chan TH. Preparation and characterization [332] Yu F, Zhu B, Dong T, Inoue Y. Effect of comonomer-unit compo-
of poly(hydroxyalkanoate) from the fermentation of Haloferax sitional distribution on thermal and crystallization behavior of
mediterranei. Journal of Biomaterials Science, Polymer Edition bacterial poly[(3-hydroxybutyrate)-co-(3-mercaptopropionate)].
2006;17:1425–38. Macromolecular Bioscience 2009;9:702–12.
[313] Yoshie N, Menju H, Sato H, Inoue Y. Complex composition distri- [333] Kumagai Y, Kanesawa Y, Doi Y. Enzymatic degradation of micro-
bution of poly(3-hydroxybutyrate-co-3-hydroxyvalerate). Macro- bial poly(3-hydroxybutyrate) films. Die Makromolekulare Chemie
molecules 1995;28:6516–21. 1992;193:53–7.
[314] Inoue Y, Sano F, Nakamura K, Yoshie N, Saito Y, Satoh H. [334] Organ SJ. Phase separation in blends of poly(hydroxybutyrate) with
Microstructure of copoly(3-hydroxyalkanoates) produced in the poly(hydroxybutyrate-co-hydroxyvalerate) – variation with blend
anaerobic–aerobic activated sludge process. Polymer International components. Polymer 1994;35:86–92.
1996;39:183–9. [335] Saito M, Inoue Y, Yoshie N. Cocrystallization and phase
[315] Cao A, Ichikawa M, Kasuya KI, Yoshie N, Asakawa N, Inoue Y, Doi Y, segregation of blends of poly(3-hydroxybutyrate) and
Abe H. Composition fractionation and thermal characterization of poly(3-hydroxybutyrate-co-3-hydroxyvalerate). Polymer
poly(3-hydroxybutyrate-co-3-hydroxypropionate). Polymer Jour- 2001;42:5573–80.
nal 1996;28:1096–102. [336] Yoshie N, Asaka A, Inoue Y. Cocrystallization and phase segregation
[316] Cao A, Ichikawa M, Ikejima T, Yoshie N, Inoue Y. Thermal in crystalline/crystalline polymer blends of bacterial copolyesters.
and morphological study of fractionated poly(3-hydroxybutyric Macromolecules 2004;37:3770–9.
acid-co-3-hydroxypropionic acid). Macromolecular Chemistry and [337] Yoshie N, Inoue Y. Cocrystallization and phase segregation in
Physics 1997;198:3539–57. blends of two bacterial polyesters. Macromolecular Symposium
[317] Cao A, Kasuya K, Abe H, Doi Y, Inoue Y. Studies on comonomer 2005;224:59–70.
compositional distribution of the bacterial poly(3-hydroxybutyric [338] Saad GR. Blends of bacterial poly[(R)-3-hydroxybutyrate] with
acid-co-3-hydroxypropionic acid)s and crystal and thermal char- oligo[(R,S)-3-hydroxybutyrate]-diol. Polymer International
acteristics of their fractionated component copolyesters. Polymer 2002;51:338–48.
1998;39:4801–16. [339] Pearce R, Jesudason J, Orts W, Marchessault RH. Blends of bacte-
[318] Cao A, Arai Y, Yoshie N, Kasuya KI, Doi Y, Inoue Y. Solid rial and synthetic poly(␤-hydroxybutyrate) – effect of tacticity on
structure and biodegradation of the compositionally fractionated melting behavior. Polymer 1992;33:4647–9.
poly(3-hydroxybutyric acid-co-3-hydroxypropionic acid)s. Poly- [340] Scandola M, Focarete ML, Adamus G, Sikorska W, Baranowska I,
mer 1999;40:6821–30. Swierczek S, Gnatowski M, Kowalczuk M, Jedliński Z. Polymer
[319] Arai Y, Cao A, Yoshie N, Inoue Y. Studies on comonomer com- blends of natural poly(3-hydroxybutyrate-co-3-hydroxyvalerate)
positional distribution and its effect on some physical properties and a synthetic atactic poly(3-hydroxybutyrate). Characteri-
of bacterial poly(3-hydroxybutyric acid-co-3-hydroxypropionic zation and biodegradation studies. Macromolecules 1997;30:
acid). Polymer International 1999;48:1219–28. 2568–74.
