You are on page 1of 25

Materials Science & Engineering C 92 (2018) 1092–1116

Contents lists available at ScienceDirect

Materials Science & Engineering C


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

Review

Recent advances in the development of biodegradable PHB-based T


toughening materials: Approaches, advantages and applications

Jayven Chee Chuan Yeoa,b, Joseph K. Muiruria,b, Warintorn Thitsartarnb, Zibiao Lib, ,

Chaobin Hea,b,
a
Department of Materials Science and Engineering, National University of Singapore, 9 Engineering Drive 1, 117576, Singapore
b
Institute of Materials Research and Engineering, Agency for Science, Technology and Research (A*STAR), 2 Fusionopolis Way, Innovis #08-03, 138634, Singapore

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

Keywords: Polyhydroxybutyrate (PHB) is a natural biodegradable polymer that is produced by many types of bacteria as an
Biodegradable intracellular energy storage material. Due to its numerous advantages such as biodegradability, biocompat-
Polyhydroxybutyrate ibility, availability and with physical properties comparable to petroleum-based thermoplastics, PHB is a po-
Material toughening tential substitute in biomedical and packaging fields. However, several physical drawbacks, such as high pro-
Polyester
duction cost, thermal instability, and poor mechanical properties, due to secondary crystallization and slow
nucleation rate, limit its competition with traditional plastics in industrial and biomedical applications. Thereby,
many attempts have been employed to improve the material performance of toughened PHB so as to achieve
greater competitiveness and sustainability. In this review, the most recent developments of PHB-based tough-
ening materials are discussed with respect to their approaches and strategies, which includes: drawing and
thermal treatment, blending with materials from natural sources and synthetic polymers, as well as forming
reinforced composites with natural fibers and inorganic fillers. The alternation of PHB chemical structure to form
various types of functional copolymers with enhanced materials performance is also summarized. The expanded
utilization of these newly developed sophisticated PHB materials as engineering materials and the biomedical
significance in different domains are also addressed.

1. Introduction single-use plastic bag in most retail stores back in the 1990s.
Subsequently, countries in Africa, Asia, Ireland, and the rest of Europe
Over the past few decades, negative environmental impacts caused gradually introduced bans or enforcement of plastic bag consumption
by petroleum-based polymers coupled by the uncertainty of crude oil tax and levy; which contributed to reducing the overall consumption
prices and depleting petroleum resources has been a major concern. At significantly [3]. To address the root cause of environmental pollution,
the end of 2015, it was estimated that 8300 million metric tons (Mt) of a viable approach is to use bio-related polymers as substitutes to replace
virgin petroleum based polymers for resins, fibers, and additives were petroleum-based polymer, largely due to their excellent eco-friendly
produced in eight industrial sectors. Nearly 6300 Mt of plastic waste attributes.
had been produced, of which 12% was incinerated, 9% was recycled, One of the most promising biopolymers is polyhydroxybutyrate
and 79% was gathered in landfills or the natural environment, as shown (PHB), a biogenic short chain polyhydroxyalkanoate (PHA) polymer
in Fig. 1 [1]. In addition, the main global source in 2015 was PP and PE, produced by fermentation process; first discovered by Lemoigne in the
contributing 46% to the global production and 63.4% in packaging 1920s as an intracellular energy and carbon storage material accumu-
industry as illustrated in Fig. 2. Based on the present production and lated in various microorganisms, such as bacteria Alcaligenis euterophus,
waste management trends, approximately 12,000 Mt of non-biode- Bacillus, and Pseudomonas [4,5]. PHB exhibits remarkable mechanical
gradable polymers waste were expected to be disposed of in the landfill properties, comparable only to 100% meso diads-polypropylene (PP)
by 2050 [2]. and polyethylene (PE). Fig. 3 shows the chemical structure of PHB. Its
Realizing the downside of these unfavorable polymers, many exceptional stereo-chemical regularity of the structure leads to a highly
countries have initiated strategies and policies targeting this global crystallized homopolymer with crystallinity up to 70%, contributing to
issue, for instance, Germany and Denmark were among the first to ban its excellent mechanical properties; high elasticity modulus of around


Corresponding authors.
E-mail addresses: lizb@imre.a-star.edu.sg (Z. Li), msehc@nus.edu.sg (C. He).

https://doi.org/10.1016/j.msec.2017.11.006
Received 16 October 2017; Received in revised form 3 November 2017; Accepted 11 November 2017
Available online 13 November 2017
0928-4931/ © 2017 Elsevier B.V. All rights reserved.
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 1. Global production, use, and fate of polymer resins,


synthetic fibers and additives from 1950 to 2015; in million
metric tons.
Copyright (2017) Science Advances [1].

Fig. 2. (a) Types of plastics produced worldwide (in per-


cent), and (b) Plastic use in the packaging industry in the
world in 2015.
Copyright (2017) Royal Society of Chemistry [2].

CH3 O inherent physical aging effect, based on secondary crystallization which


lead to embrittlement [23], (2) slow crystallization rate and low nu-
cleation density promotes formation of large spherulites lead to easier
crack and fracture, evident from the low elongation at break of 5–7%
H [24], (3) thermal instability due to narrow thermal processing window,
O OH where PHB degrades via random chain scission on the ester bond in the
temperature range of 170–200 °C [25,26] and lastly, (4) high produc-
Fig. 3. Chemical structure of PHB.
tion cost which limit its competitiveness in industrial and commercial
applications [27–29].
Many strategies have been devised to toughen PHB such as mod-
2.5–3 GPa and tensile strength at break of 35–40 MPa. In addition, the ification through drawing and thermal treatment, blending with ma-
lamellar structure contributes to superior gas barrier properties with its terials from natural sources and synthetic polymers with suitable mo-
water vapor permeability at about 560 g μm/m2/day, making it sui- lecular structures; inclusion of natural fibers or rigid fillers to form
table for low-end food packaging applications [6–15]. reinforced composites and lastly, modification by chemical functiona-
Furthermore, PHB is biodegradable and biocompatible, which in lization as illustrated in Fig. 4. The enhancement in toughness and
turn dictates its environmental fate in terms of ecotoxicity and human flexibility are normally at the expense of material stiffness and strength
toxicity [16]. As such, PHB has found valuable applications in tissue [30–34]. Therefore, mutual enhancement of these properties has been a
engineering and other biomedical related applications, such as surgical grand challenge thus far and currently, this challenge has attracted the
sutures, thermogels as a controlled-release drug delivery vehicle, sur- interest of many research groups. However, there are limited reviews
gical meshes, wound dressing and absorbable nerve guides, tissue on PHB toughening. In this contribution, we intend to provide a com-
scaffolding for bone and nerve regeneration, cardiovascular and carti- prehensive review of PHB toughening materials including the future
lage support [17–22]. perspectives of PHB biopolymer. Table 1 is a summary of various
While PHB has many desirable properties, its widespread applica- toughened PHB systems which show promising applications in con-
tions have been limited due to myriad challenges, which includes (1) struction, packaging and biomedical applications.

1093
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 4. Graphical illustration of the different PHB mod-


ification approaches.

2. PHB toughening through drawing and thermal treatment crystallinity. The two-step drawn (600% + 1000%) and annealed fiber
showed an overall enhancement compared to as-spun native PHB,
Crystallization and the size of the crystals have a great impact on the reaching a tensile strength of 1.3 GPa, elongation at break of 35% and
mechanical and thermal properties of polymers. PHB's exceptional Young's modulus of 18.1 GPa. The strengthening effect was due to the
stereochemical regularity and slow nucleation density results in the formation of β-form lamellar crystals being restrained in the 2D
formation of very large spherulites which promote inter-spherulitic amorphous chains between α-form (21 helix conformation) crystals,
cracking. In addition, secondary crystallization ascribed to crystal generating a planar zigzag conformation which increased the polymer
perfection occur over time, where the amorphous segments are re- chains alignment, contributing to the high mechanical properties as
stricted in between the crystalline segments lead to a lamellar structural shown in Fig. 5 [38,39].
reorganization, causing embrittlement which severely reduces its me- Kabe et al. fabricated annealed melt-spun fibers of PHB with ultra-
chanical performance [35]. Therefore, the combination of drawing high-molecular-weight PHB (UHMW-PHB) at the composition of 5/95
which remodels the molecular chains orientation along the drawing and 10/90 w/w% by similar cold-drawing methods. The annealed fiber
direction and thermal annealing at elevated temperatures and re-aging of the 5/95 blend showed improved performance in tensile strength,
at room temperature could expunge secondary crystallization, im- Young's modulus, and elongation at break achieving 242 MPa,
proving overall toughness and ductility [36,37]. In this section, various 1.50 GPa, and 88% respectively, comparable to poly(ethylene ter-
formulation of PHB that were drawn and annealed are discussed. ephthalate) (PET). The wide-angle X-ray diffraction results indicated
For example, a group of researchers from RIKEN Institute reported a that UHMW-PHB act as nucleating agent and addition of small amount
series of novel drawing and annealing processes to mitigate negative caused the formation of β-form in blend films. [40] The combined
effects of aging and restore toughness by tuning the draw ratios and process of fiber drawing and annealing also showed a significant effect
annealing temperatures. The resulted ultra-high-molecular-weight PHB on the strengthening of PHB-co-HHx fiber [41]. In addition, Lee et al.
film (UHMW-PHB) which was drawn by uniaxial one-step loose showed that annealing allows two immiscible blends to bind together,
hanging weight method at around 160 °C silicone oil, followed by an- obtaining a single Tg. In this work, the authors utilized PHB and
nealing at 160 °C for 2 h showed a remarkable increase of elongation at epoxidized natural rubber (ENR-50), in which the reaction mechanism
break from 6 to 67% and tensile strength from 77 to 100 MPa but at the involved carboxyl end groups of degraded short PHB chains reacting
expense of Young's modulus as it decreased from 2.3 to 1.8 GPa [38]. with epoxide group of the ENR-50 [42]. Kurusu and his co-workers [43]
Amazingly, the properties were maintained even after 6 months with a investigated the effect of annealing on mechanical properties of pure
higher Young's modulus at 2.5 GPa, tensile strength at 100 MPa and PHB and its blends with poly(ethylene-co-methyl-acrylate-co-glycidyl-
elongation at break remain unchanged. After which, the authors re- methacrylate) (PEMAGMA). Aging occurred in the non-annealed
ported that film annealed at low temperature (100 °C for 2 h) displayed sample just after 18 h, showing a drastic difference in mechanical
higher tensile strength while high temperature (160 °C for 2 h) dis- properties between the sample at 6 h and 1 day. Similar observations
played higher Young's modulus. Second drawing and annealing further were revealed in the DSC and SAXS analysis, where the melting en-
increased the tensile strength and Young's modulus to about 388 MPa dotherms of the samples presented a single peak, 24 h after processing
and 2.94 MPa respectively, accompanied by 25% decrease in elonga- [43].
tion [38]. This clearly indicated that the second phase, PEMAGMA, acted as a
The same research group further improvised the process by using a nucleating agent. The same authors further investigated the effect of
lab-scale extruder to melt spin the fibers, followed by a combination of annealing on the injected molded samples under two conditions (110 °C
cold-drawing using iced water at 4 °C, second drawing at room tem- for 10 min and 125 °C for 1000 min) and proposed a toughening me-
perature and a final thermal annealing process to increase the overall chanism [43]. It was revealed that sample aging resulted in thinner

1094
Table 1
Material characteristic properties of toughened PHB via different strategies and their applications.

Strategies Material type Enhanced thermal and mechanical features Application(s) Ref
J.C.C. Yeo et al.

Drawing and thermal PHB film 1st drawing (DR: 10): σ = 100 MPa, εb = 67%, Marine, agricultural and medical 38
treatment 2nd drawing (DR: 1.5): σ = 388 MPa, E = 2.94 MPa
UHMW-PHB fiber 2-step drawing (DR: 60): E = 8.1 GPa, σ = 1.3 GPa, εb = 35% Fishing line and suture 39
UHMW-PHB/PHB fiber 1-step drawing (DR: 12): E = 1.5–2.51 GPa, σ = 211–242 MPa, ɛb = 74–88%, Tm = 170 °C – 40
PHB/ENR film Tg = −14.1 °C – 42
PHB/P(E-MA-GMA) film ISunotched = ~ 1100 J/m, εb = 14–17%, σUTS = 17 MPa, E = 1.2–1.3 GPa Agricultural and food packaging 43
PHB/P(E-MA-GMA)/TEC film ISunotched = ~ 180 J/m, εb = 2.2%, σUTS = 9.4–11 MPa, E = 0.8–1.1 GPa, Tg ~ 19 °C Agricultural and food packaging 44
PHB/P(E-MA-GMA)/TBC film
PHB/TBC/BN fiber σUTS = 215 MPa, εmax = 40% Textile manufacturing and medical 45
Blending PHB/soluble potato Starch Tg ~ 63–87 °C, Tm = 165–168 °C, σUTS = 31.5 MPa Paper coating & cardboard 46
packaging
PHB/maize starch Tg ~ 37 °C, Tm = 165–168 °C, σ= 31.5 MPa – 47
PHB/destructured starch σ = 20 MPa, εb = 0.8%, E = 4 GPa Medical 488
PHB/maize starch σ = 18.7–22.0 MPa, εb = 2.62%, E = 0.37–1.32 GPa Agricultural & marine 49
PHB/starch Tg,1 = − 58.1 °C, Tg,2 = − 19.2 °C, Tm = 141–159.1 °C, σ = 3–11.3 MPa, ε = 2.3–15.5%, – 50
E = 0.12–0.21 GPa
PHB/starch/glycerol σ = 5.9 MPa, Tear strength = 45 kJ/m2 – 51
PHB/EVA/starch/MA σ = 15–20 MPa, εb = 50%, Work = 0.09 Nm – 52
PHB or PHB-g-AA/starch Tm = 167.8–171.3 °C, σ = 17.5 MPa Agricultural, marine & medical 53
PHB/PVAc modified starch Tm = 160–170 °C, σ = 6.4–15.4 MPa, εb = 2.9–21%, E = 0.33–1.06 GPa – 54
PHB/cellulose nanocrystals Tm = 150–172.5 °C, σ = 20–30 MPa, εb = 1.4–2.2%, E = 3.5–4 GPa Packaging 55
(CNC)
PHB/cellulose nanowhiskers Tm = 156–163 °C, σ = 17 MPa, εb = 300%, E = 0.1–0.6 GPa Medical & agricultural 56
(CNW)
PHB/cellulose/DCP Tm = 155–171 °C, σ = 25.9 MPa, εb = 13.2%, E = 5.5 GPa Packaging and semistructural 57

1095
profiles
PHB/bacterial cellulose (BC) εmax = 10.5%, σmax = 91.4 MPa Tissue engineering and medical 58
PHB/hydroxethyl cellulose Tm = 155–175 °C – 60
acetate
PHB/ethyl cellulose (EC) Tm = 170–175.5 °C – 61
PHB/ethyl cellulose (EC) IS = 7 kJ/m2, εb = 1.5%, σ = 33.3 MPa, Tm = 163.6–175.9 °C Biotechnology 62
PHB/CAP and PHB/CAB εb = 0–5%, σ = 2–10 MPa – 63
PHB/insoluble lignin/or Tm = 172–175 °C Diverse applications 66
holocellulose
PHB/crude lignin and PHB/ Tm = ~ 143.9–175 °C Agriculture, cosmetics and packaging 67
soda lignin
PHB/lignin Tm ~ 151–174 °C Packaging & pharmaceutical 68
PHB/chitosan εb = 43.3–82.9%, σ = 3.4–4 MPa, E = 4.9–8.7 MPa Tissue engineering 69
PHB/chitosan/or α-chitin Tm ~ 155–170 °C Biomedical 70
PHB/chitin Tm ~ 166–170 °C Biomedical, agricultural, and paints 71
PHB/PHBV film E′ = 3.8–4.2GPa – 72
PHB/PHBHHx film εb = 106–120%, Tc ~ 50 °C Tissue engineering 74
PHB/PLA/TBL Tm ~ 165–167 °C, σ = 9–14.4 MPa, ε = 12.5–193.2%, E = 0.07–0.5 GPa – 84
PHB/PDLLA Tm ~ 147–165 °C, σ = 35–45 MPa, σFlex = 45–65 MPa, EFlex = 2–3.5 GPa, IS = 15–35 kJ/m2 Medical 85
PLA/PHB Tm ~ 155–165 °C, σ = 27.5–37.5 MPa, ε = 5–20% Medical 86
PHB/Sc-PLA Tm ~ 210–230 °C Medical devices 87
PHB/PCL film εb = 7.09%, σ = 1.43 MPa, E = 7.3 MPa Drug delivery and biomedical 96
(continued on next page)
Materials Science & Engineering C 92 (2018) 1092–1116
Table 1 (continued)

Strategies Material type Enhanced thermal and mechanical features Application(s) Ref

