You are on page 1of 10

J Polym Eng 2017; aop

Chin-San Wu* and Hsin-Tzu Liao

Fabrication, characterization, and application of


polyester/wood flour composites
DOI 10.1515/polyeng-2016-0284 materials are not environmentally friendly. Furthermore,
Received August 1, 2016; accepted November 4, 2016 toxic substances can be released during printing, which
can cause serious harm to the environment and the
Abstract: The mechanical properties, thermal properties,
human body [7, 8].
antibacterial activity, and fabrication of three-dimensional
In response to the international trends of energy
(3D) printing strips of composite materials containing
savings, carbon reduction, and sustainable development,
polyhydroxyalkanoate (PHA) and wood flour (WF) were
more environmentally friendly aliphatic polyesters have
evaluated. Maleic anhydride (MA)-grafted PHA (PHA-g-
been developed; they can be decomposed by microorgan-
MA) and WF were used to enhance the desired character-
isms in the wild after being mixed with CO2, water, and
istics of these composites. The PHA-g-MA/WF composites
other harmless residues [9–11]. Among these polyesters,
had better mechanical properties than the PHA/WF com-
polyhydroxyalkanoate (PHA) is one of the most promising
posites did. This effect was attributed to a greater compat-
because microorganisms are polymerized by β-hydroxyl
ibility between the grafted polyester and WF. Additionally,
eicosanoic acid into aliphatic polyester [12]. This provides
the PHA-g-MA/WF composites provided higher quality 3D
PHA with good biocompatibility, biodegradability, non-
printing strips and were more easily processed because of
toxicity, processability, and mechanical properties [13, 14].
ester formation. The water resistance of the PHA-g-MA/
However, PHA’s high production cost is the main bottle-
WF composite was greater than that of PHA/WF. Moreo-
neck to its commercialization. Most PHA products cannot
ver, WF enhanced the antibacterial activity of the compos-
compete with petrochemical plastics in the market, limit-
ites. Composites of PHA-g-MA or PHA containing WF had
ing their commercial development and applications. The
better antibacterial activity.
linking of PHA with 3D printing technology is one way
Keywords: 3D printing strips; antibacterial activity; poly- to reduce PHA’s cost and increase its added-value. For
mer-matrix composites; wood flour. example, when PHA/wood flour (WF) composite is made
into 3D printing strips, the value and competitiveness of
the composite are enhanced.
1 Introduction This study used waste WF (Chamaecyparis formosen-
sis) as an additive. Chamaecyparis formosensis is hard,
Three-dimensional (3D) printing technology arose in contains wood alcohol, is moth-proof antibacterial and
the 1980s and has been gradually used in the field of deodorizing, and has anticancer properties. It is a valuable
rapid prototyping to accelerate commercial product crop for these reasons [15–17]. Chamaecyparis formosensis
research and development. It is free of expensive tools, is commonly used in buildings, furniture, woodcarvings,
instruments, and waste materials; 3D printing has been tableware, and other commodities [18, 19], but the manu-
identified as a key aspect of a new Industrial Revolution facturing process is accompanied by residues and chips
[1–3]. Fused deposition modeling is the most widely used that cause environmental problems. To reduce the waste
process for 3D printing strips; it uses acrylonitrile-buta- and increase the reuse effectiveness, C. formosensis chips
diene-styrene, polycarbonate, polyamide, and mixtures were ground into WF and mixed with PHA; then, maleic
of thermoplastic materials [4–6]. However, most of these anhydride (MA) was used as a compatibilizing agent to
enhance the adhesion between PHA and the WF [20]. This
composite was formed into a 3D printing strip, which dif-
fuses the aroma of wood alcohol during the 3D printing
*Corresponding author: Chin-San Wu, Department of Applied
process, thus increasing the added-value of the product.
Cosmetology, Kao Yuan University, Kaohsiung County, Taiwan 82101,
Republic of China, e-mail: t50008@cc.kyu.edu.tw
The waste from this composite strip after 3D printing can
Hsin-Tzu Liao: Department of Applied Cosmetology, Kao Yuan be decomposed by microorganisms in soil by composting
University, Kaohsiung County, Taiwan 82101, Republic of China and will not pollute the environment.

Brought to you by | Iowa State University


Authenticated
Download Date | 1/14/17 9:41 AM
2      C.-S. Wu and H.-T. Liao: Polyester/wood flour composites