[320] Wang Y, Ichikawa M, Cao A, Yoshie N, Inoue Y. Comonomer compo- [341] Hirota Y, Yoshie N, Ishii N, Kasuya KI, Inoue Y. Correlation between
sition distribution of P(3HB-co-3HP)s produced by Alcaligenes latus solid-state structures and enzymatic degradability of cocrystallized
at several pH conditions. Macromolecular Chemistry and Physics blends. Macromolecular Bioscience 2005;5:1094–100.
1999;200:1047–53. [342] Feng LD, Watanabe T, He Y, Wang Y, Kichise T, Fukuchi T. Phase
[321] Yu F, Nakamura N, Inoue Y. Comonomer-unit compositional dis- behavior and thermal properties for binary blends of bacterial
tribution and its effect on thermal and crystallization behavior of poly(3-hydroxybutyrate-co-3-hydroxyhexanoate)s with narrow-
bacterial poly(3-hydroxybutyrate-co-3-hydroxyhexanoate). Poly- comonomer-unit compositional distribution. Macromolecular Bio-
mer 2010;51:4408–18. science 2003;3:310–9.
[322] Watanabe T, He Y, Fukuchi T, Inoue Y. Comonomer compositional [343] de Koning D. Physical properties of bacterial poly((R)-
distribution and thermal characteristics of bacterially synthesized 3-hydroxyalkanoates). Canadian Journal of Microbiology
poly(3-hydroxybutyrate-co-3-hydroxyhexanoate)s. Macromolec- 1995;41(13):303–9.
ular Bioscience 2001;1:75–83. [344] Noohom W, Jack KS, Martin D, Trau M. Understanding the roles of
[323] Feng LD, Watanabe T, Wang Y, Kichise T, Fukuchi T, Chen G-Q, Doi nanoparticle dispersion and polymer crystallinity in controlling the
Y, Inoue Y. Studies on comonomer compositional distribution of mechanical properties of HA/P(3HB)V nanocomposites. Biomedical
bacterial poly(3-hydroxybutyrate-co-3-hydroxyhexanoate)s and Materials 2009;4, 015003/1–13.
thermal characteristics of their factions. Biomacromolecules [345] Bloembergen S, Holden DA, Bluhm TL, Hamer GK, Marchessault
2002;3:1071–7. RH. Isodimorphism in synthetic poly(␤-hydroxybutyrate-co-␤-
[324] Asrar J, Valentin HE, Berger PA, Tran M, Padgette SR, Garbow hydroxyvalerate) – stereoregular copolyesters from racemic
JR. Biosynthesis and properties of poly(3-hydroxybutyrate- ␤-lactones. Macromolecules 1989;22:1663–9.
co-3-hydroxyhexanoate) polymers. Biomacromolecules [346] Zhao W, Chen GQ. Production and characterization of
2002;3:1006–12. terpolyester poly(3-hydroxybutyrate-co-3-hydroxyvalerate-co-3-
[325] Doi Y, Segawa A, Kunioka M. Biosynthesis and characterization hydroxyhexanoate) by recombinant Aeromonas hydrophila 4AK4
of poly(3-hydroxybutyrate-co-4-hydroxybutyrate) in Alcaligenes harboring genes phaAB. Process Biochemistry 2007;42:1342–7.
eutrophus. International Journal of Biological Macromolecules [347] Shimamura E, Scandola M, Doi Y. Microbial synthesis and char-
1990;12:106–11. acterization of poly(3-hydroxybutyrate-co-3-hydroxypropionate).
[326] Shi FY, Ashby RD, Gross RA. Fractionation and characteriza- Macromolecules 1994;160:4429–35.
tion of microbial polyesters containing 3-hydroxybutyrate and [348] Lemos PC, Viana C, Salgueiro EN, Ramos AM, Crespo JPSG, Reis MAM.