2
J.C.C. Yeo et al.

PHB/PCL composite IS = 10.6 J/m , εb = 11.2–1000%, σ = 17.3–21.4 MPa, E = 690–1643 MPa, Food packaging 97
EFlex = 716–1605 MPa, σFlex = 27–38.5 MPa, Tm = 171.2–172.7 °C
PHB/PCL/DCP composite εb = 21.3%, IS = 6.3 kJ/m2, Tm = 55.5 °C Food packaging 98
PHB/PPC/PC film IS = ~ 45 kJ/m2 Biomedical 101
PHB/PPC/PVAc/ATEC films εb = 335–475%, Tc = 63 to 85 °C, Tg = − 8 °C Air and water filter, wound dressing, 102
drug delivery, surgical suture and
mat fiber used in children diapers
PHB/PBS/DCP film IS = 3.5–5.5 kJ/m2, ISunotched = 17–82 kJ/m2, ɛb = 4–15%, σflex,yield = 54–60 MPa, Diverse applications 103
σys = 37–40 MPa
PHB/PCL-b-PEG εb = 350–382%, Tm,1 = 174.8–177.9 °C, Tm,2 = 157–159.4 °C Biomedical 107
PHB/PEO Ti,1 = 216–239 °C, Ti,2 = 351–357 °C Biomedical 108
PHB/PVA film εb = ~ 210%, Tm,1 = 178–186 °C, Tm,2 = 123–132 °C Packaging, agricultural and 110
biomedical
PHA/PVAs/peroxide composite εb = ~ 211–557%, IS = 21.3–42.7, tear strength = 10–80 g/mil, E = 494–1241 GPa, Automotive, consumer durable, 113
σb = 21.1–24.5 MPa construction, electrical, medical,
packaging products
PHB/NR/MP composite IS = 124 J/m Structural 114,
PHB/EPR/MP composite 115
PHB/PP composite IS = 22.5–29 J/m, εb: 4.5% E = 1.5–1.88 MPa, σ = 24.5–27.5 MPa, shore D hardness = 61–65, – 118
Tm = 177–180 °C
PHB/PP-PHB/manganese εb = 8%, σ = 35 MPa. EFlex = 2.32 GPa, σFlex = 50 MPa Cosmetic containers; cell phones; 120
stearate composite laptops; and packaging
PP/PHB/compatibilizer* IS = 80 J/m, εb = ~ 80% – 121
*PP-MAH, P(E-MA), P(E-GMA)
& P(E-MA-GMA)
LDPE/PHB film Tm,1 = 105.9–106.8 °C, Tm,2 = 171.8–173.2 °C Packaging materials 123

1096
PHB reinforced composite PHB/flax fiber/TDP Tm = 168–172 °C, E′ = ~ 13 GPa Automotive (door and rear-parcel 125
shelf panels)
PHB/flax E = 8 GPa, σ = 40 MPa, IS = 135 J/m Automotive 126
PHB/agave IS = 34.4 J/m, σ = 14 MPa, E = 770 MPa, EFlex = 2154 MPa, σflex = 17 MPa, E′ = 3733 MPa, Construction 128
Tg = 11.1C, Td = 25.2 °C
PHB/Hemp fabric σwarp = 55.9 MPa, εb,weft = 9%, Ewarp = 5.47 MPa, EFlex = 5.05 GPa, Tm = 117 °C Construction 129
PHB/jute, PHB/hemp, PHB/ σmax = 17 MPa, EFlex = 4 GPa, IS = 0.25 J Biomedical 130
lyocell
PHB/carnauba fibers Tm = 117 °C, σUTS = 27.5 MPa, EFlex = 3.25 GPa Automotive 131
PHB/cellulose fibers/DBEEA Tm = 158.1 °C Packaging 132
PHB/curaua fibers/TEC Tm = 117 °C, σUTS = 25 MPa, εmax = 14%, E = 178 MPa, IS = 70 J/m Automotive and appliances 133
PHB/SWCNTs E = 11.74 GPa, hardness = 0.35 GPa Tissue engineering 136
PHB/acid-treated or alkylated Tm = 171–177C, E′ = 4000–5200 MPa Medical and pharmaceutical 137
MWCNTs
PHB/NaMMT or OMMT (30B) E = 3.44 GPa, σ = 27 MPa Biomedical, agricultural, and 138
packaging
PHB/C18MMT E′ = 552 MPa, Tg = 29.5 °C, Tm,1 = 156.4 °C, Tm,2 = 166.13 °C Tissue engineering 139
PHB/OMMT E = 15.5 MPa, σ = 10 MPa, εb = 0.72%, Tm,1 = 172.9 °C, Tm,2 = 169.3 °C Packaging 140
PHB/A-fnSiO2 Tm,1 = 158 °C, Tm,2 = 167 °C, pencil hardness = HB Tissue engineering, medical devices 141
PHB/GNP and food packaging
PHB/MGO E = 83.6–114.6 MPa, σ = 20.7–37.4 MPa Electrical 142
(continued on next page)
Materials Science & Engineering C 92 (2018) 1092–1116
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

boron nitrate, MA: maleic anhydride, AA: acrylic acid, DCP: dicumyl peroxide, PHBV or PHB-co-HV: poly(3-hydroxybutyrate-hydroxyvalerate), PHBHHx or PHB-co-HHx: poly(3-hydroxybutyrate-co-3-hydroxyhexanoate), TBL: tributyrin, PLA:

ethylene oxide, PVA: polyvinyl alcohol, EPR: ethylene-propylene rubber, NR: natural rubber, MR: maleated rubber, MP: maleated polybutadiene, PP: polypropylene, HDPE: high density polyethylene, PA6: polyamide 6, ABS: acrylonitrile butadiene
polylactic acid, PDLLA: poly-DL-lactic acid, PCL: polycaprolactone, PPC: poly(propylene carbonate), PC: polycarbonate, PVAc/ATEC: polyvinyl acetate/acetyl triethyl citrate, PBS: polybutylene succinate, PEG: polyethylene glycol, PEO: poly-
PHB: polyhydroxybutyrate, UHMW-PHB: ultrahigh molecular weight polyhydroxybutyrate, ENR: epoxidized natural rubber, P(E-MA-GMA): poly(ethylene-co-methyl acrylate-co-glycidyl methacrylate), TEC: triethyl citrate, TBC: tributyl citrate, BN:

epoxy-1,2,3,6-tetrahydrophthalic anhydride, MAc: maleic acid. DR: drawing ratio, ɛb: elongation at break, σ: tensile strength, σb: tensile strength at break, σUTS: ultimate tensile strength, σFlex: flexural strength, E: Young's modulus, EFlex: flexural
styrene, PP–MAH: poly(propylene-g-maleic anhydride, P(E-MA): poly(ethylene-co-methyl acrylate), P(E-GMA): poly(ethylene-co-glycidyl methacrylate), DBEEA: bis[2-(2-butoxyethoxy)ethyl] adipate, SWCNTs: single-walled carbon nanotubes,
MWCNT: multi-walled carbon nanotubes, NaMMT: natural montmorillonite, OMMT: organo-modified montmorillonite, A-fnSiO2: amino-functionalized nano-silica, GNP: graphene nano-platelets, MGO: modified graphene oxide, ETA: exo-3,6-
lamellae and imperfect crystals between stable crystals, with a more

106
148
149

150

151
152

153
154
155
156
Ref
interfacial area between crystalline and amorphous regions as shown in
Fig. 6. Annealing treatment reduced the chain density in the amorphous
regions, the defective crystals in the melt, whereas, the stable crystals
become thicker with increasing L, lc and la, enhancing the ability to
dissipate energy [43]. Longer annealing time produced a homogeneous
microstructure with higher crystallinity, lamellae thickness and crys-

Packaging and drug delivery


Cardiovascular engineering
talline − amorphous interfacial area, which led to remarkable increase
in impact strength from 400 to ~1000 J/m and 6-fold increase in
Tissue engineering

elongation from 2.5% to ~ 17% compared with the aged sample [43].
The effect of aging and annealing on PHB blends with three industrial
Application(s)

plasticizers were subsequently investigated, including triethyl citrate


Biomedical

Biomedical

Biomedical

Biomedical

(TEC), tri(ethylene glycol) bis(2-ethyl hexanoate)(TEG-EH) and tributyl


citrate (TBC) [44]. The results showed an increase in impact strength


for all the blends with plasticizers. However, both the elongation at
break and tensile strength decreased due to cracks and voids emerging
from contraction of the amorphous interlamellar layer in the matrix.
Edry state = 616.4 MPa, Ehydrated state = 342.7 MPa, σb,dry state = 18 MPa, σb,hydrated state = 14.5 MPa
εb = 280–448%, σUTS = 9.2–10.5 MPa, Tm,1 = 126–131 °C, Tm,2 = 139–141 °C, Ti = 203–228 °C

The plasticizer was forced to leach out, which eventually led to dete-
riorating of properties instead of toughening effect [44].
A recent investigation done by Hufenus et al. [45] focused on the
effect of additives, various filament drawing setups and stress annealing
on PHB fiber with an addition of commercial additives, boron nitride
εb = 1912%, σys = 10.2 MPa, σb = 10.7 MPa, E = 120 MPa, Ti = 227 & 350 °C

(BN) as a nucleating agent and tri-n-butyl (TBC) as a plasticizer to


modulus, E′: storage modulus, Tm: melting temperature, Tg: glass transition temperature, Ti: onset decomposition temperature, Td: decomposition temperature.

improve its melt-spinning. It was observed that crystallization was


stabilized by the nucleating agents, and the damping viscosity fluc-
tuations were achieved from plasticizers, thus, minimizing thermal
Tg,1 = − 31.2-23.5 °C, Tg,2 = 11.2–32.8 °C, Tm = 123.4–130.2 °C
2nd drawing (DR = 25): σ = 150 MPa, εb = 110%, E = 1.5 GPa
1st drawing (DR = 50): σ= 80 MPa, εb = 258%, E = 0.87 GPa

degradation. On the other hand, the introduction of the modified draw-


off unit enhanced the initial crystallization, thus, suppressing the post-
crystallization [45]. The resulted drawn fibers crystallized in an or-
ientation prior to a complete solidification, forming highly oriented
amorphous phase (β-form) being kinetically confined between the
aligned α-phase crystalline as observed by the wide-angle X-ray dif-
εb,dry state = 1090%, εb,hydrated state = 1962%,
Enhanced thermal and mechanical features

Ti = 223.4 °C, Td = 270.3 °C, Tm = 165 °C

fraction (WAXD). The fiber exhibited the highest tensile strength of


εb = 21%, Tg = 4.8 °C, Tm = 208.9 °C

215 MPa at annealing condition 110 °C for 60 min, a 39% increase from
neat fiber but with a decrease of 16% in tensile strain [45]. While an-
nealing seems to be a remedial and effective method in the reduction of
Tm = 145.4 °C, Tg = − 2.2 °C

the brittleness of PHB-based materials and elimination subsequent


εb = 970%, Tg = − 9 °C
εb = 850%, Tg = − 4 °C

aging of the samples, the main concern still lies in the compensation of
toughness and elongation at break. In addition, the intricacy and high
cost remain a challenge for up-scaling and adoption from the laboratory
scale to industrial scale.

3. PHB toughening through blending

Blending is by far the easiest, most effective and economical ap-


proach for obtaining new materials with improved physical and me-
PHB-co-PCL-PEG-PCL urethane

chanical properties, in which the drawbacks of the parent components


can be altered and specially tailored by choosing and varying the
compositions of the blend and preparation parameters. More im-
PHB-co-PEG urethane
PHB-co-PEG urethane

PHB-co-atactic PHB

PLA-co-PHB-co-PLA
PHB-co-PLA-co-PCL

portantly, blends of PHB with biodegradable polymers have been of


PHB-co-HHx film
PHB-co-HHx film

much interest, in line with the concept of sustainability and eco-


Material type

friendly. The concept alludes to the fact that the choice of materials,
PHB-g-MAc
PHB-co-HV

PHB-g-ETA

both matrix, and fillers, should be sourced from renewable resources. In


this section, various blends of PHB with natural or synthetic sources
and also non-biodegradable polymers are discussed.

3.1. Blending with materials from natural sources

3.1.1. Starch
Chemical modification

Starch is an amylose-amylopectin storage polysaccharide as shown


in Fig. 7, which is abundant, renewable and biodegradable. This bio-
Table 1 (continued)

filler has found applications in textile industry, as sizing/stiffening


agents and, as adhesives in papermaking. Blending of PHB with starch
Strategies

has been widely investigated with the sole aim of exploiting inherent
properties of starch, in order to overcome PHB limitations and over the
years, remarkable studies have been done on physical mixtures of

1097
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 5. Schematic illustrations showing the pro-


posed mechanism for generating β-form struc-
tures in blend and PHB films. (A) Initial crystal-
lization of lamellar crystals (α-form). (B)
Generating the β-form in an amorphous region
during further crystallization. (B′) Movement of
lamellar crystals during further crystallization.
The β-form was formed from tie-chains fixed
between lamellar crystals during annealing.
Copyright (2012) American Chemical Society
Macromolecules [40].

Fig. 6. Proposed mechanism for the lamellar structure al-


teration after the annealing treatment.
Copyright (2014) Wiley Society of Chemical Industry [44].

OH OH OH peaks in the range of 165–168 °C. From the various proportions studied,
O O O an optimum tensile strength was reported at PHB/starch ratio of 7:3, an
OH O OH increase of about 72% compared to neat PHB's strength of 18.29 MPa
O
OH
[46]. It is worth noting that blending of 30 parts of inexpensive starch
HO OH OH OH OH
n to 70 parts of costly PHB, results in substantial cost reduction. Ad-
ditionally, these blends were thermally stable at about 203–223 °C,
Fig. 7. Chemical structure of starch.
which was 30 °C lower compared to PHB/thermoplastic starch blends.
The newly affordable bioplastics, with uncompromised properties,
starch and PHB. widens PHB applications landscape, in paper coating and cardboard
For instance, in 2002, Godbole et al. [46] investigated the mis- packaging fields [46]. In a similar study, Zhang et al. [47] examined
cibility, morphology and thermomechanical properties of PHB/Starch PHB properties by blending with dissimilar maize starch, one con-
blends, prepared through solvent casting technique. Thermal char- taining 70% amylase (starch 1) and, another containing 72% amylo-
acterization of the blends revealed single glass transition temperature pectin (starch 2). In this work, similar PHB/starch ratio of 70:30 was
(Tg) in the range of 63 °C–87 °C, a strong suggestion that all the blends employed and processed through melt compounding. The results re-
were miscible, in all the proportions tested [46]. On the other hand, vealed the dual action of starch, both as filler and nucleating agent in
DSC analysis exhibited crystalline nature of the blends, with melting PHB/starch blends, which led to considerable reduction in the size of

1098
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

PHB spherulites [47]. As a result, thermomechanical properties were Besides the unmodified binary PHB/starch blends, researchers have
significantly enhanced, especially, for PHB and high amylose starch. further studied complex PHB/starch blend systems, including ternary
From dynamical mechanical thermal analysis (DMTA), PHB/starch 1 blends. For instance, Ma et al. [52] investigated the physical properties
blend exhibited a lower and narrower tan delta peak, compared to of PHB/EVA/starch blends by reactive compatibilization using maleic
PHB/starch 2 blends. These results indicated that PHB/starch 1 blend, anhydride (MA) and benzoyl peroxide (BPO). In this work, EVA/starch/
dissipated less energy due to strong interfacial bonding. This was as- MA/BPO/glycerol were pre-compounded, and then blended with PHB
cribed to uninterrupted hydrogen bonding between PHB and high [52]. Fig. 8 shows SEM images gelanatized starch, uncompatibilized
amylose starch, which is linear but coiled, compared to branched EVA/starch, compatibilized EVA-g-starch and proposed reaction me-
amylopectin starch (starch 2) [47]. In addition, the Tg of PHB/starch 1 chanism for compatibilization. As shown, starch particles in compati-
was 2 °C lower than that of PHB/starch 2 (Tg ~ 37 °C), which was at- bilized blend are homogeneously dispersed in EVA matrix, which was
tributed to linear structure of amylose, compared to cross-linked amy- then introduced into PHB to produce PHB/EVA-g-starch blends [52].
lopectin. Furthermore, PHB/starch 1 blends had higher thermal stabi- The authors reported that tensile strength increased with MA con-
lity, as a result of PHB's carbonyl and starch's hydroxyl groups' tent up to 0.09%, but later decreased due to change in phase mor-
interaction, which inhibited ring ester decomposition (chain scission) phology and size of starch particles and EVA domains. Additionally, the
mechanism [47]. co-continuous morphology resulted in an increase in elongation at
In an earlier work, Koller and Owen [48] studied the structure and break and, hence improved toughness of the blends. The enhanced
mechanical properties of maize starch-filled PHB, as well as PHB/HV toughness was ascribed to the improved affinity of EVA-g-starch and
copolymer prepared by compression molding at 190 °C for PHB/starch. PHB, compared to PHB/starch blend [52]. Cavitation, fibrillation and
In this work, PHB/starch exhibited brittle failure at below 1% strain, matrix yielding were identified as the major toughening mechanisms in
indicating that starch caused PHB embrittlement, with increased reactively compatibilized PHB/EVA/starch blend system. On the other
modulus and lower strength [48]. In addition, PHB filled with de- hand, thermal properties of PHB were somehow inert to reactive
structured starch (treated in water under heat and shear) had better compatibilization process. However, the Tg of EVA component in-
properties, especially, modulus compared to native starch granules. The creased by 9 °C with 36% increase in MA content, attributed to mis-
effect was attributed to small particles size, asymmetrical shape and cibility and morphological changes, arising from composition array of
surface roughness of destructured starch [48]. Similarly, Thiré et al. EVA domains [52]. In another interesting study by Liao and Wu [53],
[49] investigated the effect of starch on properties of compression PHB-g-acrylic acid (PHB-g-AA) was synthesized by grafting acrylic acid
molded PHB/starch blends at different starch contents, up to 50%. The (AA) to PHB, initiated BPO, and subsequently melts blended with starch
results showed that thermal stability was not affected by the addition of for comparison with PHB/starch blend, in terms of thermomechanical
starch, in all compositions. The onset degradation temperature was properties. The mechanical properties of PHB-g-acrylic/starch blend
about 274 °C. On the other hand, elongation at break and tensile were found to be superior to PHB/starch blend up to 50% starch con-
strength decreased with increased starch content, whereas, Young's tent [53]. The enhancement in properties was ascribed to the reaction
modulus was unaffected up to 30% starch content [49]. Further, the between hydroxyl groups in starch and carbonyl group in PHB-g-AA
modulus declined rapidly by about 70%, resulting in flexible materials, and hence, improved compatibility in these blends compared to PHB/
and this was as envisaged and, attributed to the ingrained rigidity of starch blends. In addition, PHB-g-AA/starch blends had improved
PHB. The overall poor mechanical properties of the blends were linked processability compared to PHB/starch blends, due to low melt visc-
to inadequate interface adhesion and heterogeneous morphology of osity [53]. The authors reported that the melting temperatures (Tm) of
PHB/starch granules blends [49]. all the blends decreased with increased starch content up to 50%, and
Innocentini-Mei and co-workers [50], explored the effects of che- were in the range of 170 °C–173 °C and 168 °C–172 °C, for PHB/starch
mical modification of starch on PHB's thermomechanical properties, by and PHB-g-AA/starch, respectively [53].
comparing natural starch, and modified starch i.e. starch adipate and Polyvinyl acetate (PVAc)-modified starch has also been synthesized
urethane-grafted starch at varying proportions. In this study, it was and used to improve the thermal and mechanical properties of PHB. In a
discovered that natural starch and its derivatives, lowered the melting recent work, Don et al. [54] grafted vinyl acetate (VAc) to potato
(Tm) and glass transition temperatures (Tg) in all the binary blends, starch, initiated by ceric ion, to obtain PVAc-modified starch, which
compared to pure PHB [50]. However, much lower Tm and Tg values was subsequently blended with PHB. In this study, it was discovered
were observed for PHB/modified starch blends, ascribed to internal that all PHB/PVAc-modified starch blends were miscible at all combi-
plasticization. On the other hand, all the PHB/natural starch and PHB/ nations [54]. This was ascertained by single glass transition tempera-
starch adipate blends were stiffer and brittle (low tensile strength at tures (Tg) for the blends, in both DSC and DMA analysis. Furthermore,
break), in spite of insignificant changes in elongation at break [50]. TGA results showed improvement in thermal stability with the addition
Unlike other studies, Lai et al. [51] discussed the possibility of enhan- of PVAc-modified starch, with a three-stage degradation profile, com-
cing properties of thermoplastic starch (TPS) by blending with PHB as a pared to single degradation pattern for PHB at 288 °C [54]. In essence,
filler and TPS as a matrix. In this work, three kinds of gelatinized starch the improved thermal stability of PHB by PVAc-modified starch meant
(i.e. potato starch, corn starch and soluble potato starch), were em- an enlarged processing window for PHB. In terms of mechanical
ployed and blended with certain amounts of PHB, in order to tune the properties, PHB/PVAc-modified starch blends were found superior in
performance of TPS [51]. Significant changes in mechanical properties toughness, compared to virgin PHB. This was ascribed to improved
for the blends were noted. For instance, with a low degree (25% gly- compatibility and leathery nature of PVAc-modified starch [54].
cerol) gelatinized potato starch/7% PHB, the tensile strength increased
1100% and was ascribed to [1] high strength of PHB and, [2] reason- 3.1.2. Cellulose and its derivatives
able low levels of polymer segregation in TPS matrix. On the contrary, Cellulose is the most abundant structural polysaccharide on earth,
33% glycerol-TPS, exhibited poor strength, even at increasing PHB mainly found in plants. It consists of glucosidic linkages, where C-1 of
content, and this was attributed to high plasticizing effects [51]. one glucose unit joins C-4 of the next glucose molecule as shown in
Moreover, TPS (potato starch) had exhibited better mechanical prop- Fig. 9. On a nanoscale, it forms nanofibrils with unique strength
erties improvement compared to TPS (soluble starch), which was linked properties that make up the structure of the plants. As a result of these
to a higher molecular weight of potato starch and it's highly bonded unique properties and nanostructure form, cellulose nanocrystals
cohesive structure. Additionally, inclusion of PHB into TPS conferred (CNCs) have been widely investigated as bio-fillers for PHB polymer.
significant thermal stability to the blends, a great potential for pro- For example, Seoane et al. [55] thoroughly studied the effects of
cessability of new environmentally benign plastics [51]. CNCs on PHB properties. In this study, DSC cooling scans demonstrated