2 Materials and methods layers of cheesecloth. The dichloromethane-insoluble


product remaining on the cheesecloth was washed with
acetone to remove the unreacted MA and was then dried
2.1 Materials
in a vacuum oven at 80°C for 24 h. The MA loading of the
dichloromethane-soluble polymer was determined by
Commercial-grade PHA (EM 5400F) was obtained from
titration and expressed as the grafting percentage [21]. The
Shenzhen Ecomann Biotechnology Co., Ltd. (Shenzhen,
grafting percentage was 0.96  wt%. The loadings of DCP
China). MA, obtained from Sigma (St. Louis, MO), was
and MA were maintained at 0.3 and 10 wt%, respectively.
purified before use by recrystallization from chloroform.
Dicumyl peroxide (DCP; Sigma) was used as an initiator
and was purified by dissolution in chloroform and repre-
cipitation in methanol. Other reagents were purified using 2.3 Preparation of WF and its composites
conventional methods. WF was obtained from Taiwan.
Chamaecyparis formosensis was purchased from the
Taiwan Forestry Bureau. After being cut at a timber mill,
2.2 M
 A-grafted PHA copolymer the residual dark brown coarse wood chips were crushed
by a coarse crusher, resulting in 0.2–1.0-cm-long fine
The reaction grafted MA onto PHA (PHA-g-MA). In a pre- wood chips. These chips were immersed in 1000 ml of dis-
liminary test using dichloromethane as the solvent in tilled water for 1 day to remove any water-soluble compo-
a round-bottomed flask, a mixture of MA and DCP was nents. After filtering, the wood chips were further ground
added in four equal portions at 2-min intervals to PHA to six times in a high-speed rotary grinder (Yowlin Industrial
allow grafting to take place. The reaction was performed Co. Ltd., Taipei, Taiwan) for 10  min and vacuum-dried
in a nitrogen (N2) atmosphere at 50 ± 2°C. Preliminary at 50–60°C for 1 day. The resulting dry wood chips were
experiments showed that equilibrium was attained in less then ground again and sieved through 400- and 500-mesh
than 12  h. Thus, reactions were allowed to progress for sieves and vacuum-dried for at least 12  h at 100–110°C
12 h with stirring at 60 rpm. Samples of 4 g of the product until the moisture content of the resulting dark brown
were dissolved in 200 ml of refluxing dichloromethane WF was 1.0 ± 0.3%. The entire WF processing procedure is
at 50 ± 2°C and the solution was filtered through several shown in Scheme 1.

WF
Drying at (Wood flour)
100~110°C
1 days
Crushed sieve
analysis

Wood chips

1 days
Wash
Water for
Drying at 1 days
50~60°C
Immerse
in distilled

Scheme 1: Modification of PHA with WF and the preparation of three-dimensional (3D) printing strips.

Brought to you by | Iowa State University


Authenticated
Download Date | 1/14/17 9:41 AM
C.-S. Wu and H.-T. Liao: Polyester/wood flour composites      3

The WF samples were washed with acetone and dried TA Instruments, New Castle, DE) operating in film tension
in an oven at 100–110°C for 1 day prior to composite prepa- mode. Test specimens were rectangular thin films measuring
ration. Composites were prepared in a Brabender (Dayton, 25 × 5 × 1 mm3. DMA tests were performed at a frequency of
OH, USA) “Plastograph” W50EHT 200-Nm mixer equipped 1 Hz and a strain level of 0.075% under conditions in which
with a rotor blade. The blends were mixed at 140–150°C for the temperature rose from − 60 to 70°C at a rate of 3°C min − 1.
15 min at a rotor speed of 50 rpm. The WF contents in the The glass transition temperature (Tg), melting temperature
hybrid materials, i.e. PHA/WF and PHA-g-MA/WF, were 0, (Tm), and heat of fusion (∆Hf) were determined using a dif-
10, 20, 30, and 40 wt%. After mixing, the composites were ferential scanning calorimeter (DSC) (Model 2010; TA Instru-
pressed into thin plates using a hot press and placed in a ments). The mass of the specimens ranged from 4–6  mg.
dryer for cooling. These thin plates were cut into standard Melting curves were recorded from − 40 to 180°C at a heating
sample dimensions for further characterization. rate of 10°C min − 1. A mechanical tester (Model LR5K; Lloyd
Instruments, Bognor Regis, West Sussex, UK) was used to
measure the tensile strength at failure, in accordance with the
2.4 F abrication of 3D printing strips and American Society for Testing and Materials (ASTM) Standard
products D638. Test samples were prepared using a hydraulic press at
150°C and conditioned at a relative humidity of 50 ± 5% for
An extruder (Model TUS-194T; Atlas Electric Devices 24 h prior to taking measurements. Testing was carried out at
Company, Chicago, IL) was used for making the 3D print- a crosshead speed of 10 mm min − 1. Mean values were deter-
ing strips at a fabrication temperature of 140–150°C and a mined from the values of five specimens. A thin membrane
screw rotation rate of 50 rpm. A sample (PHA or PHA-g-MA of each composite was prepared with a hydraulic press and
and their composites) was placed in the single extruder heated in a vacuum oven at 60°C for 24 h to obtain samples
and heated until it was molten. The sample was slowly with dimensions of 150 × 150 × 1  mm3. Specimens were cut
extruded and cooled by a cooling system. The coiler speed according to ASTM D638 specifications. Following failure, a
was used to control the diameter of the 3D printing strip thin section of the fracture plane was removed, coated with
to 1.75 ± 0.05  mm. Scheme 1 shows the fabrication of 3D gold and imaged using scanning electron microscopy (SEM;
printing strips and products from the composite materials. Model S-1400; Hitachi, Tokyo, Japan) to observe the fracture
surface morphology.

