Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Polym. Bull.

DOI 10.1007/s00289-015-1535-9

ORIGINAL PAPER

Accelerated hydrolytic degradation of poly(lactic acid)


achieved by adding poly(butylene succinate)

Yang-peng Wang1 • Yan-jun Xiao1 •


Jin Duan1 • Jing-hui Yang1 • Yong Wang1 •

Chao-liang Zhang2

Received: 27 April 2015 / Revised: 17 August 2015 / Accepted: 30 September 2015


Ó Springer-Verlag Berlin Heidelberg 2015

Abstract In this work, different contents (10–40 wt%) of poly(butylene succinate)


(PBS) were introduced into poly(lactic acid) (PLA) through the common melt
compounding processing. The microstructure and morphologies of the blends were
investigated through differential scanning calorimetry (DSC), wide-angle X-ray
diffraction (WAXD) and scanning electron microscope (SEM). The results showed
that the presence of PBS neither induces the crystallization nor enhances the
crystallinity of PLA matrix, and completely amorphous PLA was obtained in all
samples. PBS exhibited dispersed particles in the PLA matrix and there were clear
gaps between components. The hydrophilicity of samples was evaluated by mea-
suring contact angles. The results demonstrated that adding PBS improved the
hydrophilicity of samples. The hydrolytic degradation measurements were carried
out at 37 °C in alkaline solution. The results showed that the presence of PBS
accelerated the hydrolytic degradation of PLA matrix. Specifically, the higher the
content of PBS was, the bigger the weight loss per unit area of sample was. The
hydrolytic degradation mechanism was then analyzed.

Keywords Poly(L-lactide)/poly(butylene succinate)  Morphology  Hydrolytic


degradation

& Yong Wang


yongwang1976@163.com
1
Key Laboratory of Advanced Technologies of Materials (Ministry of Education), School of
Materials Science and Engineering, Southwest Jiaotong University, Chengdu 610031, China
2
State Key Laboratory of Oral Diseases, Sichuan University, Chengdu, 610041, China

123
Polym. Bull.

Introduction

Poly(lactic acid) (PLA) has attracted much attention of researchers in both the
academic and industrial fields since it has been synthesized. There are at least two
aspects that arouse the research interests. Firstly, PLA can be prepared from
completely natural resources such as corn, wheat and/or rice and consequently, the
wide application of PLA is favorable for reducing the degree of dependence on
fossil resources. Secondly, PLA can be completely degraded. For example, through
hydrolytic degradation process, the molecular weight of PLA can be greatly reduced
and finally, the oligomer or lactide monomer can be obtained again [1]. This means
that the articles made of complete or partial PLA material are recyclable to a certain
degree. However, through enzymolysis process, PLA can be completely decom-
posed into water and carbon dioxide [2–4]. Obviously, this greatly reduces the
problem of environmental pollution and meets the requirement of sustainable
development. Therefore, PLA is also regarded as a ‘‘green polymer’’ and it is of
great interest as an alternative to traditional commodity plastics.
Studying the hydrolytic degradation behavior of PLA and/or PLA-based
materials is very significant, and the corresponding researches have been carried
out by many researchers in the past years. So far, it has already been clarified that
the hydrolytic degradation of PLA is influenced by many factors, including the
microstructural factors such as the content of D-isomer content [5, 6], the molecular
weight and its distribution [6], the degree of crystallinity [7–14], the spherulite
morphology [15] and the end group of PLA [16, 17], etc., and environmental factors
such as temperature [15, 18, 19], pH value of solution [13, 18, 20, 21] and the
presence of ultraviolet light [22], etc.
PLA composites, as one type of PLA-based materials, have been one of the main
subjects relating to the researches about PLA. To date, the researches about PLA
composites are mainly focused on several aspects, including accelerating the
crystallization of PLA matrix [23–29], enhancing the mechanical properties [29–
35], and improving the thermal stability [36–40], etc. In addition, much attention
has also been focused on studying the hydrolytic degradation behavior of PLA
composites. Different kinds of nanofillers, including zero-dimensional nanoparticles
such as titanium dioxide (TiO2) [41], zinc oxide (ZnO) [42], silica (SiO2) [43], and
calcium carbonate (CaCO3) [44], one-dimensional fibers (nanotubes) or whiskers
such as carbon nanotubes (CNTs) [26, 27, 45, 46] and cellulose whisker [47], and
two-dimensional nanoparticles such as clay [8, 19, 48–50] and layered double
hydroxides [51], etc. The hydrolytic degradation of PLA composites is also
influenced by many factors. Besides the influence originated from the variation of
PLA microstructure induced by nanofillers, the other factors, including the
dispersion and content of nanofillers [42, 51] and the variation of hydrophilicity
of composites [43, 46, 50], also determine the hydrolytic degradation rate of PLA
matrix. Specifically, it has been clarified that with the presence of nanofillers, a large
number of interfaces are introduced into the material and consequently, hydrolytic
degradation of PLA composites usually occurs at the interface between PLA and
nanofillers rather than in the PLA bulk [41, 43].