1099
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 8. SEM images showing morphologies of gelatinized


starch, EVA/starch (50/50 wt%) and EVA/starch (50/50 wt
%) compatibilized with 0.09% MA.
Copyright (2014) Elsevier [52].

reduced by ~9%, compared to PHB. The TGA and DSC results indicated
OH a widened processing window for PHB, by incorporating CNWs. In
OH terms of mechanical properties, toughness (viz. strain at break) of the
O OH
O nanocomposites improved 50-fold with up to 0.45 wt% CNW and 15 wt
O % PEG, with strength remaining almost unchanged at 17 MPa. This was
OH O attributed to preferential chain orientation in the tensile direction.
OH
However, Young's modulus declined remarkably due to reduced re-
OH n sistance to chain flow, arising from capsulation of CNW by PEG. These
materials potentially extend application possibilities for PHB [56].
Fig. 9. Chemical structure of cellulose.
In one recent study, Wei and co-workers [57] reported on in-situ
compatibilized PHB-cellulose materials obtained from free radical
that CNCs were effective nucleating agents for PHB in all the nano- polymerization, initiated by dicumyl peroxide (DCP). Fig. 10 shows the
composites, with an exothermic peak at ~83 °C, compared to ~62.5 °C predicted schematic structures obtained through the generation of free
for pristine PHB. This indicated that CNCs effectively reduced the en- radicals and grafting sites in PHB/cellulose. From this study, the au-
ergy barrier for PHB nuclei formation [55]. Moreover, CNCs were found thors found that grafting was dependent on time and DCP concentration
to favor PHB crystallization, with 1.4% change of in crystallinity. Worth [57]. Consequently, the thermal stability of PHB was enhanced, as a
noting, pristine PHB exhibited single degradation pattern, with max- result of the formation of new bonds. Although mechanical properties
imum DTG peak at ~295 °C, whereas, PHB/CNC nanocomposites had were not reported, chemical cross-linking in PHB-cellulose materials
two-step profile, with lower decomposition temperatures compared to could facilitate effective stress transfer. [57]
pristine PHB. This was attributed to the presence of sulfate and hy- In another work, Zhijiang et al. [58] prepared PHB-bacterial cellu-
droxyl groups on CNC, which assisted chain scission of PHB at elevated lose (BC) biocompatible scaffolds by impregnation of BC into PHB-
temperatures [55]. On the other hand, with up to 6% CNC content, chloroform solution, and subsequent freeze-drying to remove the sol-
tensile modulus and strength of the nanocomposites increased by vent. In this study, it was discovered that incorporation of BC, increased
~ 50% and ~35%, respectively. Furthermore, nanoindentation tests overall crystallinity of PHB in PHB/BC scaffolds. This suggested that BC
exhibited high elastic modulus with CNCs up to 2% and hardness values nanofibrils have crystallization capability influence on PHB molecules.
in the range of 155–165 MPa, which suggested the effective reinforcing On the other hand, tensile strength and elongation at break, increased
effect of CNCs and good CNC-PHB matrix interactions [55]. In a similar 13-fold and 2-fold, respectively [58]. This was attributed to reinforcing
study, S de O Patrício, Patrícia, et al. [56] evaluated the influence of effects of BC in PHB matrix, arising from the inherent high strength of
cellulose nanowhiskers (CNWs) on thermomechanical properties of BC. Similarly, Barud and co-workers [59] also prepared BC/PHB
PHB. In this research, CNWs dispersion in PHB matrix was facilitated by membranes, with up to 90 wt% PHB by solution casting technique. The
polyethylene glycol (PEG) by solution casting method. TGA results authors reported improvement in thermal stability of BC/PHB mem-
showed the enhanced thermal stability of the blends in two stages at branes with maximum temperature of ~360 °C, compared to ~320 °C
228 °C and 295 °C, corresponding to PEG and PHB degradation, re- for pure PHB [59]. In addition, above 50 wt% PHB, overall crystallinity
spectively [56]. Additionally, melting temperatures of the blends were was enhanced, although lower than pure PHB. Young's modulus, tensile

1100
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 10. Schematic illustration showing possible


free radical sites and grafting sites in PHB/
Cellulose blends initiated by DCP. Reprinted with
permission from ref.
Copyright (2015) American Chemical Society
[57].

strength and elongation at break were found to be dependent on BC/ Yamaguchi and Arakawa [63] investigated the structure-properties
PHB combinations, and were reported to be as high 144 MPa, 9.5% and relationships of PHB with four different kinds of cellulose derivatives
15.5 GPa, respectively [59]. (i.e. CAP and CAB), including the mechanical properties. The blends
Interestingly, cellulose being a unique polymer with hydroxyl were prepared by mixing and compression molding at 180 °C, followed
groups offers plenteous room for surface modification on the poly- by crystallization at 40 °C for 5 min. In this work, the authors reported
saccharide backbone, such as, etherification, esterification and so on. A anomalous viscoelastic behavior of the blends observed through DMA.
number of cellulose derivatives have been produced, including cellu- Although, high molecular weight cellulose derivatives CAP46 and
losic derived plastics such as cellulose acetate (CA), ethyl cellulose, CAB37 were brittle, their blends with PHB showed unlikely plastic
cellulose acetate butyrate (CAB), and cellulose acetate propionate deformation [63]. This was attributed to reduced crystallinity of PHB,
(CAP). In recent years, these cellulose derivatives have become of as a result of more entrapped floating chains. These floating chains
particular interest to research groups in tuning the properties of bio- resulted to more amorphous PHB blends, thus, ductile failure mode
degradable PHB polymer. For instance, Zhang and Deng [60] prepared under DMA and tensile tests was reported. These results agreed with the
biodegradable PHB/hydroxyethyl cellulose acetate (HECA) blends by shift in β-relaxation peaks to lower temperature [63]. On the other
solution casting method. Although the authors did not report on the hand, PHB blend with low molecular weight CAB52 and CAB38 ex-
mechanical properties, thermal changes of the blends were clearly ob- hibited low elongation at break (˂1%) and comparable β-relaxation
served. In this regard, melting temperatures of blends with PHB above peak as PHB and PHB-rich blends at about 20 °C. In conclusion, me-
20% was about 175 °C, and were independent of the blend composition chanical properties of PHB/cellulose derivatives blends can be tailored
[60]. Low PHB content (˂20%) blends exhibited depressed equilibrium by fine-tuning the amorphous fraction [63].
melting temperature of PHB, and this was attributed to mesophase-
isotropic phase transition of HECA within the same temperature range.
Furthermore, Tg of PHB in all the blend compositions remained un- 3.1.3. Lignocellulosic
changed at about 8 °C. In addition, HECA content in the blends mark- Plant matter (biomass) is the most abundant resource on earth,
edly affected both nonisothermal and cold crystallization, with a shift to widely used in biorefineries. The biomass is composed of cellulose,
lower temperatures and higher temperatures, respectively [60]. Zhang hemicelluloses and lignin, which have important roles on overall
et al. [61] also explored thermal properties of solution cast PHB/ethyl structural properties of lignocellulosic materials [64,65]. In light of
cellulose (EC) binary blends. In this study, increase in EC content in the this, value addition of plant matter, as additives in polymers, through
blends caused an upward shift in the observed composition-dependent valorization is advantageous. In an earlier study, Angelini and cow-
Tg. However, Tg of PHB remained unchanged at approximately 5–9 °C orkers [66] fabricated PHB blends with crude lignin-rich residues (LRR)
for all the blends. Both, the melting temperatures and crystallinities of and alkaline lignin (AL) through compression molding. In this work,
PHB/EC blends were superior to that of pure PHB. Mechanical prop- LRR had no significant effect on thermal stability of PHB, whereas, AL
erties of these blends were not reported [61]. In a similar work, Chen acted as a strong pro-degrading agent. As anticipated, TGA analysis
and co-workers [62] investigated mechanical properties of PHB/EC revealed that LRR had higher mass loss and lower char yield, which was
blends prepared through melt mixing technique. In this work, PHB/EC due to decomposition of cellulose fraction present [66]. Moreover, the
blend system was reported to be heterogeneous; in which the dispersed Tg values of AL and LRR were 112 and 113 °C, respectively. Tensile and
EC domains acted as melt viscosity modifiers in processing and, as impact tests showed overall decrease in mechanical properties of PHB/
stiffeners under shear flow [62]. The dual character of EC domains LRR blends. For example, PHB/30% LRR blend showed a decrease of
influenced the spherulites structure of PHB, as well as its kinetics of about 60% and 55% in tensile strength and impact strength, respec-
crystallization. For example, inclusion of EC led to (i) reduced growth tively. This was ascribed to presence of defect sites between the phases,
rate of PHB spherulites, (ii) reduced crystallinity and (iii) lamellar de- filler heterogeneity, poor compatibility and weak filler/matrix interface
fects, which subsequently, affected the mechanical properties of PHB. adhesion. On the other hand, flexural modulus of PHB/30% LRR blend
With low EC (˂1 wt%), EC acted both as reinforcing and toughening was about 40% higher than that of pure PHB. This was ascribed to
agent. The impact and tensile strength of blends with 1 wt% EC were flexural modulus being dependent on the combined action of individual
enhanced by 1.22 and 1.15 times, respectively [62]. components in the blend and only slightly sensitive to components'
mutual interfacial interactions [66].

1101
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

The same research group [67] further investigated the effects of separately, admixed, cast on poly(tetrafluoroethylene) (PTFE) dish,
isolated fractions of lignocellulosic (LC) biomass on PHB properties. In dried at RT and under vacuum for a couple of days [70]. From DSC
this work, the authors obtained two fractions i.e. insoluble lignin (IL) characterization, the authors revealed that both α-chitin and chitosan
and hollocellulose (HC), from biorefinery waste and used it as fillers in hampered PHB's crystallization. This observation was ascribed to the
PHB matrix, in which purified LC was the reference [67]. LC 30% w/w reduced PHB's lamellar thickness, intermolecular filler/matrix interac-
was found to have remarkable enhancement on PHB's moduli and tions, resulting in confined and trapped PHB molecules in the blends.
complex viscosity, which was almost independent of time and fre- Compared to α-chitin, chitosan had about 10% more suppressing
quency. These results were supported by SEM observations, showing power, obtained from the heat of fusion (ΔH) of varying blend com-
good dispersion of LC in the matrix that resulted in a tightened co- positions [70]. In another recent study, Khasanah et al. [71] prepared
continuous network, with good filler-matrix interface. However, IL did PHB/Chitin blends through solution casting technique. In this work, the
not promote PHB crystallization compared to HC. This was ascribed to authors explored the effects of intermolecular hydrogen bonding be-
the higher melt viscosity of PHB/LC blends, which hampered the chain tween carbonyl groups of PHB and amine groups of chitin. From the
mobility [67]. DSC cooling scans, the crystallization temperature of PHB was de-
Mousavioun et al. [68] investigated the thermal and rheological pressed at low chitin loadings (≤ 10 wt%), an indication of increased
properties of PHB/lignin blends in different compositions. The authors crystallization [71]. This improvement was linked to nucleating effect
reported 50% drop in thermal decomposition activation energy of PHB/ of chitin, which promoted rapid growth of PHB crystals. However, at
lignin blends compared to pure PHB, which suggested that presence of higher chitin loadings, crystallinity decreased and, this ascribed to re-
lignin reduced the thermal stability of the blends. With up to 40% lignin duced PHB chain mobility, due to intermolecular hydrogen bonds be-
content, the blends exhibited single phase morphology, single Tg (15 °C) tween PHB and chitin. Fig. 12 shows the proposed structural changes
and miscibility [68]. However, above 50% lignin content, phase se- and intermolecular hydrogen bonds interactions in PHB/Chitin blends
paration occurred, with two Tg (18 °C and 130 °C) of the components. [71].
On the other hand, viscoelastic properties were dependent on lignin
content. At low lignin content (≤ 30%), dynamic and loss modulus
were lower than that of pure PHB, whereas, higher loadings of lignin 3.1.5. Polyhydroxyalkanoates (PHA)
exhibited higher moduli. These viscoelastic response differences be- In an earlier work, Gassner & Owen [72] reported a blend of PHB
tween low lignin and high lignin content PHB blends were attributed to with another homopolymer, PHBV. The melt-pressed film formed phase
lower crystallinity and enhanced rigidity, respectively [68]. separated domains, nucleation, and growth of two kinds of spherulites
was observed, ascribed to the presence pure PHB and PHBV. DTA
measurements exhibited two separate melting endotherms, further
3.1.4. Chitin
confirming the presence of the two components [72]. Overall, the sto-
Chitin is the second most ubiquitous structural polysaccharide after
rage modulus decreased with % HV loading, while the Tg remain un-
cellulose, found in exoskeletons of crabs, lobster, beetles, shrimps, as
changed compared to neat PHB, but was generally stiffer than the
well as cell walls of fungi and yeasts. Fig. 11 show structures of (a)
comparative P(HB-co-18% HV) film due to the high crystallinity [72].
chitin and (b) chitosan. Chitin consists of both, crystalline and amor-
PHBV is a copolymer of PHB with incorporation of HV units in the PHB
phous domains; the crystalline domains can potentially serve as re-
backbone, the amount of HV% grafted determines the degree of flex-
inforcements. Furthermore, chitin biofiller has many desired properties,
ibility of the blend. In their review in 1993, Verhoogt et al. [73] re-
including, biocompatibility, biodegradability, nontoxicity and high
ported that blends of PHB with P (HB-co-8% HV) and PHB with P (HB-
modulus. Upon hydrolysis, chitin yields chitosan.
co-18.4% HV) were miscible, with one single sharp endotherm. How-
Research studies have shown that introduction of chitin/chitosan as
ever, at higher concentration of P (HB-co-76% HV) copolymer, im-
rigid filler into PHB matrix improved the mechanical properties.
miscibility and phase separation took place [73].
However, publications in this area of research are limited, probably due
On the other hand, Zhao et al. [74] blended PHB with PHB-HHx and
to (i) high cost of the common solvent for PHB/chitin/chitosan in case
found increased flexibility and elongation at break with increasing
of solvent casting and, (ii) susceptibility of chitin/chitosan to carboni-
PHB-HHx content in the blend. Addition of flexible PHB-HHx also in-
zation under high temperature in melt blending [69]. However, prior
creased the crystallization temperature indicating the decline in crys-
arts in this area are worth highlighting. For instance, Cao and collea-
tallization degree of PHB. The resulting scaffold with 60% PHB-HHx
gues prepared PHB/chitosan films and scaffolds using emulsion
exhibited good mechanical properties as compared to pure PHB and
blending technique, in which PHB in chloroform and chitosan in acetic
also showed strong provision in the growth and physiological function
acid were mixed, followed by casting/evaporation processes. Compared
of chondrocytes, with great potential for tissue engineering application
to neat chitosan films, 30% PHB/chitosan blend exhibited higher tensile
[74]. Kabe et al. [40] subsequently fabricated annealed melt-spun fibers
strength and elongation at break, by an increment of 38% and 63%,
of PHB with ultra-high-molecular-weight PHB (UHMW-PHB) at the
respectively. Furthermore, these properties coupled with the porous
composition of 5/95 and 10/90 w/w% by similar cold-drawing
structure of the PHB/chitosan films accentuate tissue engineering ap-
methods, as mentioned in Section 1. The annealed fiber showed high
plication possibilities for these composites [69]. In a particular study,
performance with tensile strength, Young's modulus, and elongation at
Ikejima et al. [70] investigated thermal properties and crystallization
break of a 5/95 blend film at 242 MPa, 1.5 GPa, and 88%, which is
behavior of PHB/α-chitin and PHB/chitosan. In this work, composites
close to poly(ethylene terephthalate) [40].
were prepared by solution casting in hexafluoroisopropyl alcohol
(HFIP) solvent. PHB, chitin and chitosan solutions were prepared