2.5 Characterization
2.7 Water absorption
Solid-state 13C nuclear magnetic resonance (NMR) spectra
were acquired using an AMX-400 NMR spectrometer (Bruker,
Samples were prepared for water absorption measurements
Billerica, MA) at 100  MHz under cross-polarization, while
by cutting the composites into 45 × 25 mm2 strips (150 ± 5 μm
spinning at the magic angle. Power decoupling conditions
in thickness) in accordance with ASTM Standard D570. The
were set with a 90° pulse and a 4-s cycle time. A contact time
samples were dried in a vacuum oven at 50 ± 2°C for 12 h,
of 2.0 ms was used. Samples were loaded into 4-mm fused
cooled in a desiccator, and then immediately weighed to
zirconia tubes and sealed with Kel-FTM caps. Spectra were
the nearest 0.001 g. This mass was designated mc. Thereaf-
obtained at a spinning rate of ca. 4700 Hz. Infrared spectra of
ter, the samples were immersed in distilled water and main-
the samples were obtained using an FTS-7PC Fourier-trans-
tained at 30 ± 2°C for a 120-day period, during which they
form infrared (FT-IR) spectrophotometer (Bio-Rad, Hercules,
were removed from the water at 20-day intervals, gently
CA). The spectra were acquired at 2 cm − 1 resolution between
blotted with tissue paper to remove excess water from the
400 and 4000 cm − 1 with collection times of ca. 1 min. X-ray
surfaces, immediately weighed to the nearest 0.001 g three
diffraction (XRD) data were recorded using a D/max 3-V
times, and then returned to the water. The average of the
X-ray diffractometer (Rigaku, Tokyo, Japan) with a Cu target
mass measured at the 20-day intervals was calculated;
and Kα radiation at a scanning rate of 2° min − 1.
these average masses were designated mw. The percentage
mass increase because of water absorption, mf, which was
calculated to the nearest 0.01% as follows:
2.6 Thermal properties

The phase compatibility in the composites was evaluated mw − mc


%m f = × 100%
using dynamic mechanical analysis (DMA) (Model 2080; mc (1)

Brought to you by | Iowa State University


Authenticated
Download Date | 1/14/17 9:41 AM
4      C.-S. Wu and H.-T. Liao: Polyester/wood flour composites

2.8 Antibacterial assay


Escherichia coli (BCRC10239) was obtained from the Biore-
source Collection and Research Center, Hsinchu, Taiwan.
E. coli was maintained in nutrient broth medium (3 g of
beef extract and 5 g of peptone in 1 l of distilled water, pH
7.0). The samples were incubated at 37 ± 1°C at a relative
humidity of ca. 90%.
The Japanese Industrial Standard JIS Z 2801:2000, a
method typically used to estimate the antibacterial activ-
ity, was used to determine the antibacterial activity of
membranes of PHA or PHA-g-AA and its WF composites
[22]. This method determines the antibacterial index (ABI)
and kill-bacterial index (KBI) according to the following:
ABI = log B − log C (2)

KBI = log A − log C (3) Figure 1: FT-IR spectra of (A) PHA, (B) PHA-g-MA, (C) PHA/WF
(20 wt%), (D) PHA-g-MA/WF (20 wt%), and (E) WF.
where A is the number of bacteria recovered from the
inoculated unfilled WF sample (native PHA or PHA-g-MA)
shoulders observed at 1785 and 1851 cm − 1 for PHA-g-MA.
immediately following inoculation, B is the number of
These features are characteristic of MA groups; similar
bacteria remaining in the inoculated unfilled WF sample
results have been reported previously [23]. The shoulders
after 18 h, and C is the number of bacteria remaining in the
in the spectra represent the anhydride carboxyl groups in
inoculated filled WF sample after 18 h.
the modified polymer, signifying grafting of MA onto PHA.
Changes in the peaks at 2800–3000 cm −1 in Figure 1A–D
were attributed to changes in alkane C-H bonds caused by
2.9 Bacteria inhibition
grafting of MA onto PHA and/or composite formation of
WF with PHA.
Escherichia coli cells were activated in the liquid medium
Peaks at 3200–3700  cm − 1 correspond to the O-H
and then transferred to fresh medium for 18 h of cultiva-
stretching mode of the -OH group of WF; these were more
tion. The samples (100 μl) were then dropped into the
intense in the PHA/WF (20  wt%) composite (Figure 1C)
culture dish. Next, a sterilized triangular glass rod was
and appeared with an extra shoulder at 3286 cm − 1. Further
used to spread the bacteria in the dish. Additionally, the
comparison of the spectra of the PHA/WF (20  wt%) and
specimen (ca. 1.30 cm in diameter and 0.05 cm thick) was
PHA-g-MA/WF (20 wt%) hybrids (Figure 1C and D) revealed
wiped with 75% alcohol for sterilization and placed into
that the vibrational band of the -OH group in WF shifted
a culture dish with clippers. The culture dish edges were
from 3286 to 3221  cm − 1. The above observations demon-
sealed using sealing membrane and placed in an incuba-
strated the existence of strong hetero-associated hydrogen
tor at 37 ± 1°C for 1 day of cultivation to evaluate the bacte-
bonds between carboxylic groups of the PHA-g-MA matrix
rial inhibitory effects.
and saccharide hydroxyl groups of the WF. Notably, the
FT-IR spectrum of the PHA-g-MA/WF (20  wt%) compos-
ite (Figure 1D) exhibited a peak at 1736 cm − 1 that was not