123
Polym. Bull.

The other type of PLA-based materials is PLA blends, which are prepared
through compounding with other polymers. PLA blends can be classified at least
two types, namely, PLA acts as the matrix while other polymers act as the dispersed
component, or PLA acts as the dispersed component while other polymers act as the
matrix. For the first type of PLA blends, it is expected that the addition of other
polymers can improve the mechanical, barrier and thermal properties of PLA [52–
56]. However, for the second type of PLA blends, it is expected that PLA can
strengthen and toughen other polymers because of the excellent intrinsic strength
and modulus of PLA on one hand. On the other hand, the addition of PLA can
reduce the consumption of petroleum-based plastics and simultaneously enhance the
recyclability of the materials [57]. However, researches about the hydrolytic
degradation behavior of PLA blends are rather few.
Oyama et al. [58] investigated the effect of oligomeric poly(aspartic acid-co-
lactide) (PAL) on the hydrolytic degradation of poly(L-lactic acid) (PLLA) and
found that the addition of PAL did not accelerate the hydrolysis of PLLA in air
(25 °C, 60 % relative humidity), but significantly accelerated it in a phosphate
buffer solution. Shirahase et al. [59] studied the hydrolytic degradation of PLLA
blends containing poly(p-vinyl phenol) (PVPh) and reported that in an alkaline
solution, degradation rate of PLLA/PVPh blends was faster than that of neat PLLA
because PVPh could dissolve in the alkaline solution. Tsuji et al. [60] reported that
the presence of hydrophilic poly(vinyl alcohol) (PVA) accelerated the non-
enzymatic and enzymatic hydrolysis of PLLA due to the increased wettability and
the simultaneous occurrence of enzymatic hydrolysis at the interfaces of PLLA-rich
and PVA-rich phases inside the blend films as well as at the film surfaces,
respectively. Huang et al. [61] comparatively investigated the hydrolytic degrada-
tion of PLA coated with poly(dodecafluorheptyl methylacrylate) (PFA) and PLA
blended with PFA. The results showed that the blended PLA sample exhibited
higher water absorption rate compared with the coated PLA sample because the
gaps between PLA and PFA, resulting from poor interaction between the two
components, reduced the penetration depth. However, they found that coating and
blending PFA with PLA delayed the hydrolytic degradation of PLA through
hindering water permeation and decreasing the autocatalysis of the hydrolytic
products.
To further understand the hydrolytic degradation behaviors of PLA blends, in the
present work, poly(butylene succinate) (PBS), which exhibits a relatively higher
hydrophilicity than PLA, is introduced into PLA. Factually, PLA/PBS blends have
been researched elsewhere and the attention of researchers is mainly focused on
studying the crystallization of both components and the mechanical property
changes of the blends [62, 63]. It is expected that similar to the addition of
hydrophilic nanofillers, which provide a large number of interfaces in the PLA
material and accelerate the hydrolytic degradation of PLA matrix, the addition of
immiscible PBS component also provides the weak interface in the blends,
providing the permeable channels in the materials for water penetration. Therefore,
we believe that the hydrolytic degradation of PLA might be accelerated by adding
PBS.

123
Polym. Bull.

Experimental part

Materials

PLA (2003D, with a D-isomer content of 4.3 %, a weight-average molecular weight


(M w ) of 2.53 9 105 g/mol, a polydispersity index (d) of 1.94, a melt flow rate
(MFR) of 5.6 g/10 min (190 °C/2.16 kg) and a density of 1.24 g/cm3) was
purchased from NatureWorksÒ, USA. PBS [HX-E101, with the density of 1.26 g/
cm3 and the MFR of 20 g/min (190 °C/2.16 kg)] was obtained from Anqing Hexing
Chemical Co., Ltd.