OH
Fig. 11. Chemical structure of (a) chitin and (b) chitosan.
OH OH OH OH OH
O O O O O O
OH O OH OH
O O O OH
HO NH OH NH OH NH NH2 NH2 NH2
n HO OH OH
n
CH3 CH3 CH3

(a) (b)

1102
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 12. Proposed carbonyl-amine hydrogen


bonds interactions in PHB/Chitin blends and the
resultant structures.
Reprinted with permission from ref. Copyright
(2015) Elsevier [71]

3.2. PHB blending with synthetic biodegradable polymers thermal stability. However, the tensile strength and elongation at break
was lower compared PLA/PHB (75/25), by 90% and 64%, respectively
3.2.1. Polylactic acid (PLA) [86].
PHB properties can also be tailored by blending with polyester de- Recently, stereocomplex PLA (Sc-PLA) has been employed to en-
rived from renewable feedstock. PLA is biodegradable polyester pro- hance the thermomechanical properties of PHB and vice versa. For
duced from annually renewable natural resources, such as corn or sugar instance, Chang and Woo [87] investigated the melt-crystallization ki-
cane [75–78]. Although PLA is inherently brittle (Tg ~50–60 °C), the netics of PHB, by incorporating 2–10 wt% Sc-PLA using solution casting
overall mechanical properties and environmental attributes have been technique. The authors reported that even with low content of Sc-PLA,
the key attraction in the polymer world [79–83]. For example, D'Amico the crystallization of PHB was enhanced with a morphology transition
and coworkers [84] processed PHB/PLA blends with a Tributyrin (spherulites to dendrites). Fig. 13 show Polarization optical microscopy
plasticizer (TBL), through melt mixing technique. In this study, DSC (POM) images of 50/50 PHB/Sc-PLA blends, crystallized at 130 °C.
thermal analysis revealed that all blends exhibited two Tg's, between From the images, spherulitic (a, b) to dendritic (c, d) morphology
− 23 °C and 19 °C, which clearly indicated of immiscibility. The me- transition is clearly depicted [87]. Additionally, DSC analysis revealed
chanical properties were significantly altered, due to the presence of that addition of 10% Sc-PLA shifted the crystallization peak tempera-
20% plasticizer and varied PLA content in the blends. Compared to neat ture (Tp) of PHB by > 20 °C from the melt. Higher Tp indicated faster
PHB, elongation at break for PHB/70% PLA improved ~50-fold and crystallization of the blend. Above 10% Sc-PLA, Tp decreased and this
consequential 58% decrease in tensile strength. The authors analyzed indicated an optimum Sc-PLA content for efficient crystallization under
the SEM images and concluded that fibrillation and debonding were the nonisothermal conditions [87].
major deformation failure modes for the blends [84]. In another work, Although mechanical properties of PHB/Sc-PLA blends were not
Dong et al. [85] studied the effects of in-situ crosslinks of PHB/PDLLA examined in this study, the presence of Sc-PLA would be expected to
blend by reactive blending using DCP. The authors analyzed a PHB/ markedly enhance the mechanical and thermal stability of the blends
PDLLA (70/30) blend as a function of DCP content. With increased DCP [87].
content, the double melting peaks slightly shifted to a lower tempera-
ture [85]. This behavior was attributed to the presence of imperfect and
thinner lamellar of PHB, due to the cross-linked and branched network. 3.2.2. Polycaprolactone (PCL)
On the other hand, the addition of 0.5% DCP enhanced the tensile PCL is an aliphatic biopolymer with a crystallinity of about 50%,
strength and impact toughness, by 14% and 30%, respectively. These glass transition at around −70 °C and melting temperature of about
enhancements were linked to improved interfacial adhesion, tailored by 60 °C. Due to its linear structure, it is generally used as a biodegradable
in-situ crosslinks/branching [85]. polymeric plasticizer in blends, thus improving the processability of the
PLA/PHB blends have also been studied by varying PHB content. In blend, where the increase in ductility and elongation at break is at-
such blends, it interesting to consider blend compositions with higher tributed to the plasticizing effect of the soft segment PCL, depending on
PHB content, in order to have the other component as the minor phase. the amount [88–95]. Therefore, it is crucial to optimize the PCL loading
For instance, Zhang and Thomas [86] prepared PLA/PHB blends with content in order to achieve overall toughening or strengthening and,
different weights ratios by melt compounding. By considering blends eventually the thermal stability of the blend. Generally, PHB and PCL
with high PHB contents, such as 100, 75 and 50, the effect of PLA as a are immiscible, with no synergetic effects between the two biopolymers
minor phase in the blend can be elucidated [86]. For instance, DSC due to immiscibility and the overall properties were dependent on
scans for PLA/PHB (50/50) and (25/75) showed the lowest melting blend composition.
temperature with a reduction of 4 °C, which indicated improved Vergara-Porras et al. [96] investigated the influence of thermal
treatment on mechanical, thermal properties and biodegradation

1103
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 13. POM images showing morphological transition


from spherulites to dendrites in PHB/Sc-PLA (1:1) blends.
Reprinted with permission from ref. Copyright (2011)
Elsevier [87].

kinetics of 70/30 (PHB/PCL) blends which were compression molded produced by copolymerization of propylene oxide and carbon dioxide,
after melting. The results showed that fast cooling rates (176 °C/min) mainly used toughening some epoxy resins. However, it has slightly low
hindered crystal formation, thus, reducing the elongation at break mechanical strength, poor thermal stability and generally immiscible
by > 90%. At a lower cooling rate of 1.3 °C/min, tensile strength in- with PHB. [99,100] Zhou et al. [101] studied the toughness as a
creased three-fold, whereas, Young's modulus slightly decreased. This function of temperature on toughening Poly(3-hydroxybutyrate)
could be correlated further with the thermal analysis by DSC where two (PHB)/poly(propylene carbonate) (PPC) blends containing various
peaks were observed in the PHB melting temperature region [96]. In amounts of propylene carbonate (PC) plasticizer, along with mor-
the case of fast or sudden cooling, the formation of nuclei was hindered phology studies of dispersed particles on the brittle-ductile transition
which in turn impeded crystallization. In contrast, slow cooling only for PHB/PPC/PC blends [101]. The result tough PHB/PPC/PC blends
showed a single melting peak indicating a more homogeneous crystal exhibited a reduction in the brittle-ductile transition temperature (TBD)
distribution with polymer chains. These chains were arranged and from 60 °C to 10 °C and a 750% increment in impact strength (5 to
packed properly, thus, creating larger crystals that were more stable, 34 kJ/m2), with plasticizer content ranging 0 to 25 wt%. Noticeably,
with a higher degree of perfection [96]. Garcia et al. [97] reported that the particle size of plasticized PPC became uniformly smaller, but this
PCL acts as an impact modifier which reduces the intrinsic fragility of accompanied by a huge reduction in Young's modulus from 1000 to
the PHB/PCL blend, thus providing higher impact resistance (impact 0.3 MPa [101]. Recently, El-Hadi [102] reported a novel PHB/PPC
energy of 10.6 J m− 2) and ductility (the elongation at break > blend using acetyl triethyl citrate/polyvinyl acetate (PVAc) as plasti-
1000%), however at the expense of both strength and modulus [97]. cizer and compatibilizing agent to improve the miscibility between the
The optimum was at 25% PCL where the impact energy and elongation PHB and PPC blend. The strong intermolecular interactions of the
at break increased slightly, 2 to 2.6 Jm− 2 and 8.1% to 11.2%, respec- electrospun fibers exhibited an increase in flexibility (elongation at
tively. Moreover, flexural strength and modulus were maintained, break ~335–475%), this was attributed to the formation of hydrogen
whereas, tensile strength and modulus decreased by 3.6% and 15%, bonds [102]. DSC results revealed that the melting temperature and
respectively [97]. The melt peak temperature for PHB and PCL showed crystallinity PHB was suppressed by increasing PVAc content, due to
no significant changes and the degradation temperatures of the blends hydrogen interactions revealed in FTIR as broadening of OH group and
revealed significant improvements compared to PHB [97]. Garcia et al. shifting of the ester band C]O [102]. The improved interactions were
[98] further studied the effects of 0–1 wt% of dicumyl peroxide (DCP) further observed in Fig. 15 showing the miscibility of PHB and PPC,
on the thermal, mechanical and surface morphology of the 75/25 with and without additives. In the POM images, it was observed that
(PHB/PCL) melted by reactive extrusion. The elongation at break and PHB and PPC phase separated with additives as shown in Fig. 15 (a, a′).
impact-absorbed energy of the blend with of 1 wt% DCP showed a re- Further increase in additives showed that PPC domains dispersed uni-
markable increase of 91% and 231%, respectively. However, the flex- formly and well embedded in the PHB matrix. This phenomenon was
ural properties and tensile modulus were sacrificed as shown in Fig. 14 ascribed to the fact that the additives aimed at the dispersion of PPC
[98]. FESEM and AFM characterization revealed that the interfacial into small domains in the PHB matrices, leading to strong interactions
adhesion and compatibility were improved between the two phases, between PHB/PPC. The developed fiber could be used in air and water
which remarkably decreased the anisotropic PCL particle sizes in the filters, absorbent pads in children diapers, wound dressing, surgical
PHB matrix with a noticeable decrease in the inter-particle distance. In suture and drug delivery applications [102].
addition, the DCP reduced the crystallinity of PHB by about 14%,
whereas, other properties remain unchanged [98].
3.2.4. Other biodegradable polymers
In addition to PLA, PCL and PPC, some other biodegradable poly-
3.2.3. Polypropylene carbonate (PPC) mers have also been utilized to blend with PHB to modulate its thermal
PPC is a relatively new biodegradable aliphatic biopolymer and mechanical properties [103–110]. For example, Poly (butylene

1104
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 14. Mechanical properties of the PHB/PCL


(75/25) blend as a function of DCP content used
for reactive extrusion: (a) tensile properties, (b)
flexural properties and (c) impact properties.
Copyright (2017) Elsevier [98].

succinate) (PBS), a biodegradable and biocompostable polyester which mechanical properties. Azuma et al. [109] investigated the thermal
can be produced from both petroleum and renewable sources, with a Tg behavior and miscibility of blends containing PHB and PVA. The au-
of −30 °C and Tm of about 112 °C. Due to its good mechanical prop- thors reported that the degree of miscibility were composition depen-
erties and processability, PBS has been used as a toughening agent in dent and the miscibility improved with increasing PVA content [109].
other polymer systems. For instance, Ma et al. [103] reported on Recently, Ol'khov et al. [110] reported that PVA/PHB extruded films
toughened PHB/PBS film via in-situ compatibilization using 0.5 wt% of with 0 to 50% of PHB showed potential in packaging, agricultural and
dicumyl peroxide as a free-radical grafting initiator. The initiator in- biomedical applications. At 50% PHB, elongation at break remained
creased the impact and tensile properties of all the blends. The PHB/ low, coupled with reduction in tensile modulus. However, with low
PBS/DCP (50/50/0.5) blend achieved a 9-fold (0.6 to 5.5 kJ/m2) and PHB content (≤30%) the tensile strength and elongation at break were
15-fold increase in impact toughness and increased elongation, re- enhanced, due to reinforcement effect of PHB fiber and uniform dis-
spectively. The enhancement was due to improved compatibility be- persion within the matrix, whereas, PVA contributed to the flexibility of
tween the partially cross-linked PBS particles and the PHB matrix, the blend [110].
which altered the plastic deformation through dilatation and fibrilla-
tion of the matrix [103].
In another aspect, polyethylene glycerol (PEG), also known as 3.3. Blending with non-biodegradable polymers
polyethylene oxide (PEO) or polyoxyethylene (POE) is an aliphatic
polyether that is significantly used in medical applications. Avella et al. 3.3.1. Elastomers
studied the miscibility of PHB and poly (ethylene oxide) (PEO) and Another strategy to obtain toughened polymer is by incorporation of
found that both the crystallizable components were miscible in the an elastomer as the second component. However, this is usually ac-
melt, characterized by single Tg, depressed equilibrium Tm and radial companied by a reduction in stiffness and tensile strength [111].
growth rate of PHB spherulites [104,105,106]. In another work, Shuai Therefore, it is of crucial importance to balance the ratio of the elas-
et al. [107] reported that addition of 5 and 10% of compatibilizing tomeric counterpart in order to achieve a toughened polymer with
agent (PCL-b-PEG) had slight improvement in elongation at break, but optimal thermal and mechanical performance. For example, Greco et al.
decrease in tensile strength. The PEG phase of the block copolymer was [112] reported on blending PHB with amorphous (ethylene-propylene
miscible with PHB, while PCL phase remain immiscible with PHB, rubber (EPR) and poly (vinyl acetate) (PVAc)). The authors reported
which resulted in the existence of isolated PCL domains free of PHB that PHB and EPR were immiscible in the melt, whereas PHB and PVAc
chains in the blend [107]. In a recent study, Jakic et al. [108] in- are compatible, with a single Tg and a drastic depression of equilibrium
vestigated the thermal stability and kinetic analysis of PHB/PEO blends Tm of PHB. The second component was found to have greatly influenced
by using non-isothermal thermogravimetry in an inert atmosphere. It the miscibility and phase structure in the melt and solid state [112]. In
was shown that PEO increased the thermal stability of PHB, with lower a recent study, Whitehouse [113] reported a novel toughening for-
degradation temperatures and occurs more slowly with the increase of mulation based on PHB, commercial PVAs and peroxide agent. The
PEO [108]. The findings agreed with studies using combined iso- optimum blended materials demonstrated higher toughness and ducti-
conversional Friedman method and multivariate non-linear regression; lity, where elongation at break increase from 23 to 257% and impact
where PEO displayed a higher value of activation energy confirming its strength increase from 0.4 to 0.8 ft lb/in with no expense in stiffness
superior thermal stability to PHB [108]. [113]. In another study, Parulekar et al. [114] developed a novel
Another polymer that has been employed in toughening PHB is the toughened PHB through reactive extrusion together with functionalized
flexible polyvinyl alcohol (PVA) with a Tm around 230 °C and good natural rubber (NR) and epoxidized natural rubber (EPR); along with a

1105
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 15. Optical micrographs of the PHB/PPC/PVAc/ATBC


blends with and without polarized (a, a1) 75/25/00/00, (b,
b1) 47/25/08/20, (c, c1) 45/25/10/20 and (d, d1) 43/25/
12/20.
Copyright (2017) Springer [102].

maleated polybutadiene as a compatibilizing agent to the PHB rubber Therefore, efforts to blend these thermoplastics with PHB have been
system. It was reported that maleated rubber with a high amount of made to increase the toughness and reduce the amount of thermoplastic
maleic anhydride (MA) grafting and low molecular weight was an ef- consumption in the materials. Although PP and PHB are immiscible as
fective compatibilizing agent. The well-dispersed rubber particles of the reported [117]. Pachekoski et al. [118] successfully prepared tough-
formulation (PHB + 30% ENR + 10% Maleated rubber) showed a ened PHB/PP injection molded samples with better mechanical prop-
tremendous enhancement in impact strength (440% increment) and a erties, improved miscibility and higher degradation in alkaline soil,
50% reduction in storage modulus [114–116]. compared to pure PHB. The PHB/PP blends showed proportionate low
crystallinity and stiffness of the polymer matrix, dependent on PP
content [118]. The PHB/PP (75/25) extruded samples showed an in-
3.3.2. Blending with polyolefin based thermoplastics crease in impact strength, elongation and Tm, by 31%, 1% and 4 °C,
Polypropylene and polyethylene are the two most commonly used respectively, compared to neat PHB. However, a slight decrease in the
petroleum-based polymers with similar physical properties to PHB.