3 Results and discussion present in the FT-IR spectrum of the PHA/WF (20  wt%)
composite. This peak was assigned to the ester carbonyl
stretching vibration of the PHA-g-MA/WF (20 wt%) com-
3.1 C
 haracterization of PHA and its posite [24]. These data suggest the formation of branched
composites and cross-linked macromolecules in the PHA-g-MA/WF
composite via covalent bonding of the anhydride car-
Figure 1A and B shows the FT-IR spectra of unmodified boxyl groups in PHA-g-MA with the saccharide hydroxyl
PHA and PHA-g-MA, respectively. The characteristic transi- groups of WF. Hence, an interfacial interaction was clearly
tions of PHA at 3300–3700, 1700–1760, and 500–1500 cm − 1 evident in PHA-g-MA/WF. The WF spectrum in Figure 1E is
appeared in the spectra of both polymers, with two extra similar to that reported by Zhang et al. [25].

Brought to you by | Iowa State University


Authenticated
Download Date | 1/14/17 9:41 AM
C.-S. Wu and H.-T. Liao: Polyester/wood flour composites      5

The solid-state 13C NMR spectra of PHA and PHA-g-


MA are shown in Figure 2A and B, respectively. Three
peaks were observed, corresponding to carbon atoms in
the unmodified PHA (1, 6: δ = 169.3; 2: δ = 38.5; 3: δ = 72.0;
4: δ = 26.9; 5: δ = 9.2; 7: δ = 40.9; 8: δ = 67.7; 9: δ = 19.6 ppm)
[26]. The 13C NMR spectrum of PHA-g-MA exhibited addi-
tional peaks (10: δ = 36.1; 11: δ = 42.3; 12: δ = 171.9  ppm).
These peaks confirmed the grafting of MA onto PHA [27].
The 13C NMR spectra of PHA/WF (20  wt%), PHA-g-
MA/WF (20 wt%), and WF are shown in Figure 2C–E. The
spectra of the WF composites are similar to those reported
by Delmotte et  al. [28]. Relative to the unmodified poly-
ester, additional peaks (peaks 10–12, Figure 2B) were
observed in the spectra of composites containing PHA-g-
MA. However, the peak at δ = 171.9 ppm (C=O) (peak 12 of
Figure 2B), which is also typical for MA grafted onto PHA,
was absent in the solid-state spectrum of PHA-g-MA/WF
(20  wt%). This absence is probably because of an addi-
tional condensation reaction between the anhydride car-
boxyl group of MA and the OH group of WF, which caused
the peak at δ = 171.9 ppm to split into two bands (at δ = 175.9
and 178.7  ppm). This additional reaction converted the
-OH groups in the original WF to esters (represented by
peaks 13 and 14 in Figure 2D) and did not occur between
PHA and WF, as indicated by the absence of correspond-
ing peaks in the FT-IR spectrum of PHA/WF (20 wt%) in
Figure 1D. The formation of ester groups significantly
affected the mechanical properties of PHA-g-MA/WF, and
this is discussed in greater detail in the following sections. Figure 2: Solid-state 13C NMR spectra of (A) PHA, (B) PHA-g-MA, (C)
The crystallization properties of neat PHA, PHA/WF PHA/WF (20 wt%), (D) PHA-g-MA/WF (20 wt%), and (E) WF.
(20  wt%), PHA-g-MA/WF (20  wt%), and WF were also
examined by XRD (Figure 3). Similar to the result proposed
by Ribeiro et al. [29], neat PHA (Figure 3A) exhibited seven
peaks at ca. 13.1°, 16.8°, 19.8°, 21.5°, 22.5°, 25.5° and 27.2°.
Comparison of the XRD patterns of neat PHA and the PHA/
WF composite (Figure 3A and B, respectively) indicated
that there was only one extra peak, at about 15.3°, in the
spectrum of the latter, which derived from the structure of
WF [30]. This result revealed that the WF was physically
dispersed in the PHA matrix. Compared with Figure 3B, a
new peak appeared at 18.1° in the XRD pattern of the PHA-
g-MA/WF composite (Figure 3C) [31].

3.2 T
 hermal properties of PHA, PHA-g-MA Figure 3: XRD patterns for (A) PHA, (B) PHA/WF, (C) PHA-g-MA/WF
and its composites and (D) WF.