Sample preparation

Before melt compounding, PLA and PBS pellets were dried at 40 °C for 24 h to
avoid the effect of moisture on processing. Different contents of PBS (10, 20, 30 and
40 wt%) were introduced into PLA through melt compounding, which was
conducted on an internal mixer. During the compounding processing, the processing
conditions were set at a melt temperature of 190 °C, a screw speed of 60 rpm and a
mixing duration of 6 min. After that, the blends were compression-molded at a melt
temperature of 190 °C and a pressure of 5 MPa to obtain the sheet-like samples for
hydrolytic degradation measurements. The sheet-like sample had a thickness of
about 0.05 cm. The sample notation was defined as PLA/PBS-x, where x represented
the content of PBS. For a comparison, pure PLA sample was also prepared through
the completely same processing procedures as was done for the blends.

Differential scanning calorimetry (DSC)

A differential scanning calorimetry (DSC) STA 449C Jupiter (Netzsch, Germany)


was used to investigate the melting behaviors of samples. A sample of about 8 mg
was heated from 30 to 200 °C at a heating rate of 10 °C/min. All the measurements
were conducted in nitrogen atmosphere.

Wide-angle X-ray diffraction (WAXD)

The crystalline structures of samples obtained before and after hydrolytic


degradation were investigated using a wide-angle X-ray diffraction (WAXD)
X’pert PRO diffractometer (Panalytical, the Netherlands) with Ni-filtered Cu Ka
radiation. The continuous scanning angle range used in this study was from 5° to
55° at 35 kV and 25 mA.

Scanning electron microscopy (SEM)

The phase morphologies of the blends as obtained and the morphological evolution
of samples during the hydrolytic degradation process were characterized using a
scanning electron microscope (SEM) Fei inspect (FEI, the Netherlands) at an

123
Polym. Bull.

accelerating voltage of 5.0 kV. To characterize the phase morphology, the


compression-molded samples were firstly immersed into liquid nitrogen for
30 min, and then they were cryogenically fractured. Either for the cryo-fractured
surfaces or for the surfaces of the hydrolyzed samples, they were coated with a thin
layer of gold before SEM characterization.

Contact angle measurement

Contact angle measurement was carried out on the surfaces of the compression-
molded sheets. The measurements were conducted on a drop shape analysis system
DSA 100 (KRÜSS, Germany) at an environmental temperature of 20 °C.
Measurement of a given contact angle was carried out for at least 5 times and the
value that was most close to an average value was given. Double distilled water
(H2O) was used as a probe liquid.

Hydrolytic degradation measurement

Hydrolytic degradation of the samples were carried out at 37 ± 0.2 °C in a solution


of sodium hydroxide (NaOH, pH = 13). For all the samples, the dimensions were
invariant (with a width of 4 cm, a length of 4 cm and a thickness of 0.05 cm). The
weight of each sample was carefully weighed before degradation. Then, the sample
was immersed into the solution and the hydrolytic degradation occurred. After being
hydrolyzed for 24 h, the sample was taken out, washed with fresh water and dried
through a cryodesiccation treatment at a temperature of -18 °C for 24 h to insure
the complete removal of water, and then the residual weight of the hydrolyzed
sample was carefully weighed to know the weight loss of the sample (including the
hydrolytic degradation of PLLA matrix and the possibly released PBS particles
from the material surface for blend samples) during the hydrolytic degradation
process. After that, the hydrolyzed sample was further hydrolyzed in the alkaline
solution, and the similar procedures were repeated until the appearance of sample
could not be well maintained. Subsequently, the variation of the weight loss per unit
area (U, g/cm2) as a function of hydrolytic degradation time (t, h) was illustrated.
Here, the weight loss per unit area was calculated according to the following
equation:
w0  wt
U¼ ; ð1Þ
A
where w0 and wt represented the initial weight (g) of sample before hydrolysis and
residual weight (g) after being degraded for a certain time, t, respectively, and A
represented the area of sample. For each sample, the hydrolytic degradation mea-
surement was repeated three times and the data that were mostly close to the
average value were reported.

123
Polym. Bull.

Results and discussion

Microstructure and morphology of samples as prepared

Studying the microstructures and morphologies of samples as prepared is crucial to