1106
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Young's modulus and tensile strength of about 19.5% and 5%, respec- or glass fibers, due to environmental concerns related to safety and
tively, was reported. These results suggested a plasticization effect of recyclability. Incorporation of natural fibers offers many advantages,
the PP in the PHB, leading to a decrease in tensile strength [118]. such as, high specific stiffness and strength, desirable fiber aspect ratio,
Girioli et al. [119] reported on a series of PHB composites blended with low density, biodegradability and most importantly, lower the cost of
different polyolefin polymer including ABS, PA6, HDPE and PP, along the final product.
with other additives for toughening effect. A wide range of properties In an earlier work, Wong et al. [125] reported a toughened PHB-flax
were achieved with different blends, showing potential in cosmetic fiber composite by incorporating 4,4′-thiodiphenol (TDP), which acted
containers; cell phones; laptops; and packaging applications [119]. In as a hydrogen bonding additive between the OH group of the fibers and
subsequent studies, the research group reported as series of PHB/PP- the carboxyl group of PHB, to improve compatibility. The resulted
PHB copolymer blends together with some compatibilizing agents. composite exhibited higher thermal stability, but lower crystallinity
Results showed that the blends without compatibilizing agent had the and melting temperature as the TDP hinder the crystallization [125]. In
highest mechanical performance while the blend with compatibilizing addition, the Tg revealed an ascending trend with increased TDP con-
agent suffered a loss in the mechanical properties. The reduction de- tent. For example, at 5% TDP, the storage modulus showed an increase
monstrated that the PHB homopolymer and the PP homopolymer are of 135% and enhanced fiber–matrix adhesion [125]. Above 5% TDP, it
incompatible, even with the presence of a compatibilizer [120]. was suspected that the additive migrates into the matrix from the fibers
Another recent work on PHB/PP blend was reported by Sadi et al. and generate a ductile matrix, thus, reducing the storage modulus. SEM
[121] The authors prepared four different types of copolymers as and polarized optical microscopy studies revealed the formation of
compatibilizers: (i) poly(propylene-g-maleic anhydride) (PP–MAH), (ii) smaller crystals in the matrix and the reduced spacing between the fiber
poly (ethylene-co-methyl acrylate) [P(E–MA)], (iii) poly(ethylene-co- and the surrounding matrix as illustrated in Fig. 16 [125].
glycidyl methacrylate) [P(E–GMA)], and (iv) poly(ethylene-co-methyl Barkoula et al. [126] investigated the toughening effect of short flax
acrylate-co-glycidyl methacrylate) [P(E–MA–GMA)]. Out of the four fiber reinforced PHB composites with the effect of the fiber, copolymer
copolymers, [P(E–MA–GMA)] had the best performance, due to com- content and manufacturing process, and processing conditions. Indeed,
bined physical and reactive compatibilization provided by the MA the addition of flax fibers along with controlled processing conditions
groups and the GMA groups. Compared to PHB, the blend showed great toughened the native PHB due to fiber de-bonding and fiber pull-out
improvement in the elongation and impact strength, due to the small mechanisms as shown in Fig. 17. Interestingly, both Young's modulus
well dispersed spherical particles in the matrix, as revealed in the SEM and absorbed impact energy increased with increasing flax fiber con-
study [121]. tent. At 40 wt% of flax fiber, Young's modulus increased from 3 to
The miscibility of PHB/PE blends were evaluated by means of the 8 GPa, impact energy increased from 15 to 135 J/m, whereas, the
molecular dynamics (MD) simulation and the dissipative particle dy- tensile strength remained constant at 40 MPa and an insignificant drop
namics (DPD) simulation methods by Yang et al. [122] at both atomic in elongation at break was noted [126].
and mesoscale level, respectively. The simulation matches qualitatively Hodzic et al. [127] investigated the mechanical properties of both
with the experimental results where two glass transition temperatures PHB and PHBV bio-composite reinforced with bagasse fibers, which is a
are obtained [122]. The results show that PHB and PE are immiscible, by-product of the sugar making process. At the optimum fiber length,
where PE aggregate to form PE-rich domains showing two phases in flexural strength increased by 50%, surpassing the strength and mod-
PHB/PE blend [122]. Ol'khov et al. [123] reported the blend between 2 ulus than the standard thermoplastics [127]. These mechanical prop-
and 32 wt% of PHB as a secondary component in the PE matrix, the erties enhancements were ascribed to the coherent xylem structure of
results showed that the two blends were incompatible with clear phase the fibers, acting as reinforcement. In addition, both washing and
boundaries between the PHB dispersed phase and PE matrix [123]. The acetone treatment of the fibers improved the interfacial stress transfer
resultant band-like/cylinder-like fibrils morphology of the PHB en- between the fiber and the matrix. However, results varied slightly due
hanced the ultimate tensile strength and elastic modulus compared to to the surface treatment of the bagasse fibers [127]. Recent studies on
original PHB or LDPE polymer films [123]. This group further reported the toughening effect of agave fiber, which is a waste product of the
that blend with > 16 wt% PHB changed from an oriented/anisotropic Tequila industry reinforced in PHB green composites were carried out
PHB structure to an isotropic structure where the PHB fibrils form a by Torres-Tolle and co-workers [128]. Bio-composites with 30 wt% of
network in the blend; and this network displayed better resistance agave fiber had improved tensile modulus by about 80% (from 413 to
against hydrolysis, making it a potential inexpensive bio-erodible 770 MPa), flexural modulus by 36% (from 1581 to 2154 MPa), impact
packaging material [124]. strength by 44% (from 24.5 to 34.4 J/m) compared to neat PHB [128].
In general, blending is undoubtedly the most effective and eco- In addition, the authors observed a 47% increase in storage modulus
nomical approach for obtaining new materials with improved physical (from 2541 to 3733 MPa), as well as increase in degradation tempera-
and mechanical properties. However, glitches on phase separation and ture by about 10 °C. The synergistic toughening effect was attributed to
immiscibility between different polymers prove to be tricky. Inferior improved compatibility between the long agave fibers (with L/D = 10)
thermal and mechanical properties may be attributed to the immiscible and the partial polar characteristic of the PHB matrix interface [128].
blend, due to the absence of chemical bonding or synergetic effect be- Christian and Billington [129] studied two bio-based composites,
tween the separated constituents. PHB/hemp fabric, and CA/hemp fabric as alternatives to wood/en-
gineered wood-based materials for the construction industry. In this
4. PHB reinforced composites work, the authors prepared eight-layer composite laminates and eval-
uated the mechanical properties of the laminates. The composites ex-
Another toughening approach is by introducing a secondary re- hibited bilinear and tri-linear responses for warp direction and weft
inforcement such as natural fibers and nanofillers into the PHB polymer direction, respectively. These differences were ascribed to nonlinear
matrix. Besides the augmentation of both thermal and mechanical behavior of the polymeric matrices and yarn crimp effect. The PHB/
properties, unique properties attributed to the reinforcement phase, hemp composites were more ductile in the weft direction compared to
including, barrier, optical, flame retardant, conductivity, etc. can be CA/hemp, by about 29%. On the other hand, flexural strength for CA/
achievable. hemp was higher compared to PHB/hemp, ascribed to high stiffness in
compression. Overall, bio-based CA/hemp and PHB/hemp flexural and
4.1. Natural fibers shear strengths were comparable to wood/wood-based products, such
as furniture, crates, pallets, shelves, and formwork. Using combined
Natural fibers are mainly used as alternatives to traditional aramid classical laminate plate theory and non-linear numerical methods, a

1107
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 16. OM photographs of (a) PHB composite, (b) with


2.5% TDP, (c) 5% TDP, (d) 7.5% TDP, (e) 10% TDP.
Copyright (2004) Elsevier. [125]

Fig. 17. SEM pictures of impact fracture surface of flax/


PHB NMT composites, showing pull-out of flax fiber bun-
dles.
Copyright (2010) Elsevier [126].

good initial stress-strain response was obtained. However, the modified reduced, especially for PHB/lyocell composites. Inclusion of 30 wt% of
non-linear laminate plate theory overestimated the strength by about high modulus (10–40 GPa) hemp and jute fibers as fillers in PHB re-
0.09–0.23 times. This overestimate was linked to the use of in- sulted in large increases in flexural modulus of about 591% and 246%
dividualized stress-strain responses for each strain component [129]. In compared to neat PHB, respectively. On the other hand, PHB/lyocell
another interesting study, Gunning and coworkers [130] produced PHB composites exhibited the lowest flexural modulus, and modulus values
bio-composites based on three natural fibers, through extrusion-in- increased with fiber loading for each fiber type, but were also depen-
jecting molding techniques. PHB was reinforced with hemp, jute and dent on chemical composition [130].
lyocell, with a loading between 10 and 30 wt% and evaluated for re- Melo et al. [131] investigated the effect of various fiber treatment
inforcing effect on PHB. The authors revealed that the processing re- methods on the thermal and mechanical properties of carnauba fibers.
sulted in fiber attrition on all the fibers incorporated in the matrix. Firstly, carnauba fibers were extracted from palm tree and surface
Consequently, the tensile and impact properties were remarkably modified using caustic soda, peroxide, potassium permanganate and

1108
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

acetylation treatments. In this work, 10 wt% modified fibers were used thermal behavior, morphology and also kinetics of isothermal crystal-
in preparation of PHB-carnauba fibers prepregs, which were then hot- lization [137]. Both the MWCNTs showed improvement in storage
pressed into composite plates at 190 °C for 2 min and cut into tensile modulus, due to the better dispersion in the PHB matrix. MWCNTs
and flexural specimens [131]. From the TGA results, modified carnauba grafted with the long alkyl chains grafted resulted in steric hindrance,
fibers were found to be thermally stable up to 240 °C, a prerequisite for causing slower nucleating efficiency and overall crystallization kinetics
thermoplastic processing. The authors reported that PHB composites during the isothermal melt crystallization reducing the overall crys-
with hydrogen peroxide-treated fibers were about 10–40% superior tallinity of PHB [137].
compared to all the other composites, in terms of strength and elastic In another work, Botana et al. [138] fabrication PHB/clay nano-
modulus. This was ascribed to improved fiber-matrix interface adhesion composites with two commercially available clays, (i) Cloisite® Na
which facilitated stress transfer. Furthermore, temperature dependent + (Na-M) is a Na-montmorillonite and, (ii) Cloisite® 30B (30B–M) is an
storage modulus of the composites revealed that below Tg, the modulus organo-montmorillonite. The TEM and XRD results showed that the
was matrix controlled, whereas, at high temperatures the modulus was intercalation and exfoliation of 30B–M were more prominent than
fiber dominated [131]. In another study, Tǎnase et al. [132] studied the PHB/Na-M. In addition, incorporation of 30B–M increased the overall
physical mechanical behavior of PHB/cellulose fibers (CF) composites crystallization temperature, with a decrease in spherulites size and
with varying fiber content (2–10%) and 10% plasticizer. The authors smaller crystals formation, which contributed to the increase in Young
obtained PHB/CF blends by melt mixing technique at 175 °C/40 rpm modulus by about 12%. The results attested to the good compatibility
for 10 min. Films were then obtained through hot-press at 170 °C/5 min between the organic modified montmorillonite and PHB. However, the
at 150 bar. DSC results revealed that addition of CF had a negative tensile strength decreased about 9%, due to the low exfoliation/inter-
effect on melting temperature of all the composites, except PHB/CF10 calation ratio while the elongation at break was not reported [138].
composite. For instance, the Tm of PHB/CF5 composites was 155 °C, Panayotidou et al. [139] reported on the structural, thermal, me-
compared to 158 °C for neat PHB, from DSC second run [132]. This was chanical, and biocompatibility characteristics of PHB nanocomposites
ascribed to high CF content, which affected the flow properties (melt with phyllosilicate clays, specifically, the octadecylamine-modified
viscosity) of PHB due to reduction in chain entanglements in PHB. In montmorillonite (C18MMT at different loadings (1, 3, 5, and 10 wt%)).
addition, the crystallinity of all PHB/CF composites decreased com- The films were prepared by melt mixing in an extruder with nitrogen
pared to neat PHB, and this attributed to confined PHB chain mobility purging to prevent thermal degradation of PHB. DMA analysis revealed
by the cellulose fibers in the composite matrix. The PHB/CF composites that at an optimum of 5 wt% clay loading, the storage modulus and Tg
had high transparency and low transmittance, making them good increased by 85.6% and 5 °C, respectively, compared to neat PHB.
candidates for packaging material [132]. These enhancements were attributed to the excellent clay intercalation
In a more recent work, Scalioni and colleagues [133] investigated and exfoliation as revealed in AFM and TEM and X-ray diffraction
the morphology, thermal, mechanical and physical properties of PHB/ analyses. In addition, AFM showed that the surface of 5 wt% clay was
curaua fiber composites containing 30% plasticizer. The authors pre- rougher compared to that of the neat PHB, which supported osteoblast
pared both PHB/10% pristine curaua and PHB/10% modified curaua cells attachment and proliferation as shown in Fig. 18, a great potential
fiber composites by melt mixing technique. Incorporation of curaua in tissue engineering applications [139].
fibers into PHB resulted in morphological changes in PHB, i.e. spher- In a recent work, Akin and Tihminlioglu [140] investigated the
ulite-axialite transition, with profound effect on thermal and mechan- thermal, optical, and mechanical properties of the PHB based biona-
ical properties of PHB [133]. The authors reported that addition of nocomposite films fabricated by solvent casting method with the ad-
plasticizer shifted the crystallization temperature (Tc) downwards dition of with a commercial organo-modified montmorillonite (OMMT-
(37 °C), whereas, modified curaua fibers shifted the Tc upwards (72 °C). Cloisite 10A) at varied weight percent for potential packaging appli-
This was an indication that curaua fibers acted as nucleating agent in cation. The authors reported that the optimum improvements in me-
PHB, and initiated crystallization through heterogeneous nucleation chanical properties were obtained at 3 wt% clay loading, with an in-
mechanism. In addition, mechanical properties were affected by the crease of Young's modulus, tensile strength and strain at break. Clay
morphological changes [133]. For instance, strain at break (~ 14%) for exfoliation was achieved at low clay loaded sample of 1% wt. while the
PHB/modified curaua fibers was comparable to that of neat PHB clay was intercalated at higher loading of 3% wt. based on the X-ray
(~ 13%). However, impact strength for PHB/modified curaua fibers diffraction analysis. In addition, an increase in crystallinity at 1 wt%
increased by about 60%, only comparable to plasticized PHB. These showed that at low loading, clay act as a nucleating agent while ag-
curaua fiber reinforced PHB composites could find applications in au- gregation of clay restricting chain mobility results in lower melt en-
tomotive, appliances and food industries [133]. thalpy and crystallinity at high loading of 3 wt%. Therefore, it was
deduced that the resultant high-performance tough PHB/clay compo-
4.1.1. Inorganic nanofillers site depended on the dispersion of clay particles in the polymer matrix
Nanofillers such as carbon nanotubes, clay, and silica etc. were [140].
added to PHB polymer matrix for enhancement, including, barrier Barletta et al. [141] fabricated compression molded PHB films re-
properties, nucleation and crystallization properties, thermal and me- inforced with amino-functionalized silica (A-fnSiO2) and nano-sized
chanical properties [134]. In an earlier work, Yun et al. [135] reported graphene nanoplatelets (GNP) respectively. The overall rigidity and
on spray-dried PHB/SWCNTs blend from solution. Interestingly, the hardness of the filler modified film were higher than neat unmodified
topology of composite microspheres changed from spherical to the ro- PHB as observed from pencil hardness tested samples using FESEM, as
sette particles with increasing SWCNTs [135]. These composite pow- presented in Fig. 19 [141].
ders proved to be very good precursors for further processing, such as As seen, with unmodified neat PHB being the softest at B, graphene
extrusion or melt blending for bulk composite fabrication [135]. Yun modified PHB at H follow by amine-functionalized silica modified PHB
et al. [136] also reported the fabrication and mechanical properties of at HB which is hardest. Both the rigid nanofillers acted as a reinforce-
PHB/SWCNTs composite films for biomedical applications. Although ment to the polymer matrix, withstanding better the penetrating force
the films were more brittle than the neat PHB films, the hardness and acting on the surface [141]. Particularly, the amine functionalized silica
Young's modulus significantly increased with the addition of 1% gained additional cohesive strength with the hydrogen bonding inter-
SWCNTs, due to the nucleating effect of SWCNTs on PHB polymer action between the hydrogen on the amino group with the oxygen of
[136]. In a recent work, Huh et al. [137] successfully fabricated a PHB the carbonyl groups of PHB. In terms of scratch resistance, graphene
film with acid-treated MWCNTs and alkylated MWCNTs as reinforce- modified PHB performed better with a higher resistance to penetration
ment and investigated the effect of various MWCNTs loadings on their under scratch [141]. This was attributed to the incorporation of rigid

1109
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 18. (left) AFM 3D representations, (center) SEM imaging of human osteoblast cells after 7 days of culture and (right) staining by DAPI of human osteoblast cells after 7 days of culture
on (a) pure PHB and (b) PHB with 5 wt% organically modified MMT.
Copyright (2015) Wiley Periodicals, Inc. [139]

particles graphene particles, with high elastic modulus in the PHB size and excellent dispersion of the secondary reinforcement filler and
matrix increasing the overall rigidity and hardness of the composites most importantly, the strong filler-matrix interaction which proves to
during the scratch tests using progressive and constant load. In addi- be crucial in achieving a high-performance composite.
tion, both as-received and modified PHBs demonstrated good recovery
after the load release due to the intrinsic ductility of the PHBs in the
5. PHB toughening through chemical modification
engineered formulations [141]. However, toughness and elongation of
the composite was not reported. With improved mechanical strength
Apart from altering PHB using drawing and thermal treatment,
and hardness, PHB/graphene could potentially be used in tissue en-
blending or addition of reinforcements, chemical modifications are
gineering, medical devices and food packaging application [141]. In a
used to modify PHB at a molecular level. This approach is aimed at
recent study, Bian et al. [142] successfully fabricated a series of mod-
achieving improved properties by strategically controlling and de-
ified graphene oxide (MGO)/PHB nanocomposite and investigated their
signing chemical composition of the polymers and their architecture
thermal, mechanical and electrical performance. The surface MGO was
[143]. In this section, we will review on the approaches employed to
grafted with functional reagents, tolylene-2,4-diisocyanate (TDI) and
improve mechanical and thermal properties of PHB by chemical func-
1,4-butanediol (BD), to improve the interaction via hydrogen bonding
tionalization.
as illustrated in Fig. 20 [142].
PHB has a relatively narrow temperature processing window, and as
It was reported that the tensile modulus increased by 46% and
such, it is sensitive to thermal degradation. At temperatures near to its
tensile strength increased by 101% at 2 wt% MGO loading, accom-
Tm (160 to 180 °C), significantly decreases in the thermal stability occur
panied by a slight reduction in elongation at break. The enhancement
due to random chain scission of the molecular chains of PHB. Above
was attributed to the strong interaction between the PHB matrix and
180 °C, a tremendous decrease in molecular weight of PHB happens as a
the well dispersed and exfoliated MGO [142]. In addition, it was ob-
result of rapid random scission according to the β-elimination me-
served that the thermal degradation was enhanced by 15.5 °C as com-
chanism involving a six-membered ring transition state, as shown in
pared to the neat PHB and the crystallinity of PHB increased by in-
Fig. 21.
corporation of MGO. Furthermore, the entrenched graphene-graphene
In order to improve the toughness and thermal stability of PHB,
network contributed to the electrical conductivity that was 14-orders
different PHA copolymers were produced based on various bacterial
higher than neat PHB, at only 4 wt% MGO loading [142]. In principle,
strain and the cultivation conditions. The alternation in chemical
the key challenges of this approach are generally the optimum particle
structure of PHB allows to tailor specific properties of the deigned PHB

2B B HB F H HB
Fig. 19. FESEM image of residual deformation pattern after pencil hardness tests: (i) PHB at 2B; (ii) PHB at B; (iii) PHB/GNP at HB; (iv) PHB/GNP at F; (v) PHB/A-fNSiO2 at H; (vi) PHB/
A-fNSiO2 at HB. Reprinted with permission from ref.
Copyright (2016) Springer Science [141].