4-hydroxybutyrate repeat units. Macromolecules 1997;30:2521–3. Effect of carbon source on the formation of polyhydroxyalkanoates
[327] Ishida K, Wang Y, Inoue Y. Comonomer unit composition (PHA) by a phosphate-accumulating mixed culture. Enzyme and
and thermal properties of poly(3-hydroxybutyrate-co-4- Microbial Technology 1998;22:662–71.
hydroxybutyrate)s biosynthesized by Ralstonia eutropha. [349] Patel M, Gapes DJ, Newman RH, Dare PH. Physico-chemical
Biomacromolecules 2001;2:1285–93. properties of polyhydroxyalkanoate produced by mixed-culture
[328] Valentin HE, Berger PA, Gruys KJ, de Andrade Rodrigues MF, nitrogen-fixing bacteria. Applied Microbiology and Biotechnology
Steinbüchel A, Tran M, Asrar J. Biosynthesis and character- 2009;82:545–55.
ization of poly(3-hydroxy-4-pentenoic acid). Macromolecules [350] Carrasco F, Dionisi D, Martinelli A, Majone M. Thermal stabil-
1999;32:7389–95. ity of polyhydroxyalkanoates. Journal of Applied Polymer Science
[329] Chen SS, Liu Q, Wang HH, Zhu B, Yu F, Chen G-Q, Inoue Y. Polymor- 2006;100:2111–21.
phic crystallization of fractionated microbial medium-chain-length [351] Dobroth ZT, Hu SJ, McDonald AG, Coats ER. Polyhydroxybutyrate
polyhydroxyalkanoates. Polymer 2009;50:4378–88. synthesis on biodiesel wastewater using mixed microbial consortia.
[330] Kato M, Bao HJ, Kang CK, Fukui T, Doi Y. Production of a novel Bioresource Technology 2011;102:3352–9.
copolyester of 3-hydroxybutyric acid and medium chain length [352] Iwata T, Tanaka T. Manufacturing of PHA as fibers. In: Chen G-Q,
3-hydroxyalkanaic acids by Pseudomonas sp 61-3 from sugars. editor. Plastics from bacteria: natural functions and applications.
Applied Microbiology and Biotechnology 1996;45:363–70. Berlin; Heidelberg: Springer-Verlag; 2010. p. 257–82.
[331] Ashby RD, Solaiman DKY, Foglia TA. The synthesis of short- [353] Miller ND, Williams DF. On the biodegradation of
and medium-chain-length poly(hydroxyalkanoate) mixtures from poly-␤-hydroxybutyrate (P(3HB)) homopolymer and poly-␤-
glucose- or alkanoic acid-grown Pseudomonas oleovorans. Journal hydroxybutyrate-hydroxyvalerate copolymers. Biomaterials
of Industrial Microbiology and Biotechnology 2002;28:147–53. 1987;8:129–37.
B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442 441

[354] Schmack G, Jehnichen D, Vogel R, Tändler B. Biodegradable fibers [377] Punrattanasin W. The utilization of activated sludge polyhydrox-
of poly(3-hydroxybutyrate) produced by high-speed melt spin- yalkanoates for the production of biodegradable plastics. PhD
ning and spin drawing. Journal of Polymer Science Part B: Polymer Dissertation. Environmental Science and Engineering, Blacksburg,
Physics 2000;38:2841–50. Virginia, USA: Virginia Polytechnic Institute and State University;
[355] Iwata T. Strong fibers and films of microbial polyesters. Macro- 2001. 120 pp.
molecular Bioscience 2005;5:689–701. [378] Serafim LS, Lemos PC, Torres C, Reis MAM, Ramos AM. The influence
[356] Iwata T, Aoyagi Y, Fujita M, Yamane H, Doi Y, Suzuki Y. Processing of process parameters on the characteristics of polyhydroxyalka-
of a strong biodegradable poly[(R)-3-hydroxybutyrate] fiber and noates produced by mixed cultures. Macromolecular Bioscience
a new fiber structure revealed by micro-beam X-ray diffraction 2008;8:355–66.
with synchrotron radiation. Macromolecular Rapid Communica- [379] Morgan-Sagastume F, Karlsson A, Johansson P, Pratt S, Boon N,
tions 2004;25:1100–4. Lant P, Werker A. Production of polyhydroxyalkanoates in open,
[357] Tanaka T, Yabe T, Teramachi S, Iwata T. Mechanical properties mixed cultures from a waste sludge stream containing high lev-
and enzymatic degradation of poly[(R)-3-hydroxybutyrate] fibers els of soluble organics, nitrogen and phosphorus. Water Research
stretched after isothermal crystallization near Tg. Polymer Degra- 2010;44:5196–211.
dation and Stability 2007;92:1016–24. [380] Chanprateep S, Buasri K, Muangwong A, Utiswannakul P.