understanding the subsequent hydrolytic degradation behaviors. Here, the effects of
PBS on microstructure of PLA matrix and the morphologies of the blends were
firstly investigated. Figure 1 shows the DSC heating curves of pure PLA and PLA/
PBS blends. For pure PLA sample, it exhibits several transitions, attributed to the
glass transition (TgPLA ) at temperature about 60.2 °C, the cold crystallization
(TccPLA ) at about 114.3 °C, and the fusion of crystalline structure (TmPLA ) at about
150.0 °C, respectively. Interestingly, the presence of PBS induces the apparent
changes of the DSC heating curves. Firstly, Tg of PLA matrix in the blends is
slightly lower than that of pure PLA. Secondly, PLA matrix exhibits double
endothermic peaks (Tm1 and Tm2 as shown in the graph). Furthermore, it can be seen
that the first endothermic peak (Tm1 ) shifts to lower temperatures and the intensity
gradually decreases with the increase of PBS content, while the second endothermic
peak (Tm2 ) is nearly invariant. Thirdly, although the fusion of PBS crystallites
(TmPBS , 109.9 °C) occurs in the same temperature ranges in which the cold
crystallization of PLA matrix occurs simultaneously, it is found that the
endothermic phenomenon of PBS component is more apparent than the exothermic
phenomenon of PLA matrix during the DSC heating scan. Even though the content
of PBS is only 10 wt%, one can clearly observe the fusion of PBS crystallites. With
the increase of PBS content, the intensity of PBS endothermic peak dramatically
increases and consequently, the exothermic peak of PLA in the DSC heating curves
is covered. Furthermore, it is worth noting that TmPBS increases with the increase of
PBS content. In addition, it is interesting to observe that a weak exothermic peak is

Tm-PBS
Tg-PLA Tm2
Tm1
PLA/PBS-40
Endothermic

PLA/PBS-30

PLA/PBS-20

PLA/PBS-10

Tm-PLA PLA

Tcc-PLA

60 80 100 120 140 160 180


o
Temperature / C

Fig. 1 DSC heating curves showing the melting behaviors of pure PLA and PLA/PBS blends as obtained

123
Polym. Bull.

present at the left side of the PBS endothermic peak. This indicates that the cold
crystallization of PLA is possibly induced at relatively lower temperature because of
the heterogeneous nucleation effect of PBS crystallites. It has been reported by
Shibata et al. [63] that the addition of a few PBS could promote the isothermal and
non-isothermal crystallization of PLA. However, because the occurrence of cold
crystallization is induced at relatively lower temperature and the presence of many
PBS crystallites in the material hinders the lamellar growth of PLA crystallites, PLA
component exhibits the slightly decreased Tm1 with the increase of PBS content.
Although DSC measurements can clearly demonstrate the crystallization of PBS
component in the blends, it cannot reflect the crystalline structure of PLA matrix in
the samples as obtained. To further clarify the doubt, WAXD measurements were
carried out. Figure 2 exhibits the WAXD profiles of pure PLA and PLA/PBS
blends. It can be seen that pure PLA sample exhibits the typical WAXD profiles of
amorphous polymers. For PLA/PBS blends, several characteristic diffraction peaks
are observed at 2h = 19.7°, 22.7° and 29.0°, attributing to the diffractions of (020),
(110) and (111) crystal planes of a-form PBS [64]. In addition, the intensities of
these diffractions increase with the increase of PBS content in the blends. This
confirms with the results obtained from DSC measurements. However, the
characteristic diffraction peaks of PLA, which are usually observed at
2h = 14.8°, 16.5° and 18.9°, attributing to the diffractions of (010), (110)/(200)
and (203) crystal planes [65, 66], cannot be differentiated from the WAXD profiles
as shown in Fig. 2. This indicates that in all the samples investigated in this work,
PLA is completely amorphous. Therefore, from a crystallinity point of view, it is
expected that adding PBS does not reduce the hydrolytic degradability of PLA
because PBS neither induces the crystallization nor enhances the crystallinity of
PLA matrix during the common melt compounding processing.

(110)PBS
o
22.7
(020)PBS
o
19.7
Intensity (a.u.)

o
29.0
PLA/PBS-40
PLA/PBS-30
PLA/PBS-20
PLA/PBS-10
PLA

5 10 15 20 25 30 35 40
2θ ( )
o

Fig. 2 WAXD profiles showing the crystalline structures of pure PLA and PLA/PBS blends

123
Polym. Bull.

The morphologies of PLA/PBS blends were characterized using SEM. As shown


in Fig. 3, all the blends exhibit the typical sea-island structures. In addition, there
are at least two features that need to be noticed. Firstly, there are clear interfaces
between PLA matrix and dispersed PBS particles. This conforms with the report in
literature that the blend is immiscible [63, 67]. Secondly, the blends containing
relatively low PBS contents (Fig. 3a, b) exhibit more apparent interfacial debonding
and some PBS particles are completely removed during the cryo-fracture process.
Thirdly, PBS component exhibits heterogeneous distribution in the PLA matrix,
especially in the blends containing relatively high PBS contents (Fig. 3c, d), and the
particle diameter varies from 0.2 to 6.5 lm. Obviously, the presence of the gaps
between PLA matrix and dispersed PBS particles provides permeation channels for
water penetration, which facilitates the occurrence of hydrolytic degradation.