1110
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 20. Schematic of synthesis route of MGO/


PHB nanocomposites. GO forms a good disper-
sion and functional reagent TDI and BD could be
grafted onto the surface of GO to form MGO
through the “grafting to” approach [142].

copolymers, such as melting point and crystallinity depending on the HV counterpart showed an enhancement of elongation at break to
monomeric composition. Among all PHB copolymers, PHB-co-PHV 970%, but with a slight decrease in Young's modulus and tensile
(PHBV) under trade name of Biopol [6] and PHB-co-PHHx (PHBHHx) strength of the films [106].
have aroused great interest. PHBV is formed by the addition of valeriote Another popular PHB-based copolymer incorporating a long ali-
(HV) side group to the main backbone and it was found that this co- phatic branch is called PHB-co-HHx [147]. Doi et al. [148] reported
polymer exhibits the phenomenon of isodimorphism as the % HV that the crystallinity of this copolymer is greatly reduced after the
content of the side group varies [144]. Incorporation of 40 mol% of HV grafting of HHx units, which eliminated the secondary crystallization
units resulted in a decrease of the Tm, from 178 °C to about 75 °C, which effect. The elongation at break of the films dramatically increased from
greatly broadened the melt processing temperature window. At 95 mol 6 to 850% at 17 mol% HHx fraction, while the tensile strength de-
%, the Tm increase to 108 °C, as reported by Konioka and his co-workers creased from 43 to 20 MPa. This is accompanied by visual inspection of
[145,146] In terms of mechanical properties, PHBV with 20 mol% HV the resulted film, which became soft and flexible [148]. The authors
content behaved like PHB homopolymer, where crystallinity increased [149] further reported on uniaxial cold-drawing and annealing of PHB-
with time, due to secondary crystallization. At 34 mol%, the dominant co-HHx films with 5 and 12 mol% HHx and evaluated the mechanical

Fig. 21. Expected thermal degradation pathways of PHB by


random chain scission based on ester decomposition me-
chanism (β-elimination) involving a six-membered ring
transition state.
Copyright (2008) Elsevier Polymer Degradation and
Stability [26].

1111
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 22. Synthetic Route of Poly(PHB/PEG urethane)s from High Molecular Weight PHB and PEG.
Copyright (2010) American Chemical Society Journal of Physical Chemistry B [153].

properties, structure and biodegradability in comparison with PHB mineralized by simple incubation in SBF [152].
homopolymer. The resulted films possessed high tensile strength and Naguib et al. [153] fabricated poly (ester-urethane) block copoly-
Young's modulus, and showed hardly any variation after storing for mers consisting PHB-diol as the hard segments and PCL-PEG-PCL blocks
60 days at room temperature. Mechanical properties of PHB-co-5%- as the soft segment for toughening effect, using 1,6-hexamethylene
HHx films were even further improved by two-step orientation at room diisocyanate (HDI) as coupling agent as shown in Fig. 22. The tensile
temperature, and these films retained a high performance. The en- properties of the cast film showed that an increased PCL-PEG-PCL soft
hancement was due to the increase in the orientation of the crystal segment content led to higher elongation at break with a concomitant
domains formed by the 21 helix conformation (α-form) avoiding sec- decrease in tensile strength. The toughening behavior was explained in
ondary crystallization [149]. In an earlier work, Hong and Lin [150] relation to DSC analysis; as the tendency of phase separation increased
studied the thermal degradation behaviors and crystallization of PHB with increasing content of soft segment in the PHB/PCL-PEG-PCL co-
grafted with exo-3,6-epoxy-1,2,3,6-tetrahydrophthalic anhydride (ETA) polymer. Additionally, the degree of phase mixing correlated with in-
and maleic acid (MAc). The results showed that the grafted PHB has a creasing Tg, as PHB content increased. This led to the stiffening of the
larger crystallinity, faster crystallization rate, and better thermal sta- soft segments, and eventual reduction in tensile strength and elongation
bility compared to the native PHB. The bulky maleic anhydride mole- at break [153]. Adamus [154] reported a novel di-block copolymers
cules hindered and inhibited the development of six-member cyclic containing PHB with atactic-PHB (a-PHB) as structural segments for
rings, which led to the thermal enhancement of PHB [150]. material for cardiovascular engineering. The copolymer was also con-
Apart from the formation of PHB copolymers by bacterial fermen- firmed by proving the reduction in vascular prostheses permeability.
tation, development of toughened PHB materials through chemical re- The DSC results showed that only one Tg was observed for PHB-block-(a-
actions has also been well explored. For example, Li et al. [151] suc- PHB) (Tg = −2.2 °C) indicating full mixing and compatibility between
cessfully synthesized a series of poly (PHB/PEG urethanes) by varying the two blocks. Furthermore, the data showed alterations in the crys-
the amount of PEG and PHB segments using HDI coupling agent. It was tallization process when a-PHB is connected chemically or physically
found that the thermal stability of the urethane film performed better with natural PHAs (PHB, PHBV) [154]. a-PHB act as the soft segment in
than the parent polymers, with reduced crystallinity [151]. The flexible the block copolymers which decrease the order and crystallinity of the
PEG toughened the overall polymer; thus, increased the strain at break compatible blends, hindering the PHB crystallization. Hence, yielding
from 2% to 227–1912% simply by tuning by the composition with more lower ΔHm values and crystallization as compared to the initial PHB
PEG content. However, similar drop was observed in the modulus and macroinitiators [154]. On the other hand, block copolymer consist of
strength of the polyurethane [151]. In another similar research, Liu atatic-PHB allows PHB to crystallize as partially separated component,
et al. [152] developed a hydrophobic polyester scaffold based on a due to the strengthened chemical bond between two blocks [154]. Wu
series polyurethane block copolymers consisting of PHB as the hard et al. [155] successfully synthesized a novel tri-block copolymer, PHB-
segment and poly(ethylene glycol) (PEG) as the soft segment without PLA-PCL by a two steps ring opening polymerization using methyl-PHB
any surface modification, such as, alkaline or plasma treatment, which (LMPHB) oligomer as starting material. The tri-block copolymer ex-
reduced the mechanical properties of the scaffold. In addition, PEG hibited high flexibility and biocompatibility which can be used in vivo
content in the polymers was kept below 50 wt% to avoid much loss of biomedical application [155]. The authors reported that the crystal-
stiffness and strength and maintain the dimensional stability of the lization of PHB was reduced by decreasing the length of LMPHB oli-
electrospun fibrous polymer scaffold immersed in simulated body fluid gomer segment, as the PLA blocks confines the chain mobility of PHB
(SBF) [152]. Mechanical test results showed that the block copolymers blocks, resulting in higher Tg [155].
with shorter PHB segments or higher PEG content in the block copo- In a more recent work, Aluthge et al. [156] investigated the me-
lymerization were more ductile, lower PHB crystallinity and no PEG chanical and rheological properties of PLLA-PHB-PLLA and PLLA-PHB-
crystallinity. Moreover, their ductility was enhanced in hydrated states, PDLA triblock copolymers synthesized by sequential addition using
which translated to enhanced strain at break from 1090 to 1962%. The dinuclear indium [(NNO)InCl]2(μ-OEt)(μ-Cl), as catalyst which is very
electrospun fibrous scaffold consisting of the block copolymers showed active for the living ring-opening polymerization of β-butyrolactone
great potential for bone regeneration application, since it can be easily (BBL) and cyclic esters lactide (LA). A series of triblock copolymers

1112
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

Fig. 23. Polymerization of Cyclic Esters by [(NNO)InCl]


2(μ-Cl)(μ-OEt).
Copyright (2013) American Chemical Society
Macromolecules [156].

were generated with different weight percentages of PHB as the middle can be fabricated easily into structural shapes, including hollow sec-
segment as illustrated in Figure 23 [156]. The thermal stability of the tions. Furthermore, two modeling approaches, (i) combined classical
PLLA-BBL-PLLA copolymer were more stable than PLLA-BBL-PDLA, but laminate plate theory and (ii) non-linear numerical methods were used
the thermal stability dropped when the amount of PHB middle segment to obtain a good initial stress-strain response [129]. Another potential
was increased. It was found that triblocks incorporating PHB reduce the application with the enhanced properties was applied in bone scaffold
tensile properties compared to those control samples with only PLA tissue engineering. While PHB offers the fundamental requirement of
blocks (i.e. PLLA-PDLA-PLLA and PLLA-PLLA-PDLA), but the elonga- biodegradability and biocompatibility, strength and cell proliferation
tion at break improved significantly due to the addition of elastomeric are equally vital for a scaffold material. An effective scaffold should
atactic PHB. Clearly, the copolymers with a shorter block of PLLA and provide adequate physical support similar to actual bone to induce
PDLA and a longer block of PHB resulted in a reduction in the tensile bone tissue regeneration while providing incessant supply of nutrients
strength of the triblock [156]. Of all the strategies mentioned, chemical and metabolites to the tissues developed on the scaffold. The in-
modification is by far the most complex and strategic approach to corporation of well exfoliation organo-modified montmorillonite clay
achieving a toughened PHB. This approach involved the tactical design into the PHB matrix enhanced both the storage modulus and Tg as re-
of chemical composition of polymers at the molecular level. Chemically vealed in the TEM and X-ray diffraction analyses. To understand the
modifying the polymer surface with different functionality, yielding a effect of temperature on the mechanical properties of the scaffold, the
high-performance PHB with strong chemical bonds. Yet, concerns over storage modulus were investigated at two temperatures, 20 °C which
the high cost and intensely usage of chemicals and solvent during the corresponds to the installation temperature and 37 °C which corre-
synthesis and purification process, and the duplicability from lab scale sponds to the human physiological temperature. It was found that at
to industrial scale remains a huge hurdle for the industrial adoption. 20 °C, storage modulus increased by 85.6% and of 39.8% for 3 and 5 wt
% clay loading, respectively. Meanwhile, the slight increment was ob-
served at 37 °C: with 47% and 25% increase for 3 and 5 wt% clay
6. Applications
loading, respectively. In addition, AFM was conducted to evaluate the
surface roughness and the result shows that the surface of 5 wt% clay
Most research and development on toughened PHB biopolymers
was rougher, as compared to that of the neat PHB. The irregular to-
targets applications such as packaging, automobile and biomedical.
pography supported osteoblast cells attachment and proliferation just
With enhanced thermal and mechanical features, these toughened PHB
3 days after the cell culture. The cells efficiently attached and spread
materials substantially open new perspectives in construction and
over a large area on the PHB/clay nanocomposite surfaces through fi-
scaffold tissue engineering applications (Table 1). Development of fully
lopodia or lamellipodia. To further investigate the proliferation rates,
bio-based natural fiber composites, also known as green composites,
staining by DAPI was carried out on the human osteoblast cells after
potentially reduce the final material cost while enhancing both the
7 days of incubation time as shown in Fig. 18 and the results was po-
thermal and mechanical performance, For instance, the use of hemp
sitive. In conclusion, the incorporation of the organo-modified mon-
fabric PHB composite as alternatives to wood or engineered wood-
tmorillonite was able to improve the thermomechanical properties of
based materials. The authors prepared an eight-layer composite lami-
the bone scaffold without affecting the adhesion and the distribution of
nates using lay-up method in three different orientation ([0/ ± 45/0],
the cells in comparison to those on pure PHB [139].
[30/−30/60/−60/0] and [0/45/−45/90]) and evaluated the me-
chanical properties of the laminates. The composites exhibited bilinear
and tri-linear responses for warp direction and weft direction, respec- 7. Conclusion and future perspectives
tively [129].
Generally, the flexural and shear strengths of the bio-based com- The development of a biodegradable toughened PHB composite
posites are comparable to the allowable wood design strengths and it with superior mechanical properties could overcome the shortcomings