Figure 4 shows the effects of WF content and mixing time the melt. Therefore, the resulting polymer composite con-
on the melt torque of the PHA/WF and PHA-g-MA/WF tained powdered filler. The torque decreased with increas-
composites. When preparing the PHA/WF or PHA-g-MA/ ing WF content and mixing time and leveled off after 8 min
WF, the polymer was first melted and WF was added to of mixing. The final torque decreased with increasing WF

Brought to you by | Iowa State University


Authenticated
Download Date | 1/14/17 9:41 AM
6      C.-S. Wu and H.-T. Liao: Polyester/wood flour composites

A 1010

E′ (Pa)
WF wt%
109
40
20
20
0
PHA-g-MA/WF 0
PHA/WF
108
– 60 – 40 – 20 0 20 40 60

B
108

E′′ (Pa)
WF wt%
40
Figure 4: Torque values as a function of mixing time for the PHA/WF 20
20
and PHA-g-MA/WF composites for various WF contents. 0
107 0

– 40 – 20 0 20 40 60
content because the melt torque of WF was lower than
that of PHA or PHA-g-MA. Additionally, because of the soft C 2.0
WF wt%
PHA-g-MA/WF
ester groups formed by reaction of PHA-g-MA and WF, the 0
1.5 0 PHA/WF
melt torque of the PHA-g-MA/WF was significantly lower 20
Tan(δ)

than that of the PHA/WF for a given WF content. 1.0


20
40
The dynamic mechanical properties of the PHA/WF
0.5
and PHA-g-MA/WF composites were measured and used to
evaluate the compatibility of hybrids. Figure 5A–C shows 0.0
the variation of the storage modulus (E′), the loss tensile – 40 – 20 0 20 40 60
modulus (E″), and the loss tangent (tan δ) as a function
of temperature for the neat PHA/WF and PHA-g-MA/WF Figure 5: Dynamic viscoelastic behavior of (A) storage modulus (E′),
(B) the loss tensile modulus (E″), and (C) the loss tangent (tan δ) of
composites. Regardless of the WF content, E′ (Figure 5A)
PHA/WF and PHA-g-MA/WF composites containing various amounts
decreased whilst E″ (Figure 5B) increased with increasing
of WF.
temperature, and there existed a transition at ca. – 5 to 3°C
for E″. These phenomena may be related to the Tg of the
PHA, which occurs at this temperature. At the Tg, the heat of and decreased in intensity with increasing WF content.
deformation and molecular motions lead to energy dissipa- It is clear that the damping temperature of the PHA/WF
tion. Note that the temperature dependences of E′ and E″ for composites only slightly increased (by ca. 5.0°C) as the WF
the PHA-g-MA/WF composites were lower than those for the content increased from 0 to 20 wt%, suggesting a smaller
PHA/WF composites; i.e. E′ and E″ of the PHA-g-MA blends enhancement in the restricted motion of the amorphous
remained relatively high over a wide temperature range. phase of the PHA/WF composites; this related to the poor
This was attributed to the formation of ester bonds. compatibility between the WF and PHA phases. Further-
Figure 5C shows that tan δ increased sharply at a more, these results indicated only a small enhancement
particular temperature, which corresponded to the onset in the restricted motion of the amorphous phase and a
of segmental motion. The temperature of maximum decrease in the chain mobility of the PHA-g-MA/WF com-
damping shifted to higher values with increasing WF posites because of the poor compatibility between the
content because of the underlying influence of the Tg PHA and WF. The ΔHf, Tm, and Tg of the PHA/WF and PHA-
of the amorphous phase in the semicrystalline PHA-g- g-MA/WF composites with different WF contents were
MA. In the presence of 20 wt% of WF, the peak damping determined using DSC. The results are shown in Figure 6
temperature had shifted by ca. 9.0°C. The shift resulted and Table 1. For both composites, the Tm decreased with
from enhanced restriction of the motion of the amor- increasing WF content, reflected by the aforementioned
phous phase, suggesting increased adhesion between the decreasing molecular weight [32]. For the same WF
polymer and WF in these composites. In the case of the content, the PHA/WF composite had a higher Tm than did
PHA/WF composites, the tan δ curve generally broadened the PHA-g-MA/WF composite. This trend was consistent

Brought to you by | Iowa State University


Authenticated
Download Date | 1/14/17 9:41 AM
C.-S. Wu and H.-T. Liao: Polyester/wood flour composites      7

PHA-g-MA/WF were ca. 4–10 J g − 1 higher than those of


PHA/WF, probably because of condensation reactions,
as discussed above. The ΔHf may be used as an indicator
of blend crystallinity. Although the ΔHf of both the PHA/
WF and PHA-g-MA/WF blends decreased with increasing
WF content (Table 1), the extent of the decrease was sig-
nificantly greater in PHA/WF, indicating a lower degree of
crystallinity. These results are similar to those obtained else-
where with composites of polymers and natural fibers [34].