Hydrophilicity of blends

A prerequisite for the hydrolytic degradation of samples in alkaline solution is that


samples have good ability to absorb water. Namely, samples exhibit excellent

Fig. 3 SEM images showing the morphologies of PLA/PBS blends with different contents of PBS. The
content of PBS is a 10 wt%, b 20 wt%, c 30 wt% and d 40 wt%

123
Polym. Bull.

hydrophilicity. It is well known to all that PBS is a typical polar polymer with many
functional groups in molecular chains. Therefore, it is easily to ask whether the
hydrophilicity of PLA material can be enhanced by adding PBS. Here, the
hydrophilicity of samples was evaluated by measuring contact angles of distilled
water on sample surfaces. The profiles of water droplets on the surfaces of pure PLA
and PLA/PBS blends are shown in Fig. 4. One can see that PLA exhibits a contact
angle of 72°, indicating the relatively low hydrophilicity of pure PLA sample. PLA/
PBS blends exhibit relatively lower contact angles compared with pure PLA sample.
For example, the PLA/PBS-10 sample exhibits a contact angle of 65°. Increasing the
content of PBS results in the further decrease of contact angle. This demonstrates
the gradually enhanced hydrophilicity of sample surface with the increase of PBS
content. The variation of contact angle can be explained as follows. Although PBS
is a dispersed component in the blends and the particle size is much smaller than the
drop size of water during the contact angle measurement, it increases water
concentration around PBS component and water supply rate to PLA [60]. In
addition, the gaps between PLA matrix and dispersed PBS particles also increase
the wettability of the blends. Consequently, contact angle decreases gradually with
increasing PBS content. However, it should be pointed out that the variation of

Fig. 4 The profiles of water droplets on the surfaces of pure PLA and PLA/PBS blends. The
corresponding contact angles are also shown in the images

123
Polym. Bull.

contact angle becomes smaller at relatively high PBS contents. Anyway, from a
hydrophilicity point of view, it is expected that PLA/PBS blends exhibit more
probability to be hydrolytically degraded in alkaline solution.

Hydrolytic degradation behavior

The above results show that in the present work, adding PBS neither induces the
crystallization nor enhances the crystallinity of PLA matrix and contrarily, the
presence of PBS not only induces a large number of permeation channels for water
penetration but also enhances the wettability of sample surface. This implies that
adding PBS will accelerate the hydrolytic degradation of PLA materials. Here, the
hydrolytic degradation behaviors were evaluated by measuring the residual weights
of samples. Figure 5 shows the variation of weight loss per unit area (U) versus the
hydrolytic degradation time (t). From Fig. 5 one can see that the variation of U is
greatly dependent upon the content of PBS in the samples. For pure PLA, U
increases apparently at relatively shorter t and changes slightly at longer t,
especially at degradation time longer than 216 h. The similar variation trends are
also observed for PLA/PBS blends containing relatively lower PBS content (10–
30 wt%). However, it is worth noting that adding PBS into the sample induces more
apparent changes of U with increasing t. Specifically, for the PLA/PBS-40 sample,
it exhibits the most apparent changes in U with increasing t until t is increased up to
216 h. This clearly indicates that PLA/PBS blends exhibit largely accelerated
hydrolytic degradation. Accordingly, the whole hydrolytic degradation process can
be roughly classified as three stages: (I) initial stage, in which U dramatically
increases with increasing t and the slopes of the curves are relatively big; (II) middle
stage, in which U slightly increases with further increasing t, and the slopes of the
curves become smaller compared with the initial stage; (III) later stage, in which the
variation of U is smaller than that in the middle stage. For the PLA/PBS-40 sample,
it can be thought that the sample exhibits only two stages during the whole

PLA
0.05 PLA/PBS-10
Weight loss per unit area (g/cm )

PLA/PBS-20
2

PLA/PBS-30
0.04 PLA/PBS-40

0.03

0.02
RIII

0.01 RII

0.00 RI
0 48 96 144 192 240 288 336
Time/h

Fig. 5 Variation of weight loss per unit area of different samples versus the hydrolytic degradation time