[358] Barham PJ, Keller A. The relationship between microstructure and Biosynthesis and biocompatibility of biodegradable poly(3-
mode of fracture in polyhydroxybutyrate. Journal of Polymer Sci- hydroxybutyrate-co-4-hydroxybutyrate). Polymer Degradation
ence Part B: Polymer Physics 1986;24:69–77. and Stability 2010;95:2003–12.
[359] Martinez-Salazar J, Sanchez-Cuesta M, Barham P, Keller A. Thermal [381] de Koning GJM, Lemstra PJ. Crystallization phenomena in bacterial
expansion and spherulite cracking in 3-hydroxybutyrate/3- poly[(R)-3-hydroxybutyrate]. 2. Embrittlement and rejuvenation.
hydroxyvalerate copolymers. Journal of Materials Science Letters Polymer 1993;34:4089–94.
1989;8:490–549. [382] Doi Y, Kitamura S, Abe H. Microbial synthesis and char-
[360] El-Hadi A, Schnabel R, Straube E, Muller G, Henning S. Cor- acterization of poly(3-hydroxybutyrate-co-3-hydroxyhexanoate).
relation between degree of crystallinity, morphology, glass Macromolecules 1995;28:4822–8.
temperature, mechanical properties and biodegradation of poly(3- [383] Gogotov IN, Gerasin VA, Knyazev YV, Antipov EM, Barazov SK. Com-
hydroxyalkanoate) PHAs and their blends. Polymer Testing posite biodegradable materials based on polyhydroxyalkanoate.
2002;21:665–74. Applied Biochemistry and Microbiology 2010;46:607–13.
[361] Kai WH, He Y, Inoue Y. Fast crystallization of [384] Gregorova A, Wimmer R, Hrabalova M, Koller M, Ters T, Mundigler
poly(3-hydroxybutyrate) and poly(3-hydroxybutyrate-co-3- N. Effect of surface modification of beech wood flour on mechani-
hydroxy-valerate) with talc and boron nitride as nucleating cal and thermal properties of poly(3-hydroxybutyrate)/wood flour
agents. Polymer International 2005;54:780–9. composites. Holzforschung 2009;63:565–70.
[362] Kai WH, He Y, Asakawa N, Inoue Y. Effect of lignin particles as [385] Howells ER. Opportunities in biotechnology for the chemical
a nucleating agent on crystallization of poly(3-hydroxybutyrate). industry. Journal of the Society of Chemical Industry, London
Journal of Applied Polymer Science 2004;94:2466–74. 1982;8:508–11.
[363] He Y, Inoue Y. ␣-Cyclodextrin-enhanced crystallization of poly(3- [386] King PP. Biotechnology – an industrial view. Journal of Chemical
hydroxybutyrate). Biomacromolecules 2003;4:1865–7. Technology and Biotechnology 1982;32:2–8.
[364] Jacquel N, Tajima K, Nakamura N, Miyagawa T, Pan PJ, Inoue [387] Mitomo H, Barham PJ, Keller A. Temperature-dependence
Y. Effect of orotic acid as a nucleating agent on the crystalliza- of mechanical-properties of poly(␤-hydroxybutyrate-␤-
tion of bacterial poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) hydroxyvalerate). Polymer Communications 1988;29:112–5.
copolymers. Journal of Applied Polymer Science 2009;114: [388] Savenkova L, Gercberga Z, Bibers I, Kalnin M. Effect of 3-hydroxy
1287–94. valerate content on some physical and mechanical properties of
[365] Volova T. Polyhydroxyalkanoates – plastic materials of the 21st cen- polyhydroxyalkanoates produced by Azotobacter chroococcum. Pro-
tury: production, properties, applications. Hauppauge, NY: Nova cess Biochemistry 2000;36:445–50.