3.3 M
 orphology and mechanical properties
of PHA, PHA-g-MA and their composites

For most composites, effective wetting and uniform dis-


persion of all components in a given matrix, as well as
strong interfacial adhesion between phases, is required
for satisfactory mechanical properties. Here, WF may be
considered as a dispersed phase within the PHA or PHA-
Figure 6: DSC heating thermograms of PHA-g-MA and its blends g-MA matrix. SEM images were acquired to evaluate the
with different contents of WF (A (0 wt%), B (10 wt%), C (20 wt%), morphology of the composite and to examine the tensile
D (30 wt%), and E (40 wt%)). fracture surfaces of PHA/WF (20  wt%) and PHA-g-MA/
WF (20  wt%). The micrographs of PHA/WF (20  wt%)
Table 1: Effect of WF content on the thermal properties of PHA/WF (Figure  6A) revealed that the WF in this composite was
and PHA-g-MA/WF composites.
evenly dispersed throughout the matrix. However, poor
adhesion between the WF and the PHA matrix occurred
WF (wt%) PHA/WF PHA-g-MA/WF
because of the formation of hydrogen bonds within WF
Tg Tm ∆Hf Tg Tm ∆Hf and the difference between the hydrophilicities of PHA
0 − 9.8 136.0 41.3 − 10.0 135.6 40.2 and WF. Poor wetting in these composites was also noted
10 − 7.0 135.1 35.0 − 3.3 134.2 38.9 (Figure 7A), which was attributed to the large difference
20 − 4.9 134.0 29.8 − 0.1 132.9 37.1 in surface energies between the PHA and WF matrix [35].
30 − 4.0 133.1 26.9 0.9 131.6 35.1
The SEM micrograph of PHA-g-MA/WF (20  wt%) shown
40 − 3.1 132.3 23.9 1.6 131.3 33.9
in Figure 7B revealed more uniform adhesion and better
wetting of WF in the PHA-g-MA matrix, as indicated by
with the corresponding torque measurements shown in the complete coverage of PHA-g-MA on the WF and the
Figure 4. The lower melt torque indicated that PHA-g-MA/ removal of both materials when WF was pulled from the
WF was more processable than PHA/WF. On the other bulk. The improved interfacial adhesion was attributed
hand, the Tg increased with increasing WF content for to similar hydrophilicities of PHA-g-MA and WF, which
both PHA/WF and PHA-g-MA/WF. This increase is prob- allowed the formation of bonds between them.
ably a result of the reduced space available for molecular Figure 7C shows the tensile strength at failure as a
motion with increasing WF content in the composites [33]. function of the WF content in the PHA/WF and PHA-g-MA/
The Tg values were higher for the PHA-g-MA composites by WF composites. The tensile strength at failure of neat PHA
approximately 6.5–12.5°C. As comparison with the values was 16.3 MPa and decreased to 15.8 MPa following grafting
examined by the DMA test (Figure 5), it can be seen that with MA. For the PHA/WF composites, the tensile strength
the DMA examinations gave slightly high Tg values due to at failure markedly decreased with increasing WF content
larger and thicker samples. and was attributed to poor dispersion of WF in the PHA
The ΔHf of pure PHA and PHA-g-MA were 41.3 and matrix. The effect of this incompatibility on the mechani-
40.2 J g − 1, respectively. The lower ΔHf of PHA-g-MA is likely cal properties of the composites was substantial. The PHA-
to be because of grafted branches that disrupted the reg- g-MA/WF composites, however, exhibited an increase in
ularity of the chain structures in PHA and increased the tensile strength at failure with increasing WF content,
spacing between polymer chains. The values of ΔHf for reaching saturation at WF > 20  wt%. This behavior was

Brought to you by | Iowa State University


Authenticated
Download Date | 1/14/17 9:41 AM
8      C.-S. Wu and H.-T. Liao: Polyester/wood flour composites

Figure 8: Percent mass gain from the absorption of water for PHA,
PHA-g-MA, PHA/WF, and PHA-g-MA/WF composites.

the PHA-g-MA/WF composites was moderate, and it is


probable that the hydrophobicity of WF was enhanced
by interactions with PHA-g-MA. For both PHA/WF and
PHA-g-MA/WF, the percent mass gain over the 120-day
test period increased with increasing WF content. As the
arrangement of the polymer chains in these systems was
expected to be random, the above results were probably
due to decreased chain mobility at high WF contents and
the hydrophilicity of WF, which adhered weakly to the
more hydrophobic PHA matrix.