123
Polym. Bull.

hydrolytic degradation process in the present work. It can be seen that in the middle
stage, increasing the content of PBS results in more apparent change of U at the
same hydrolytic degradation time, which further demonstrates that adding PBS
accelerates the hydrolytic degradation of PLA materials. Obviously, it can be
deduced that the increased interfacial gaps and the enhanced hydrophilicity of
sample surface mainly contribute to the accelerated hydrolytic degradation of
sample.
Studying surface morphologies of degraded samples is favorable for further
understanding the variation of hydrolytic degradation behaviors. Here, the samples
that were obtained after hydrolysis for 312 h were characterized. Figure 6 shows the
surface morphologies of the representative PLA/PBS-10 and PLA/PBS-30 samples.
For the PLA/PBS-10 sample (Fig. 6a), it can be seen that after being degraded,
some dispersed PBS particles are completely exposed on the sample surface on one
hand. On the other hand, for the PBS particles that still encapsulated by PLA matrix,
the gaps between PBS particles and PLA matrix become more apparent as compared
with that observed in the samples before hydrolytic degradation (Fig. 3). This
indicates that the hydrolytic degradation of PLA matrix occurs firstly in the
interface regions between the two components. More serious degradation of PLA
matrix can be observed for the PLA/PBS-30 sample (Fig. 6b), demonstrating that
adding PBS accelerates the hydrolytic degradation of PLA. It is well known to all
that PBS is also a degradable polymer and it exhibits the hydrolytic degradation in
the alkaline solution [68]. However, from the surface morphologies as shown in
Fig. 6, the presence of a large number of PBS particles on the surfaces of the
degraded samples implies that in PLA/PBS blends, the hydrolytic degradation rate
of PBS particles is possibly lower than that of PLA matrix.
Furthermore, the surface morphologies in the cross-section direction are also
characterized. As shown in Fig. 7, one can clearly see that the hydrolytic
degradation of PLA/PBS blends mainly occurs on the surface of sample, and with
increasing of hydrolytic degradation time, the inner part of the sample begins

Fig. 6 SEM images showing the surface morphologies of the a PLA/PBS-10 and b PLA/PBS-30
samples after being degraded at 37 °C for 312 h

123
Polym. Bull.

Fig. 7 SEM images showing the surface morphologies in the cross-section direction of the PLA/PBS-30
sample after being degraded at 37 °C for 312 h. Images were obtained at low (a) and high magnifications
(b)

degradation. In other words, the erosion of PLA/PBS blends in alkaline solution


occurs in a similar way as is observed for pure PLA, namely, the blends still obey
the surface-erosion mechanism [13, 20]. This indicates that the presence of PBS
particles does not change the hydrolytic degradation mechanism of PLA matrix.
To well understand the hydrolytic degradation behavior of PLA/PBS blends,
more visualized illustrations are shown in Fig. 8. For a comparison, the hydrolytic
degradation of pure PLA is also shown. For pure PLA, the hydrolytic degradation in
alkaline solution proceeds mainly via a surface-erosion mechanism [13, 20], and the
sample size and shape gradually change with increasing hydrolytic degradation time
(from the initial stage to the later stage). For PLA/PBS blends, although the
presence of PBS component enhances the wettability of sample surface, the sample
still obeys the surface-erosion mechanism and in the initial stage, PLA/PBS sample
still exhibits the similar degree of hydrolytic degradation to that of pure PLA
possibly due to that few PBS particles are exposed on sample surface. That is also

Fig. 8 Schematic representations showing the hydrolytic degradation process of pure PLA and PLA/PBS
blends

123
Polym. Bull.

the reason why all PLA/PBS blends exhibit the similar variations of weight loss per
unit area to that of pure PLA in the initial stage. However, once the sample surface
is degraded and more PBS particles are exposed, water has more chance to penetrate
into the gaps between PBS particles and PLA matrix and consequently, the
hydrolytic degradation of PLA matrix is greatly accelerated. Obviously, the higher
the content of PBS in the sample is, the bigger the hydrolytic degradation rate is.

Microstructure evolution during hydrolytic degradation

Many researches have already demonstrated that during the hydrolytic degradation
process, PLA also exhibits microstructure rearrangement, relating to the occurrence
of the molecular ordering and crystallization. Tsuji et al. [69] investigated the
crystalline structure evolution of PLA using small-angle X-ray scattering and firstly
reported the occurrence of growth of PLA crystalline regions during hydrolysis in
phosphate-buffered solution. Subsequently, they comparatively investigated the
variations of crystalline structures of amorphous PLA and crystalline PLA samples
during the hydrolysis. They found that a large number of crystalline structures were
induced for the amorphous PLA sample [22]. In contrast, not growth but significant
erosion of the crystalline regions of PLA occurred in the alkaline solution. However,
in our previous work, it was observed that even in the alkaline solution, a large
number of crystallites could be induced in the PLA samples containing a few CNTs
[46]. The enhanced chain segment mobility and the heterogeneous nucleation effect
of CNTs were suggested the main reasons for the occurrence of PLA crystallization
during the hydrolytic degradation process. In the present work, the crystalline
structures of hydrolyzed samples were also characterized using WAXD. As shown
in Fig. 9, which exhibits the WAXD profiles of pure PLA and the representative

(110)PBS
(020)PBS

PLA/PBS-30 After hydrolysis


Intensity (a.u.)