Publishers; 2004, 282 pp. [389] Cai ZJ, Yang G. Optical nanocomposites prepared by incorporating
[366] Choi JS, Park WH. Effect of biodegradable plasticizers on thermal bacterial cellulose nanofibrils into poly(3-hydroxybutyrate). Mate-
and mechanical properties of poly(3-hydroxybutyrate). Polymer rials Letters 2011;65:182–4.
Testing 2004;23:455–60. [390] Wang SA, Ma PM, Wang RY, Wang SF, Zhang Y,
[367] Bordes P, Pollet E, Averous L. Nano-biocomposites: biodegrad- Zhang YX. Mechanical, thermal and degradation
able polyester/nanoclay systems. Progress in Polymer Science properties of poly(d,l-lactide)/poly(hydroxybutyrate-co-
2009;34:125–55. hydroxyvalerate)/poly(ethylene glycol) blend. Polymer
[368] Galego N, Rozsa C, Sanchez R, Fung J, Vazquez A, Tomas JS. Charac- Degradation and Stability 2008;93:1364–9.
terization and application of poly(␤-hydroxyalkanoates) family as [391] Wang W, Zhang Y, Zhu MF, Chen YM. Effect of graft modification
composite biomaterials. Polymer Testing 2000;19:485–92. with poly(N-vinylpyrrolidone) on thermal and mechanical proper-
[369] Sherman LM. Enhancing biopolymers: additives are needed ties of poly(3-hydroxybutyrate-co-3-hydroxyvalerate). Journal of
for toughness, heat resistance and processability; 2012, Applied Polymer Science 2008;109:1699–707.
http://www.allbusiness.com/chemicals/plastics-rubber-industry- [392] Modi SJ. Assessing the feasibility of poly(3-hydroxybutyrate-co-3-
resins/11421097-1.html#ixzz1lKy5XtiG [accessed 31 January hydroxyvalerate) (P(3HB)V) and poly(lactic acid) for potential food
2012]. packaging applications. MS Dissertation. Ohio State University.
[370] Barham PJ, Organ SJ. Mechanical properties of http://rave.ohiolink.edu/etdc/view?acc num=osu1268921056;
polyhydroxybutyrate–hydroxybutyrate–hydroxyvalerate 2010. 123 pp.
copolymer blends. Journal of Materials Science 1994;29:1676–9. [393] Avella M, La Rota G, Martuscelli E, Raimo M, Sadocco P, Ele-
[371] Satoh H, Yoshie N, Inoue Y. Hydrolytic degradation of blends gir G, Riva R. Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) and
of poly(3-hydroxybutyrate) with poly(3-hydroxybutyrate-co-3- wheat straw fibre composites: thermal, mechanical proper-
hydroxyvalerate). Polymer 1994;35:286–90. ties and biodegradation behaviour. Journal of Materials Science
[372] Yang F, Li Z, Qiu Z. Miscibility and crystallization 2000;35:829–36.
behavior of biodegradable poly(3-hydroxybutyrate-co-3- [394] Fei B, Chen C, Chen S, Peng S, Zhuang Y, An Y, Dong L.
hydroxyvalerate)/phenolic blends. Journal of Applied Polymer Crosslinking of poly[(3-hydroxybutyrate)-co-(3-hydroxyvalerate)]
Science 2012;123:2781–6. using dicumyl peroxide as initiator. Polymer International 2004;53:
[373] Ha CS, Cho WJ. Miscibility, properties, and biodegradability of 937–43.
microbial polyester containing blends. Progress in Polymer Science [395] Zhang HF, Ma L, Wang ZH, Chen GQ. Biosynthesis and charac-
2002;27:759–809. terization of 3-hydroxyalkanoate terpolyesters with adjustable
[374] Yu L, Dean K, Li L. Polymer blends and composites from renewable properties by Aeromonas hydrophila. Biotechnology and Bioengi-
resources. Progress in Polymer Science 2006;31:576–602. neering 2009;104:582–9.