3.5 A
 ntibacterial activity of PHA, PHA-g-MA,
and its composites

Figure 9 illustrates the antibacterial properties of PHA,


Figure 7: SEM images showing the distribution and wetting of WF
in (A) PHA/WF (20 wt%) and (B) PHA-g-MA/WF (20 wt%) composites.
PHA-g-MA, and their composites. Figure 9A shows that the
(C) Effect of WF content on the tensile strength at failure for PHA/WF composites displayed antibacterial zones; pure PHA-g-MA
and PHA-g-MA/WF composites. did not possess antibacterial properties. However, some
micro-antibacterial zones were found with the PHA-g-MA/
WF membrane specimens having 20 wt% WF content. The
attributed to improved dispersion of WF in the PHA-g-MA antibacterial zone became more apparent with increasing
matrix, resulting from the formation of branched or cross- WF content because of easier detection.
linked macromolecules. The tensile strength of the PHA-g- Antibacterial activity was evaluated using clinically
MA/WF composites was ca. 6–18 MPa greater than that of infectious E. coli. Figure 9B shows that when in contact
the PHA/WF composites. with PHA or PHA-g-MA, the E. coli cell count increased
with time from 1.36 × 106 to 2.36 × 108 CFU ml − 1 and from
1.36 × 106 to 1.30 × 108 CFU ml − 1, respectively, after incu-
3.4 W
 ater absorption of PHA, PHA-g-MA, and bation at 37°C for 24 h. Conversely, under the same con-
their composites ditions, the bacterial cell count rapidly decreased to
nearly zero when in contact with PHA/WF or PHA-g-MA/
At a given fiber content, the PHA-g-MA/WF composites WF membranes containing more than 20  wt% WF. At
had greater resistance to water absorption than did the 20  wt% WF, the onset of E. coli reduction was observed
PHA/WF composites (Figure 8). The water resistance of at 3  h, which was significantly shorter than the ca. 6  h

Brought to you by | Iowa State University


Authenticated
Download Date | 1/14/17 9:41 AM
C.-S. Wu and H.-T. Liao: Polyester/wood flour composites      9

A corresponding samples of PHA/WF. This antibacterial


effect behavior was enhanced in PLA-g-MA/WF because
the PHA-g-MA is stabilized in a fixed orientation by WF
resulting in bacteria death due to increasing effective con-
centration of WF to bacteria or contacting time elongation.

4 Conclusions
The fabrication, structure, antibacterial activity, and
mechanical and thermal properties of 3D printing strips
made of PHA/WF and PHA-g-MA/WF composites were
PHA
PHA-g-MA investigated. FT-IR, 13C-NMR, and XRD analyses revealed
that condensation reactions occurred between WF
PHA/WF (10 wt%)
PHA-g-MA/WF (10 wt%)
B PHA/WF (20 wt%)
PHA-g-MA/WF (20 wt%) and anhydride carboxylic groups in PHA-g-MA, which
1010 PHA/WF (20 wt%)
resulted in significantly different structures in the two
PHA-g-MA/WF (20 wt%)
Concentration of E. coli (CFU)

109
108 materials. The morphology of the various PHA-g-MA/WF
107 composites was consistent, with good adhesion between
106 the WF phase and the PHA-g-MA matrix. Mechanical
105 tests showed that improved adhesion between WF and
104
PHA-g-MA enhanced the mechanical properties of the
103
102 composite, especially the tensile strength. Although
101 DSC tests revealed that the melting temperatures of both
100 PHA/WF and PHA-g-MA/WF decreased with increasing
0 3 6 9 12 15 18
Exposure time (h) WF content, the PHA-g-MA/WF composites were more
processable because of their lower melting tempera-
Figure 9: Exposure time course of the (A) inhibition zones and (B)
tures. The glass transition temperature of PHA-g-MA/WF
growth of E. coli cells during exposure to PHA or PHA-g-MA and its
composite surfaces.
was higher than that of PHA/WF, indicating more hin-
dered molecular motion. The water resistance of PHA-g-
MA/WF was greater than that of PHA/WF. Antibacterial
onset time observed for the 20 wt% WF composites. This activity was enhanced with the addition of 20  wt% WF
effect is probably because of a difference in WF surface to PHA-g-MA or PHA, resulting in ABI and KBI values of
densities. As noted in Table 2, according to the guidelines more than 8.00 and 6.00, respectively. The antibacte-
set forth by JIS L 1902:1998 and the JAFET, PHA/WF and rial properties of PHA-g-MA/WF were superior to those
PHA-g-MA/WF suppressed the growth of E. coli. Further- of PHA/WF. Thus, the current study demonstrated an
more, all of the samples containing PHA-g-MA/WF exhib- enhancement in the compatibility between PHA and WF,
ited a higher degree of bacterial suppression than did the which resulted in improved antibacterial properties.

Table 2: Effect of WF content on the antibacterial properties of PHA/WF and PHA-g-MA/WF composites.

  PHA/WF   PHA-g-MA/WF
   
10 wt%   20 wt%   40 wt% 10 wt%   20 wt%   40 wt%

A   1.36 × 106   1.36 × 106   1.36 × 106   1.36 × 106   1.36 × 106   1.36 × 106


B   2.36 × 108   2.36 × 108   2.36 × 108   1.30 × 108   1.30 × 108   1.30 × 108
C   4.91 × 103   0   0   2.31 × 103   0   0
ABI   4.68   > 8.37   > 8.37   4.75   > 8.11   > 8.11
KBI   2.44   > 6.13   > 6.13   2.77   > 6.13   > 6.13

A, number of bacteria recovered from the inoculated unfilled WF sample (native PHA or PHA-g-MA) immediately following inoculation;
B, number of bacteria remaining in the inoculated unfilled WF sample after 18 h; C, number of bacteria remaining in the inoculated filled WF
sample after 18 h.