Before hydrolysis

PLA/PBS-30

PLA After hydrolysis

PLA Before hydrolysis

5 10 15 20 25 30 35
2θ ( )
o

Fig. 9 WAXD profiles showing the crystalline structures of pure PLA and the PLA/PBS-30 samples
before and after being degraded at 37 °C for 216 h

123
Polym. Bull.

PLA/PBS-30 samples before and after being degraded for 216 h. One can see that
both pure PLA and the blend exhibit very similar WAXD profiles before and after
hydrolytic degradation. There is no evidence to demonstrate the occurrence of PLA
crystallization possibly due to the fact that the hydrolytic degradation occurs at
relatively low temperature (37 ± 0.2 °C) and in this condition, the motion of PLA
chain segment is greatly restricted on one hand. On the other hand, the hydrolytic
degradation time is possibly shorter compared with that used in the literature and
there is not enough time for the crystallization of PLA in this work.

Conclusions

In summary, PLA/PBS blends containing different PBS contents have been


prepared. The presence of dispersed PBS particles neither induces the crystallization
nor enhances the crystallinity of PLA matrix when the samples are prepared through
the common melt compounding processing. However, the presence of immiscible
PBS component not only induces a large number of gaps in the materials, but also
enhances the hydrophilicity of the samples and consequently, accelerated hydrolytic
degradation is achieved for PLA/PBS blends. The higher the content of PBS is, the
bigger the weight loss per unit area of sample is. It is demonstrated that the
hydrolytic degradation of PLA/PBS blends in the alkaline solution obeys the
surface-erosion mechanism and the degradation occurs mainly in the interface
regions between PLA and PBS. Furthermore, it is found that no crystallization
occurs during the hydrolytic degradation process.

Acknowledgments Authors express their sincere thanks to the National Natural Science Foundation of
China (51473137) for financial support.

References
1. Hoglund A, Hakkarainen M, Edlund U, Albertsson AC (2009) Surface modification changes the
degradation process and degradation product pattern of polylactide. Langmuir 26:378–383
2. Tsuji H, Tezuka Y, Yamada K (2005) Alkaline and enzymatic degradation of L-lactide copolymers.
II. Crystallized films of poly(L-lactide-co-D-lactide) and poly(L-lactide) with similar crystallinities.
J Polym Sci Part B Polym Phys 43:1064–1075
3. Li SM, Tenon M, Garreau H, Braud C, Vert M (2000) Enzymatic degradation of stereocopolymers
derived from L-, DL- and meso-lactides. Polym Degrad Stab 67:85–90
4. Liu L, Li SM, Garreau H, Vert M (2000) Selective enzymatic degradations of Poly(L-lactide) and
poly(e-caprolactone) blend films. Biomacromolecules 1:350–359
5. Saha SK, Tsuji H (2006) Effects of molecular weight and small amounts of D-lactide units on
hydrolytic degradation of poly(L-lactic acid)s. Polym Degrad Stab 91:1665–1673
6. Ray SS, Bousmina M (2005) Biodegradable polymers and their layered silicate nanocomposites: in
greening the 21st century materials world. Prog Mater Sci 50:962–1079
7. Fukuzaki H, Yoshida M, Asano M, Kumakura M (1989) Synthesis of copoly(D,L-lactic acid) with
relative low molecular weight and in vitro degradation. Eur Polym J 25:1019–1026
8. Zhou Q, Xanthos M (2008) Nanoclay and crystallinity effects on the hydrolytic degradation of
polylactides. Polym Degrad Stab 93:1450–1459
9. Gorrasi G, Pantani R (2013) Effect of PLA grades and morphologies on hydrolytic degradation at
composting temperature: assessment of structural modification and kinetic parameters. Polym Degrad
Stab 98:1006–1014