[375] Zhang JW, McCarthy S, Whitehouse R. Reverse temperature injec- [396] Chen GX, Hao GJ, Guo TY, Song MD, Zhang BH. Crystallization
tion molding of BiopolTM and effect on its properties. Journal of kinetics of poly(3-hydroxybutyrate-co-3-hydroxyvalerate)/clay
Applied Polymer Science 2004;94:483–91. nanocomposites. Journal of Applied Polymer Science
[376] Chua H, Hu PHF, Ma CK. Accumulation of biopolymers in acti- 2004;93:655–61.
vated sludge biomass. Applied Biochemistry and Biotechnology [397] Gursel I, Balcik C, Arica Y, Akkus O, Akkas N, Hasirci V. Synthe-
1999;77(9):389–99. sis and mechanical properties of interpenetrating networks of
442 B. Laycock et al. / Progress in Polymer Science 39 (2014) 397–442

polyhydroxybutyrate-co-hydroxyvalerate and polyhydroxyethyl polyester poly-(3-hydroxybutyrate-co-3-hydroxyvalerate). Poly-


methacrylate. Biomaterials 1998;19:1137–43. mer 1992;33:817–22.
[398] Jiang L, Huang JJ, Qian J, Chen F, Zhang JW, Wolcott MP, [404] Tanadchangsaeng N, Kitagawa A, Yamamoto T, Abe H, Tsuge
Zhu Y. Study of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) T. Identification, biosynthesis, and characterization of poly-
(P(3HB)V)/bamboo pulp fiber composites: effects of nucleation hydroxyalkanoate copolymer consisting of 3-hydroxybutyrate
agent and compatibilizer. Journal of Polymers and the Environment and 3-hydroxy-4-methylvalerate. Biomacromolecules 2009;10:
2008;16:83–93. 2866–74.
[399] Avella M, Bogoeva-Gaceva G, Buzarovska A, Errico ME, Gentile [405] Hassan MK, Abou-Hussein R, Zhang XJ, Mark JE, Noda I. Biodegrad-
G, Grozdanov A. Poly(3-hydroxybutyrate-co-3-hydroxyvalerate)- able copolymers of 3-hydroxybutyrate-co-3-hydroxyhexanoate
based biocomposites reinforced with kenaf fibers. Journal of (NodaxTM ), including recent improvements in their mechan-
Applied Polymer Science 2007;104:3192–200. ical properties. Molecular Crystals and Liquid Crystals
[400] Bhardwaj R, Mohanty AK, Drzal LT, Pourboghrat F, Misra M. Renew- 2006;447:341–62.
able resource-based green composites from recycled cellulose [406] Shen L, Haufe J, Patel MK. Product overview and market projection
fiber and poly(3-hydroxybutyrate-co-3-hydroxyvalerate) bioplas- of emerging biobased plastics (PROBIP 2009). Commissioned by
tic. Biomacromolecules 2006;7:2044–51. European Polysaccharide Network of Excellence (EPNOE) and Euro-
[401] Dagnon KL, Chen HH, Lnnocentini-Mei LH, D’Souza NA. pean Bioplastics. Report No.: NWS-E-2009-32. Science, Technology
Poly[(3-hydroxybutyrate)-co-(3-hydroxyvalerate)]/layered and Society Group, Copernicus Institute for Sustainable Develop-
double hydroxide nanocomposites. Polymer International ment and Innovation Utrecht The Netherlands: Utrecht University;
2009;58:133–41. 2009.
[402] Saito Y, Doi Y. Microbial synthesis and properties of poly(3- [407] Zoller DL, Johnston MV. Microstructures of butadiene copolymers
hydroxybutyrate-co-4-hydroxybutyrate) in Comamonas determined by ozonolysis/MALDI mass spectrometry. Macro-
acidovorans. International Journal of Biological Macromolecules molecules 2000;33:1664–70.
1994;16:99–104. [408] Semler JJ, Jhon YK, Tonelli A, Beevers M, Krishnamoorti R, Genzer
[403] Nakamura K, Kamiya N, Sakurai M, Inoue Y, Chujo R. A J. Facile method of controlling monomer sequence distributions in
molecular mechanics study on conformations of bacterial random copolymers. Advanced Materials 2007;19:2877–83.

Você também pode gostar