Brought to you by | Iowa State University


Authenticated
Download Date | 1/14/17 9:41 AM
10      C.-S. Wu and H.-T. Liao: Polyester/wood flour composites

Acknowledgments: The authors thank the Ministry of [15] Imberger KT, Chiu CY. Soil Biol. Biochem. 2002, 34, 711–720.
Science and Technology (Taipei City, Taiwan, R.O.C.) for [16] Pavela R. Ind. Crop Prod. 2015, 76, 174–187.
[17] Hsieh YH, Kuo PM, Chien SC, Shyur LF, Wang SY. Phytomedicine
financial support (MOST 104-2221-E-244-011-).
2007, 14, 675–680.
[18] Obayashi J, Okochi T. Dendrochronologia 2013, 31, 52–57.
[19] Chen Z, Lei Q, He RL, Zhang ZF, Chowdhury AJK. Saudi J. Biol.
References Sci. 2016, 23, 142–147.
[20] Wu CS. Polym. Degrad. Stab. 2009, 94, 1076–1084.
[1] Stansbury JW, Idacavage MJ. Dent. Mater. 2016, 32, 54–64. [21] Wu CS. Mater. Sci. Eng. C 2015, 48, 310–319.
[2] Rayna T, Striukova L. Technol. Forecasting Social Change. 2016, [22] Wu CS. Mater. Sci. Eng. C 2016, 69, 27–36.
102, 214–224. [23] Ma P, Cai X, Lou X, Dong W, Chen M, Lemstra PJ. Polym. Degrad.
[3] Berman B. Bus. Horiz. 2015, 55, 155–162. Stabil. 2014, 100, 93–100.
[4] Postiglione G, Natale G, Griffini G, Levi M, Turri S. Composites [24] Vaidya AA, Gaugler M, Smith DA. Carbohydr. Polym. 2016, 136,
Part A 2015, 76, 110–114. 1238–1250.
[5] Casavola C, Cazzato A, Moramarco V, Pappalettere C. Mater. [25] Zhang K, Zong L, Tan Y, Ji Q, Yun W, Shi R, Xia Y. Carbohydr.
Des. 2016, 90, 453–458. Polym. 2016, 136, 121–127.
[6] Ning F, Cong W, Qiu J, Wei J, Wang S. Composites Part B 2015, [26] Phukon P, Radhapyari K, Konwar BK, Khan R. Mater. Sci. Eng. C
80, 369–378. 2014, 37, 314–320.
[7] Nuñez PJ, Rivas A, García-Plaza E, Beamud E, Sanz-Lobera A. [27] Schoukens G, Martins J, Samyn P. Polymer 2013, 54,
Procedia Eng. 2015, 132, 856–863. 349–362.
[8] McCullough EJ, Yadavalli VK. J. Mater. Process. Technol. 2013, [28] Delmotte L, Ganne-Chedeville C, Leban JM, Pizzi A, Pichelin F.
213, 947–954. Polym. Degrad. Stab. 2008, 93, 406–412.
[9] Terzopoulou ZN, Papageorgiou GZ, Papadopoulou E, [29] Ribeiro PLL, da Silva ACMS, Filho JAM, Druzian JI. Ind. Crop
­Athanassiadou E, Alexopoulou E, Bikiaris DN. Ind. Crop Prod. Prod. 2015, 69. 212–223.
2015, 68, 60–79. [30] Liu R, Peng Y, Cao J, Chen Y. Compos. Sci. Technol. 2014, 103,
[10] Bikiaris DN. Polym. Degrad. Stab. 2013, 98, 1908–1928. 1–7.
[11] Commereuc S, Askanian H, Verney V, Celli A, Marchese P, Berti [31] Yang Y, Tang Z, Xiong Z, Zhu J. Int. J. Biol. Macromol. 2015, 77,
C. Polym. Degrad. Stab. 2013, 98, 1321–1328. 273–279.
[12] Kumar M, Gupta A, Thakur IS. Bioresour. Technol. 2016, 213, [32] Tnasea EE, Popa ME, Râpa M, Popa O. Agric. Agric. Sci. Proce-
249–256. dia 2015, 6, 608–616.
[13] Modjinou T, Lemechko P, Babinot J, Versace DL, Langlois V, [33] Majhi SK, Nayak SK, Mohanty S, Unnikrishnan L. Int. J. Plast.
Renard E. Eur. Polym. J. 2015, 68, 471–479. Technol. 2010, 14, S57–S75.
[14] Laycock B, Halley P, Pratt S, Werker A, Lan P. Prog. Polym. Sci. [34] Wu CS, Liao HT. Polym. Bull. 2013, 70, 3443–3462.
2013, 38, 536–583. [35] Abdullah NM, Ahmad I. Fibers Polym. 2013, 14, 584–590.

Brought to you by | Iowa State University


Authenticated
Download Date | 1/14/17 9:41 AM

You might also like