1113
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

and open up new possibilities for industrial application. Hereby, we [16] C. Isola, H.L. Sieverding, R. Raghunathan, M.P. Sibi, D.C. Webster, J. Sivaguru,
reviewed a series of approaches by researcher in achieving this goal, J.J. Stone, Life cycle assessment of photodegradable polymeric material derived
from renewable bioresources, J. Clean. Prod. 142 (2017) 2935–2944.
which includes modification through drawing and thermal treatment, [17] I. Manavitehrani, et al., Biomedical applications of biodegradable polyesters,
blending with materials from natural sources, synthetic biodegradable Polymers (Basel). 8 (2016).
and non-biodegradable polymers, inclusion of natural fibers or rigid [18] Y.L. Wu, et al., PHB-based gels as delivery agents of chemotherapeutics for the
effective shrinkage of tumors, Adv. Healthc. Mater. 5 (2016) 2679–2685.
fillers to form reinforced composites and lastly, modification by che- [19] Z. Li, X.J. Loh, Recent advances of using polyhydroxyalkanoate-based nanove-
mical functionalization discussing the structure-property studies on the hicles as therapeutic delivery carriers, Wiley Interdiscip. Rev. Nanomedicine
thermal and mechanical properties of the PHB based blends and com- NanoBiotechnology 9 (2017) 19–22.
[20] M. Meischel, J. Eichler, E. Martinelli, U. Karr, J. Weigel, G. Schmoller,
posites. Collective efforts of scientist, biologist and engineers working E.K. Tschegg, S. Fischerauer, A.M. Weinberg, S.E. Stanzl-Tschegg, Adhesive
on PHB are essential rise the popularity and adoption of PHB into in- strength of bone-implant interfaces and in-vivo degradation of PHB composites for
dustrial. The key areas include [1] development of a cost-effective load-bearing applications, J. Mech. Behav. Biomed. Mater. 53 (2016) 104–118.
[21] Z. Li, J. Yang, X.J. Loh, Polyhydroxyalkanoates: opening doors for a sustainable
fermentation process that lower the production cost, making it more
future, NPG Asia Mater. 8 (2016) 1–20.
favorable over traditional petroleum plastics; [2] design of a recycl- [22] Z. Li, X.J. Loh, Water soluble polyhydroxyalkanoates: future materials for ther-
ability and reusability process for all commercialized PHB products apeutic applications, Chem. Soc. Rev. 44 (2015) 2865–2879.
prior to production; [3] creating a comprehensive life assessment of [23] G.-Q.A. Chen, Microbial polyhydroxyalkanoates (PHA) based bio- and materials
industry, Chem. Soc. Rev. 38 (2009) 2434–2446.
these PHB; and [4] enhancement of the overall mechanical and thermal [24] J.K. Hobbs, P.J. Barham, Fracture of poly(hydroxybutyrate). Part III. Fracture
performance of PHB. morphology in thin films and bulk systems, J. Mater. Sci. 34 (1999) 4831–4844.
While biodegradability and biocompatibility of PHB remain ad- [25] N. Grassie, E.J. Murray, P.A. Holmes, The thermal degradation of poly(-(d)-B-
hydroxybutyric acid): part 3-the reaction mechanism, Polym. Degrad. Stab. 6
vantageous, strategies to tailor material architecture for toughening (1984) 127–134.
PHB are vital to tackle its shortcoming of embrittlement and thermal [26] H. Ariffin, H. Nishida, Y. Shirai, M.A. Hassan, Determination of multiple thermal
instability. Despite the strategies of toughening PHB as reviewed in this degradation mechanisms of poly(3-hydroxybutyrate), Polym. Degrad. Stab. 93
(2008) 1433–1439.
article, development of new approaches and processes could be ex- [27] Y. Wang, J. Yin, G.Q. Chen, Polyhydroxyalkanoates, challenges and opportunities,
plored and investigated. With an ultimate goal of achieving a fully Curr. Opin. Biotechnol. 30 (2014) 59–65.
biodegradable tough PHB composite with high strength and high [28] Y. Ke, X.Y. Zhang, S. Ramakrishna, L.M. He, G. Wu, Synthetic routes to degradable
copolymers deriving from the biosynthesized polyhydroxyalkanoates: a mini re-
thermal stability, PHB will become a more competitive candidate to view, Express Polym Lett 10 (2016) 36–53.
replace petroleum-based polymers for wide applications such as auto- [29] G.-Q. Chen, I. Hajnal, The ‘PHAome, Trends Biotechnol. 33 (2015) 559–564.
motive, construction, consumer products and even biodegradable toys, [30] J.K. Muiruri, S. Liu, W.S. Teo, J. Kong, C. He, Highly biodegradable and tough
polylactic acid-cellulose nanocrystal composite, ACS Sustain. Chem. Eng. 5 (2017)
in addition to the current biomedical application in tissue engineering
3929–3937.
and high-end packaging material. It is anticipated that PHB will become [31] Y. Sun, L. Yang, X. Lu, C. He, Biodegradable and renewable poly(lactide)–lignin
the most sustainable biopolymer for our future generations. composites: synthesis, interface and toughening mechanism, J. Mater. Chem. A
Mater. Energy Sustain. 3 (2015) 3699–3709.
[32] Y. Sun, C. He, Biodegradable ‘core-shell’ rubber nanoparticles and their tough-
References ening of poly(lactides), Macromolecules 46 (2013) 9625–9633.
[33] W. Thitsartarn, et al., Simultaneous enhancement of strength and toughness of
[1] R. Geyer, J.R. Jambeck, K.L. Law, Production, use, and fate of all plastics ever epoxy using POSS-rubber core-shell nanoparticles, Compos. Sci. Technol. 118
made, Sci. Adv. 3 (2017) e1700782. (2015).
[2] M. Rabnawaz, I. Wyman, R. Auras, S.A. Cheng, Roadmap towards green packa- [34] J. Sun, J. Wang, J.C.C. Yeo, D. Yuan, H. Li, L.P. Stubbs, C. He, Lignin epoxy
ging: current status and future outlook for polyesters in the packaging industry, composites: preparation, morphology, and mechanical properties, Macromol.
Green Chem. 18 (2017) 1–3. Mater. Eng. 301 (2016).
[3] D. Xanthos, T.R. Walker, International policies to reduce plastic marine pollution [35] M.L. Di Lorenzo, M.C. Righetti, Evolution of crystal and amorphous fractions of
from single-use plastics (plastic bags and microbeads): a review, Mar. Pollut. Bull. poly[(R)-3-hydroxybutyrate] upon storage, J. Therm. Anal. Calorim. 112 (2013)
118 (2017) 17–26. 1439–1446.
[4] R.W. Lenz, R.H. Marchessault, Bacterial polyesters: biosynthesis, biodegradable [36] G.J.M. de Koning, A.H.C. Scheeren, P.J. Lemstra, M. Peeters, H. Reynaers,
plastics and biotechnology, Biomacromolecules 6 (2005) 1–8. Crystallization phenomena in bacterial poly[(R)-3-hydroxybutyrate]: 3.
[5] G.Q. Chen, I. Hajnal, H. Wu, L. Lv, J. Ye, Engineering biosynthesis mechanisms for Toughening via texture changes, Polymer (Guildf). 35 (1994) 4598–4605.
diversifying polyhydroxyalkanoates, Trends Biotechnol. 33 (2015) 565–574. [37] R. Crétois, et al., Physical explanations about the improvement of poly-
[6] P.A. Holmes, Applications of PHB — a microbially produced biodegradable ther- hydroxybutyrate ductility: hidden effect of plasticizer on physical ageing, Polym.
moplastic, Phys. Technol. 16 (1985) 32–36. (United Kingdom) 102 (2016) 176–182.
[7] E. Bugnicourt, P. Cinelli, A. Lazzeri, V. Alvarez, Polyhydroxyalkanoate (PHA): [38] T. Iwata, M. Fujita, Y. Aoyagi, Y. Doi, T. Fujisawa, Time-resolved X-ray diffraction
review of synthesis, characteristics, processing and potential applications in study on poly[(R)-3-hydroxybutyrate] films during two-step-drawing: generation
packaging, Express Polym Lett 8 (2014) 791–808. mechanism of planar zigzag structure, Biomacromolecules (2005) 1803–1809.
[8] R. Pantani, L.S. Turng, Manufacturing of advanced biodegradable polymeric [39] T. Iwata, Y. Aoyagi, M. Fujita, H. Yamane, Y. Doi, Y. Suzuki, A. Takeuchi,
components, J. Appl. Polym. Sci. 132 (2015). K. Uesugi, Processing of a strong biodegradable poly[(R)-3-hydroxybutyrate] fiber
[9] H. Cheng, Z. Wu, C. Wu, X. Wang, S.S. Liow, Z. Li, Y.L. Wu, Materials Science & and a new fiber structure revealed by micro-beam X-ray diffraction with syn-
Engineering C Overcoming STC2 mediated drug resistance through drug and gene chrotron radiation, Macromol. Rapid Commun. 25 (2004) 1100–1104.
co — delivery by PHB-PDMAEMA cationic polyester in liver cancer cells, Mater. [40] T. Kabe, T. Tsuge, K.I. Kasuya, A. Takemura, T. Hikima, M. Takata, T. Iwata,
Sci. Eng. C (2017), http://dx.doi.org/10.1016/j.msec.2017.08.075 (in press). Physical and structural effects of adding ultrahigh-molecular-weight poly[(R)-3-
[10] Q. Liu, S. Cheng, Z. Li, K. Xu, G. Chen, Characterization, biodegradability and hydroxybutyrate] to wild-type poly[(R)-3-hydroxybutyrate], Macromolecules 45
blood compatibility of poly [(R)-3-hydroxybutyrate] based poly (ester-urethane) s, (2012) 1858–1865.
J. Biomed. Mater. Res. Part A 90A (2008) 1162–1176. [41] T. Kabe, C. Hongo, T. Tanaka, T. Hikima, M. Takata, T. Iwata, High tensile
[11] Z. Chen, S. Cheng, Z. Li, K. Xu, G. Chen, Synthesis, characterization and cell strength fiber of poly[(R)-3-hydroxybutyrate-co-(R)-3-hydroxyhexanoate] pro-
compatibility of novel poly (ester urethane) s based on poly (3-hydroxybutyrate- cessed by two-step drawing with intermediate annealing, J. Appl. Polym. Sci. 132
co-4-hydroxybutyrate) and prepared by melting polymerization, J. Biomater. Sci. (2015) 1–8.
Polym. Ed. 20 (2009) 1451–1471. [42] H.K. Lee, J. Ismail, H.W. Kammer, M.A. Bakar, Melt reaction in blends of poly(3-
[12] Z. Li, et al., Novel amphiphilic poly (ester-urethane) s based on poly [(R)-3-hy- hydroxybutyrate) (PHB) and epoxidized natural rubber (ENR-50), J. Appl. Polym.
droxyalkanoate]: synthesis, biocompatibility and aggregation in aqueous solution, Sci. 95 (2005) 113–129.
Polym. Int. 57 (2008) 887–894. [43] R.S. Kurusu, N.R. Demarquette, C. Gauthier, J.M. Chenal, Effect of ageing and
[13] J. Lim, M. You, J. Li, Z. Li, Emerging bone tissue engineering via annealing on the mechanical behaviour and biodegradability of a poly(3-hydro-
Polyhydroxyalkanoate (PHA)-based scaffolds, Mater. Sci. Eng. C 79 (2017) xybutyrate) and poly(ethylene-co-methyl-acrylate-co-glycidyl-methacrylate)blend,
917–929. Polym. Int. 63 (2014) 1085–1093.
[14] C.Y. Wee, S.S. Liow, Z. Li, Y. Wu, X.J. Loh, New poly [(R)-3-hydroxybutyrate-co-4- [44] R.S. Kurusu, C.A. Siliki, E. David, N.R. Demarquette, C. Gauthier, J.M. Chenal,
hydroxybutyrate] (P3HB4HB)-based thermogels, Macromol. Chem. Phys. 218 Incorporation of plasticizers in sugarcane-based poly(3-hydroxybutyrate)(PHB):
(2017) 1–13. changes in microstructure and properties through ageing and annealing, Ind. Crop.
[15] X. Wang, S.S. Liow, Q. Wu, C. Li, C. Owh, Z. Li, Codelivery for paclitaxel and Bcl-2 Prod. 72 (2015) 166–174.
conversion gene by PHB-PDMAEMA Amphiphilic cationic copolymer for effective [45] R. Hufenus, F.A. Reifler, M.P. Fernández-Ronco, M. Heuberger, Molecular or-
drug resistant cancer therapy, Macromol. Biosci. 1700186 (2017) 1–11. ientation in melt-spun poly(3-hydroxybutyrate) fibers: effect of additives, drawing
and stress-annealing, Eur. Polym. J. 71 (2015) 12–26.

1114
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

[46] S. Godbole, S. Gote, M. Latkar, T. Chakrabarti, Preparation and characterization of [76] Z. Li, P.L. Chee, C. Owh, R. Lakshminarayanan, X.J.R.S.C. Loh, Advances safe and
biodegradable poly-3-hydroxybutyrate–starch blend films, Bioresour. Technol. 86 efficient membrane permeabilizing polymers based on PLLA for antibacterial, RSC
(2003) 33–37. Adv. 6 (2016) 28947–28955.
[47] M. Zhang, N.L. Thomas, Preparation and properties of polyhydroxybutyrate [77] Z. Li, D. Yuan, X. Fan, H. Tan, C. He, Poly (ethylene glycol) conjugated poly
blended with different types of starch, J. Appl. Polym. Sci. 116 (2010) 688–694. (lactide)-based polyelectrolytes: synthesis and formation of stable self-assemblies
[48] I. Koller, A.J. Owen, Starch-filled PHB and PHB/HV copolymer, Polym. Int. 39 induced by stereocomplexation, Langmuir 31 (2015) 2321–2333.
(1996) 175–181. [78] X. Fan, Z. Wang, D. Yuan, Y. Sun, C. He, Novel linear-dendritic-like amphiphilic
[49] R.M.S.M. Thiré, T.A.A. Ribeiro, C.T. Andrade, Effect of starch addition on com- copolymers: synthesis and self-assembly characteristics, Polym. Chem. 5 (2014)
pression-molded poly (3-hydroxybutyrate)/starch blends, J. Appl. Polym. Sci. 100 4069–4075.
(2006) 4338–4347. [79] X. Fan, H. Tan, Z. Li, X.J. Loh, Control of PLA stereoisomers-based polyurethane
[50] L.H. Innocentini-Mei, J.R. Bartoli, R.C. Baltieri, Mechanical and thermal properties elastomers as highly efficient shape memory materials, ACS Sustain. Chem. Eng. 5
of poly (3-hydroxybutyrate) blends with starch and starch derivatives, (2017) 1217–1227.
Macromolecular Symposia, 197 Wiley Online Library, 2003, pp. 77–88. [80] X. Fan, S. Jiang, Z. Li, X. Jun, Conjugation of poly (ethylene glycol) to poly (lac-
[51] S.M. Lai, T.M. Don, Y.C. Huang, Preparation and properties of biodegradable tide)-based polyelectrolytes: an effective method to modulate cytotoxicity in gene
thermoplastic starch/poly(hydroxy butyrate) blends, J. Appl. Polym. Sci. 100 delivery, Mater. Sci. Eng. C 73 (2017) 275–284.
(2006) 2371–2379. [81] X. Fan, M. Cao, X. Zhang, Z. Li, Synthesis of star-like hybrid POSS- (PDMAEMA-b-
[52] P. Ma, P. Xu, M. Chen, W. Dong, X. Cai, Pauline Schmit, A.B. Spoelstra, PDLA) 8 copolymer and its stereocomplex properties with PLLA, Mater. Sci. Eng. C
P.J. Lemstra, Structure–property relationships of reactively compatibilized PHB/ 76 (2017) 211–216.
EVA/starch blends, Carbohydr. Polym. 108 (2014) 299–306. [82] B.H. Tan, J.K. Muiruri, Z. Li, C. He, Recent progress in using stereocomplexation
[53] H.-T. Liao, C.-S. Wu, Performance of an acrylic-acid-grafted poly (3-hydro- for enhancement of thermal and mechanical property of polylactide, ACS Sustain.
xybutyric acid)/starch bio-blend: characterization and physical properties, Des. Chem. Eng. 4 (2016) 5370–5391.
monomers Polym. 10 (2007) 1–18. [83] Z. Li, D. Yuan, G. Jin, H. Tan, C. He, Facile layer-by-layer self-assembly toward
[54] T. Don, C. Chung, S. Lai, H. Chiu, Preparation and properties of blends from poly enantiomeric poly (lactide) stereocomplex coated magnetite nanocarrier for
(3-hydroxybutyrate) with poly (vinyl acetate)-modified starch, Polym. Eng. Sci. 50 highly tunable drug deliveries, ACS Appl. Mater. Interfaces 8 (2016) 1842–1853.
(2010) 709–718. [84] D.A. D'Amico, M.L.I. Montes, L.B. Manfredi, V.P. Cyras, Fully bio-based and bio-
[55] I.T. Seoane, E. Fortunati, D. Puglia, V.P. Cyras, L.B. Manfredi, Development and degradable polylactic acid/poly (3-hydroxybutirate) blends: use of a common
characterization of bionanocomposites based on poly (3-hydroxybutyrate) and plasticizer as performance improvement strategy, Polym. Test. 49 (2016) 22–28.
cellulose nanocrystals for packaging applications, Polym. Int. 65 (2016) [85] W. Dong, P. Ma, S. Wang, M. Chena, X. Cai, Y. Zhang, Effect of partial crosslinking
1046–1053. on morphology and properties of the poly (β-hydroxybutyrate)/poly (D,L-lactic
[56] S. de O Patrício, P. Pereira, F.V. dos Santos, M.C. de Souza, P.P. Roa, P.B.O. Juan, acid) blends, Polym. Degrad. Stab. 98 (2013) 1549–1555.
L. Rodrigo, Increasing the elongation at break of polyhydroxybutyrate biopolymer: [86] M. Zhang, N.L. Thomas, Blending polylactic acid with polyhydroxybutyrate: the
effect of cellulose nanowhiskers on mechanical and thermal properties, J. Appl. effect on thermal, mechanical, and biodegradation properties, Adv. Polym.
Polym. Sci. 127 (2013) 3613–3621. Technol. 30 (2011) 67–79.
[57] L. Wei, A.G. McDonald, N.M. Stark, Grafting of bacterial polyhydroxybutyrate [87] L. Chang, E.M. Woo, Effects of molten poly (3-hydroxybutyrate) on crystalline
(PHB) onto cellulose via in situ reactive extrusion with dicumyl peroxide, morphology in stereocomplex of poly (L-lactic acid) with poly (D-lactic acid),
Biomacromolecules 16 (2015). Polymer (Guildf). 52 (2011) 68–76.
[58] C. Zhijiang, Y. Guang, J. Kim, Biocompatible nanocomposites prepared by im- [88] F. Gassner, A.J. Owen, Physical properties of poly(β-hydroxybutyrate)-poly(ɛ-ca-
pregnating bacterial cellulose nanofibrils into poly (3-hydroxybutyrate), Curr. prolactone) blends, Polymer (Guildf). 35 (1994) 2233–2236.
Appl. Phys. 11 (2011) 247–249. [89] M.A.T. Duarte, R.G. Hugen, E.S.A. Martins, A.P.T. Pezzin, S.H. Pezzin, Thermal
[59] H.S. Barud, et al., Bacterial cellulose/poly (3-hydroxybutyrate) composite mem- and mechanical behavior of injection molded poly(3-hydroxybutyrate/poly(ε-ca-
branes, Carbohydr. Polym. 83 (2011) 1279–1284. prolactone)) blends, Mater. Res. 9 (2006) 25–27.
[60] L. Zhang, X. Deng, S. Zhao, Z. Huang, Biodegradable polymer blends of poly (3- [90] M. Bothe, K.Y. Mya, E.M.J. Lin, C.C. Yeo, X. Lu, C. He, T. Pretsch, Triple-shape
hydroxybutyrate) and hydroxyethyl cellulose acetate, Polymer (Guildf). 38 (1997) properties of star-shaped POSS-polycaprolactone polyurethane networks, Soft
6001–6007. Matter 8 (2012) 965–972.
[61] L. Zhang, X. Deng, Z. Huang, Miscibility, thermal behaviour and morphological [91] C.P. Teng, K.Y. Mya, K.Y. Win, C.C. Yeo, M. Low, C. He, M.-Y. Han, Star-shaped
structure of poly (3-hydroxybutyrate) and ethyl cellulose binary blends, Polymer polyhedral oligomeric silsesquioxane-polycaprolactone-polyurethane as bioma-
(Guildf). 38 (1997) 5379–5387. terials for tissue engineering application, NPG Asia Mater. 6 (e142) (2014).
[62] J. Chen, D. Wu, K. Pan, Effects of ethyl cellulose on the crystallization and me- [92] X. Liu, X. Chen, M.X. Chua, Z. Li, X.J. Loh, Injectable Supramolecular Hydrogels as
chanical properties of poly(3-hydroxybutyrate), Int. J. Biol. Macromol. 88 (2016). Delivery Agents of Bcl-2 Conversion Gene for the Effective Shrinkage of
[63] M. Yamaguchi, K. Arakawa, Control of structure and mechanical properties for Therapeutic Resistance Tumors, 6 (2017), pp. 1–11.
binary blends of poly (3-hydroxybutyrate) and cellulose derivative, J. Appl. [93] Z. Li, J. Li, Control of hyperbranched structure of polycaprolactone/poly (ethylene
Polym. Sci. 103 (2007) 3447–3452. glycol) polyurethane block copolymers by glycerol and their hydrogels for po-
[64] H. Li, et al., Lignin-derived interconnected hierarchical porous carbon monolith tential cell delivery, J. Phys. Chem. B 117 (2013) 14763–14774.
with large areal/volumetric capacitances for supercapacitor, Carbon N. Y. 100 [94] X. Fan, X. Wang, M. Cao, C. Wang, Z. Hu, Y.-L. Wu, Z. Li, X.J. Loh, ‘Y’-shape armed
(2016) 151–157. amphiphilic star-like copolymers: design, synthesis and dual-responsive unim-
[65] D. Kai, K. Zhang, L. Jiang, H.Z. Wong, Z. Li, Z. Zhang, X.J. Loh, Sustainable and olecular micelle formation for controlled drug delivery, Polym. Chem. 8 (2017)
antioxidant lignin-polyester copolymers and nanofibers for potential healthcare 5611–5620.
applications, ACS Sustain. Chem. Eng. 5 (2017) 6016–6025. [95] Z. Li, X. Liu, X. Chen, M. Xuan, Y. Wu, Targeted delivery of Bcl-2 conversion gene
[66] S. Angelini, P. Cerruti, B. Immirzi, G. Santagata, G. Scarinzi, M. Malinconico, From by MPEG-PCL-PEI-FA cationic copolymer to combat therapeutic resistant cancer,
biowaste to bioresource: effect of a lignocellulosic filler on the properties of poly Mater. Sci. Eng. C 76 (2017) 66–72.
(3-hydroxybutyrate), Int. J. Biol. Macromol. 71 (2014) 163–173. [96] B. Vergara-Porras, J.N. Gracida-Rodriguez, F. Perez-Guevara, Thermal processing
[67] S. Angelini, P. Cerruti, B. Immirzi, G. Scarinzi, M. Malinconico, Acid-insoluble influence on mechanical, thermal, and biodegradation behavior in poly(β-hydro-
lignin and holocellulose from a lignocellulosic biowaste: bio-fillers in poly (3- xybutyrate)/poly(ϵ-caprolactone) blends: a descriptive model, J. Appl. Polym. Sci.
hydroxybutyrate), Eur. Polym. J. 76 (2016) 63–76. 133 (2016) 1–12.
[68] P. Mousavioun, P.J. Halley, W.O.S. Doherty, Thermophysical properties and [97] D. Garcia-Garcia, J.M. Ferri, T. Boronat, J. Lopez-Martinez, R. Balart, Processing
rheology of PHB/lignin blends, Ind. Crop. Prod. 50 (2013) 270–275. and characterization of binary poly(hydroxybutyrate) (PHB) and poly(capro-
[69] W. Cao, A. Wang, D. Jing, Y. Gong, N. Zhao, X. Zhang, Novel biodegradable films lactone) (PCL) blends with improved impact properties, Polym. Bull. 73 (2016)
and scaffolds of chitosan blended with poly (3-hydroxybutyrate), J. Biomater. Sci. 3333–3350.
Polym. Ed. 16 (2005) 1379–1394. [98] D. Garcia-Garcia, E. Rayon, A. Carbonell-Verdu, J. Lopez-Martinez, R. Balart,
[70] T. Ikejima, K. Yagi, Y. Inoue, Thermal properties and crystallization behavior of Improvement of the compatibility between poly(3-hydroxybutyrate) and poly(e-
poly (3-hydroxybutyric acid) in blends with chitin and chitosan, Macromol. Chem. caprolactone) by reactive extrusion with dicumyl peroxide, Eur. Polym. J. 86
Phys. 200 (1999) 413–421. (2017) 41–57.
[71] K. Raghunatha, H. Sato, I. Takahashi, Intermolecular hydrogen bondings in the [99] C. Guo, L. Zhou, J. Lv, Effects of expandable graphite and modified ammonium
poly (3-hydroxybutyrate) and chitin blends: their effects on the crystallization polyphosphate on the flame-retardant and mechanical properties of wood flour-
behavior and crystal structure of poly (3-hydroxybutyrate), Polymer (Guildf). 75 polypropylene composites, Polym. Polym. Compos. 21 (2013) 449–456.
(2015) 141–150. [100] S. Zhang, X. Sun, Z. Ren, H. Li, S. Yan, The development of a bilayer structure of
[72] F. Gassner, A.J. Owen, Some properties of poly(3-hydroxybutyrate)- poly(3-hy- poly(propylene carbonate)/poly(3-hydroxybutyrate) blends from the demixed
droxyvalerate) blends, Polym. Int. 39 (1996) 215–219. melt, Phys. Chem. Chem. Phys. 17 (2015) 32225–32231.
[73] H. Verhoogt, B.A. Ramsay, B.D. Favis, Polymer blends containing poly (3-hydro- [101] L. Zhou, G. Zhao, J. Yin, W. Jiang, Toughening poly(3-hydroxybutyrate) with
xyalkanoate)s. Polymer review, Polymer (Guildf). 35 (1994) 5155–5169. propylene carbonate plasticized poly(propylene carbonate), E-Polymers 14 (2014)
[74] K. Zhao, Y. Deng, J.C. Chen, G.Q. Chen, Polyhydroxyalkanoate (PHA) scaffolds 283–288.
with good mechanical properties and biocompatibility, Biomaterials 24 (2003) [102] A.M. El-Hadi, Improvement of the miscibility by combination of poly(3-hydroxy
1041–1045. butyrate) PHB and poly(propylene carbonate) PPC with additives, J. Polym.
[75] Z. Li, B.H. Tan, T. Lin, C. He, Recent advances in stereocomplexation of en- Environ. 25 (2016) 1–11.
antiomeric PLA-based copolymers and applications, Prog. Polym. Sci. 62 (2016) [103] P. Ma, D.G. Hristova-Bogaerds, P.J. Lemstra, Y. Zhang, S. Wang, Toughening of
22–72. PHBV/PBS and PHB/PBS blends via in situ compatibilization using dicumyl

1115
J.C.C. Yeo et al. Materials Science & Engineering C 92 (2018) 1092–1116

peroxide as a free-radical grafting initiator, Macromol. Mater. Eng. 297 (2012) carnauba fibers, Compos. Part B Eng. 43 (2012) 2827–2835.
402–410. [132] E.E. Tănase, M.E. Popa, M. Râpă, O. Popa, PHB/cellulose fibers based materials:
[104] M. Avella, E. Martuscelli, Poly-d-(-)(3-hydroxybutyrate)/poly(ethylene oxide) physical, mechanical and barrier properties, Agric. Agric. Sci. Procedia 6 (2015)
blends: phase diagram, thermal and crystallization behaviour, Polymer (Guildf). 608–615.
29 (1988) 1731–1737. [133] L.V. Scalioni, M.C. Gutiérrez, M.I. Felisberti, Green composites of poly (3-hydro-
[105] M. Avella, E. Martuscelli, P. Greco, Crystallization behaviour of poly(ethylene xybutyrate) and curaua fibers: morphology and physical, thermal, and mechanical
oxide) from poly(3-hydroxybutyrate)/poly(ethylene oxide) blends: phase struc- properties, J. Appl. Polym. Sci. 134 (2017).
turing, morphology and thermal behaviour, Polymer (Guildf). 32 (1991) [134] V. Ojijo, S. Sinha Ray, Processing strategies in bionanocomposites, Prog. Polym.
1647–1653. Sci. 38 (2013) 1543–1589.
[106] M. Avella, E. Martuscelli, M. Raimo, Properties of blends and composites based on [135] S.I. Yun, V. Lo, J. Noorman, J. Davis, R.A. Russell, P.J. Holden, G.E. Gadd,
poly(3-hydroxy)butyrate (PHB) and poly(3-hydroxybutyrate-hydroxyvalerate) Morphology of composite particles of single wall carbon nanotubes/biodegradable
(PHBV) copolymers, J. Mater. Sci. 35 (2000) 523–545. polyhydroxyalkanoates prepared by spray drying, Polym. Bull. 64 (2010) 99–106.
[107] X. Shuai, Y. He, Y. Na, Y. Inoue, Miscibility of block copolymers of poly (ɛ-ca- [136] S.I. Yun, G.E. Gadd, B.A. Latella, V. Lo, R.A. Russell, P.J. Holden, Mechanical
prolactone) and poly (ethylene glycol) with poly (3-hydroxybutyrate) as well as properties of biodegradable polyhydroxyalkanoates/single wall carbon nanotube
the compatibilizing effect of these copolymers in blends of poly (ɛ-caprolactone) nanocomposite films, Polym. Bull. 61 (2008) 267–275.
and poly (3-hydroxybutyrate), Polymer (Guildf). (2001) 2600–2608. [137] M. Huh, M.H. Jung, Y.S. Park, B.-J. Kim, M.S. Kang, P.J. Holden, S.I. Yun, Effect of
[108] M. Jakic, N. Stipanelov Vrandeecic, M. Erceg, Thermal degradation of poly(3- carbon nanotube functionalization on the structure and properties of poly(3-hy-
hydroxybutyrate)/poly(ethylene oxide) blends: thermogravimetric and kinetic droxybutyrate)/MWCNTs biocomposites, Macromol. Res. 22 (2014) 765–772.
analysis, Eur. Polym. J. 81 (2016) 376–385. [138] A. Botana, M. Mollo, P. Eisenberg, R.M. Torres Sanchez, Effect of modified mon-
[109] Y. Azuma, N. Yoshie, M. Sakurai, Y. Inoue, R. Chûjô, Thermal behaviour and tmorillonite on biodegradable PHB nanocomposites, Appl. Clay Sci. 47 (2010)
miscibility of poly(3-hydroxybutyrate)/poly(vinyl alcohol) blends, Polymer 263–270.
(Guildf). 33 (1992) 4763–4767. [139] E. Panayotidou, A. Kroustalli, A. Baklavaridis, I. Zuburtikudis, D.S. Achilias,
[110] A.A. Ol'Khov, A.L. Iordanskii, T.P. Danko, Morphology of poly(3-hydro- D. Deligianni, Biopolyester-based nanocomposites: structural, thermo-mechanical
xybutyrate)-polyvinyl alcohol extrusion films, J. Polym. Eng. 35 (2015) 765–771. and biocompatibility characteristics of poly(3-hydroxybutyrate)/montmorillonite
[111] Z. Bartczak, M. Grala, Toughening of semicrystalline and amorphous polylactide clay nanohybrids, J. Appl. Polym. Sci. 132 (2015) 1–11.
with atactic poly(hydroxy butyrate), Polym.-Plast. Technol. Eng. 56 (2016) 29–43. [140] O. Akin, F. Tihminlioglu, Effects of organo-modified clay addition and tempera-
[112] P. Greco, E. Martuscelli, Crystallization and thermal behaviour of poly(d(-)-3-hy- ture on the water vapor barrier properties of polyhydroxy butyrate homo and
droxybutyrate)-based blends, Polymer (Guildf). 30 (1989) 1475–1483. copolymer nanocomposite films for packaging applications, J. Polym. Environ. 0
[113] Robert S. Whitehouse, Lexington, M. (US), Toughened Polyhydroxyalkanoate (2017) 1–12.
Compositions, (2016). [141] M. Barletta, F. Trovalusci, M. Puopolo, V. Tagliaferri, S. Vesco, Engineering and
[114] Y. Parulekar, A.K. Mohanty, Biodegradable toughened polymers from renewable processing of poly(hydroxybutyrate) (PHB) modified by nano-sized graphene na-
resources: blends of polyhydroxybutyrate with epoxidized natural rubber and noplatelets (GNP) and amino-functionalized silica (A-fnSiO2), J. Polym. Environ.
maleated polybutadiene, Green Chem. 8 (2006) 206. 24 (2016) 1–11.
[115] Y.S. Parulekar, A.K. Mohanty, Biobased nanocomposites from toughened bacterial [142] J. Bian, H.L. Lin, G. Wang, Q. Zhou, Z.J. Wang, X. Zhou, Y. Lu, X.W. Zhao,
bioplastic and titanate modifed layered silicate: potential replacement for re- Morphological, mechanical and thermal properties of chemically bonded graphene
inforced TPO, Vth Annu. SPE Automot. Compos. Conf. (2005) 1–10. oxide nanocomposites with biodegradable poly (3-hydroxybutyrate) by solution
[116] Y. Parulekar, Methods of Making Nanocomposites and Compositons of Rubber intercalation, Polym. Polym. Compos. 24 (2016) 133–141.
Toughened Polyhydroxyalkanoates, (2008). [143] A.M. Gumel, M.H. Aris, M.S.M. Annuar, Polyhydroxyalkanoate (PHA) Based
[117] D. Graebling, P. Bataille, Polypropylene/polyhydroxybutyrate blends: preparation Blends, Composites and Nanocomposites, (2016), pp. 141–182.
of a grafted copolymer and its use as surface-active agent, Polym.-Plast. Technol. [144] T.L. Bluhm, G.K. Hamer, R.H. Marchessault, C.A. Fyfe, R.P. Veregin,
Eng. 33 (1994) 341–356. Isodimorphism in bacterial poly(B-hydroxybutyrate-co-B-hydroxyvalerate),
[118] W.M. Pachekoski, J.A.M. Agnelli, L.P. Belem, Thermal, mechanical and morpho- Macromolecules 19 (1986) 2871–2876.
logical properties of poly (hydroxybutyrate) and polypropylene blends after pro- [145] M. Kunioka, A. Tamaki, Y. Doi, Crystalline and thermal properties of bacterial
cessing, Mater. Res. 12 (2009) 159–164. copolyesters: poly(3-hydroxybutyrate-co-3-hydroxyvalerate) and poly(3-hydro-
[119] J.C. Girioli, D. dos Santos Rosa, P.A. dos Santos, Biodegradable Thermoplastic xybutyrate-co-4-hydroxybutyrate), Macromolecules 22 (1989) 694–697.
Compositions, (2010). [146] M. Kunioka, Y. Doi, Thermal degradation of microbial copolyesters:poly(3-hy-
[120] J.C. Girioli, D. dos Santos Rosa, P.A. dos Santos, Biodegadable Thermoplastic droxybutyrate-co-3-hydroxyvalerate) and poly(3-hydroxybutyrate-co-4-hydro-
Compositions, (2013), p. 1. xybutyrate), Macromolecules 23 (1990) 1933–1936.
[121] R.K. Sadi, R.S. Kurusu, G.J.M. Fechine, N.R. Demarquette, Compatibilization of [147] C. Zhu, Q. Chen, Polyhydroxyalkanoate-based biomaterials for applications in
polypropylene/poly(3-hydroxybutyrate) blends, J. Appl. Polym. Sci. 123 (2012) biomedical engineering, Adv. Healthc. Mater. (2014) 439–464.
3511–3519. [148] Y. Doi, S. Kitamura, H. Abe, Microbial synthesis and characterization of poly(3-
[122] H. Yang, Z.-S. Li, Z.-Y. Lu, C.-C. Sun, Computer simulation studies of the miscibility hydroxybutyrate-co-3-hydroxyhexanoate), Macromolecules 28 (1995) 4822–4828.
of poly(3-hydroxybutyrate)-based blends, Eur. Polym. J. 41 (2005) 2956–2962. [149] J.J. Fischer, Y. Aoyagi, M. Enoki, Y. Doi, T. Iwata, Mechanical properties and
[123] A.A. Ol'khov, A.L. Iordanskii, G.E. Zaikov, L.S. Shibryaeva, I.A. Litvinov, enzymatic degradation of poly([R]-3-hydroxybutyrate-co-[R]-3-hydro-
S.V. Vlasov, Morphologically special features of poly(3-hydroxybutyrate)/low- xyhexanoate) uniaxially cold-drawn films, Polym. Degrad. Stab. 83 (2004)
density polyethylene blends, Polym.-Plast. Technol. Eng. 39 (2000) 783–792. 453–460.
[124] Y.N. Pankova, A.N. Shchegolikhin, A.L. Iordanskii, A.L. Zhulkina, A.A. Ol'khov, [150] S.G. Hong, C.H. Lin, Improvement of thermal properties of polyhydroxybutyrate
G.E. Zaikov, The characterization of novel biodegradable blends based on poly- by grafted chemicals, E-Polymers 47 (2010) 1–14.
hydroxybutyrate: the role of water transport, J. Mol. Liq. 156 (2010) 65–69. [151] X. Li, X.J. Loh, K. Wang, C. He, J. Li, Poly (ester urethane) s consisting of poly [(R)-
[125] S. Wong, R. Shanks, A. Hodzic, Interfacial improvements in poly(3-hydro- 3-hydroxybutyrate] and poly (ethylene glycol) as candidate biomaterials: char-
xybutyrate)-flax fibre composites with hydrogen bonding additives, Compos. Sci. acterization and mechanical property study, Biomacromolecules 6 (2005)
Technol. 64 (2004) 1321–1330. 2740–2747.
[126] N.M. Barkoula, S.K. Garkhail, T. Peijs, Biodegradable composites based on flax/ [152] K.L. Liu, E.S.G. Choo, S.Y. Wong, X. Li, C.B. He, J. Wang, J. Li, Designing poly[(R)-
polyhydroxybutyrate and its copolymer with hydroxyvalerate, Ind. Crop. Prod. 31 3-hydroxybutyrate]-based polyurethane block copolymers for electrospun nano-
(2010) 34–42. fiber scaffolds with improved mechanical properties and enhanced mineralization
[127] A. Hodzic, R. Coakley, R. Curro, C.C. Berndt, R.A. Shanks, Design and optimization capability, J. Phys. Chem. B 114 (2010) 7489–7498.
of biopolyester bagasse fiber composites, J. Biobased Mater. Bioenergy 1 (2007) [153] H.F. Naguib, M.S.A. Aziz, S.M. Sherif, G.R. Saad, Synthesis and thermal char-
46–55. acterization of poly(ester-ether urethane)s based on PHB and PCL-PEG-PCL blocks,
[128] E.V. Torres-Tello, J.R. Robledo-Ortíz, Y. González-García, A.A. Pérez-Fonseca, J. Polym. Res. 18 (2011) 1217–1227.
C.F. Jasso-Gastinel, E. Mendizábal, Effect of agave fiber content in the thermal and [154] G. Adamus, W. Sikorska, H. Janeczek, M. Kwiecien, M. Sobota, M. Kowalczuk,
mechanical properties of green composites based on polyhydroxybutyrate or poly Novel block copolymers of atactic PHB with natural PHA for cardiovascular en-
(hydroxybutyrate-co-hydroxyvalerate), Ind. Crop. Prod. 99 (2017) 117–125. gineering: synthesis and characterization, Eur. Polym. J. 48 (2012) 621–631.
[129] S.J. Christian, S.L. Billington, Mechanical response of PHB-and cellulose acetate [155] L. Wu, S. Chen, Z. Li, K. Xu, G.-Q. Chen, Synthesis, characterization and bio-
natural fiber-reinforced composites for construction applications, Compos. Part B compatibility of novel biodegradable poly[((R)-3-hydroxybutyrate)-block-(D,L-
Eng. 42 (2011) 1920–1928. lactide)-block-(ε-caprolactone)] triblock copolymers, Polym. Int. 57 (2008)
[130] M.A. Gunning, L.M. Geever, J.A. Killion, J.G. Lyons, C.L. Higginbotham, 939–949.
Mechanical and biodegradation performance of short natural fibre poly- [156] D.C. Aluthge, C. Xu, N. Othman, N. Noroozi, S.G. Hatzikiriakos,
hydroxybutyrate composites, Polym. Test. 32 (2013) 1603–1611. P. Mehrkhodavandi, PLA-PHB-PLA triblock copolymers: synthesis by sequential
[131] J.D.D. Melo, L.F.M. Carvalho, A.M. Medeiros, C.R.O. Souto, C.A. Paskocimas, A addition and investigation of mechanical and rheological properties,
biodegradable composite material based on polyhydroxybutyrate (PHB) and Macromolecules 46 (2013) 3965–3974.

1116

You might also like