123
Polym. Bull.

10. Pantani R, Sorrentino A (2013) Influence of crystallinity on the biodegradation rate of injection-
moulded poly(lactic acid) samples in controlled composting conditions. Polym Degrad Stab
98:1089–1096
11. Li SM, Garreau H, Vert M (1990) Structure-property relationships in the case of the degradation of
massive poly(a-hydroxy acid) in aqueous media Part 3 influence of the morphology of poly(L-lactic
acid). J Mater Sci Mater Med 1:198–206
12. Vert M, Li SM, Garreau H (1994) Attempts to map the structure and degradation characteristics of
aliphatic polyesters derived from lactic and glycolic acid. J Biomater Sci Polym Edn 6:639–649
13. Tsuji H, Ikada Y (1998) Properties and morphology of poly(L-lactide). II. Hydrolysis in alkaline
solution. J Polym Sci Part A Polym Chem 36:59–66
14. Tsuji H, Mizuno A, Ikada Y (2000) Properties and morphology of poly(L-lactide). III. Effects of
initial crystallininty on long-term in vitro hydrolysis of high molecular weight poly(L-lactide) film in
phosphate-buffered solution. J Appl Polym Sci 77:1452–1464
15. Tsuji H, Nakahara K, Ikarashi K (2001) Poly(L-lactide), 8—high temperature hydrolysis of poly(L-
lactide) films with different crystallinities and crystalline thicknesses in phosphate-buffered solution.
Macromol Mater Eng 286:398–406
16. Andersson SR, Hakkarainen M, Inkinen S (2010) Customizing the hydrolytic degradation rate of
stereocomplex PLA through different PDLA architectures. Biomacromolecules 13:1212–1222
17. Chung S (1995) Chain-end scission in acid catalyzed hydrolysis polylactide in solution. J Control
Release 34:9–15
18. Xu L, Crawford K, Gorman C (2011) Effects of temperature and pH on the degradation of poly(lactic
acid) brushes. Macromolecules 44:4777–4782
19. Fukushima K, Tabuani D, Dottori M, Armentano I, Kenny JM, Gamino G (2011) Effect of tem-
perature and nanoparticle type on hydrolytic degradation of poly(lactic acid) nano- composites.
Polym Degrad Stab 96:2120–2129
20. Tsuji H, Nakahara K (2002) Poly(L-lactide). IX. Hydrolysis in acid media. J Appl Polym Sci
86:186–194
21. Tsuji H, Ikarashi K (2004) In vitro hydrolysis of poly(L-lactide) crystalline residues as extended-
chain crystallites. III. Effect of pH and enzyme. Polym Degrad Stab 85:647–656
22. Tsuji H, Shimizu K, Sato Y (2012) Hydrolytic degradation of poly(L-lactic acid): combined effects of
UV Treatment and crystallization. J Appl Polym Sci 125:2394–2406
23. Picard E, Espuche E, Fulchiron R (2011) Effect of an organo-modified montmorillonite on PLA
crystallization and gas barrier properties. Appl Clay Sci 53:58–65
24. Shieh YT, Twu YK, Su CC, Lin RH, Liu GL (2010) Crystallization Kinetics Study of Poly(L-lactic
acid)/Carbon Nanotubes Nanocomposites. J Polym Sci Part B: Polym Phys 48:983–989
25. Shieh YT, Liu GL, Twu YK, Wang TL, Yang CH (2010) Effects of carbon nanotubes on dynamic
mechanical property, thermal property, and crystal structure of poly(L-lactic acid). J Polym Sci Part B
Polym Phys 48:145–152
26. Zhao YY, Qiu ZB, Yang WT (2008) Effect of functionalization of multiwalled nanotubes on the
crystallization and hydrolytic degradation of biodegradable poly(L-lactide). J Phys Chem B
112:16461–16468
27. Zhao YY, Qiu ZB, Yang WT (2009) Effect of multi-walled carbon nanotubes on the crystallization
and hydrolytic degradation of biodegradable poly(L-lactide). Compos Sci Technol 69:627–632
28. Xu HS, Dai XJJ, Lamb PR, Li ZM (2009) Poly(L-lactide) crystallization induced by multiwall carbon
nanotubes at very low loading. J Polym Sci Part B Polym Phys 47:2341–2352
29. Suksut B, Deeprasertkul C (2011) Effect of nucleating agent on physical properties of poly(lactic
acid) and its blend with natural rubber. J Polym Environ 19:288–296
30. Jiang L, Zhang JW, Wolcott MP (2007) Comparison of polylactide/nano-sized calcium carbonate and
polylactide/montmorillonite composites: reinforcing effects and toughening mechanisms. Polymer
48:7632–7644
31. Wu DF, Wu L, Zhou WD, Sun YR, Zhang M (2010) Relations between the aspect ratio of carbon
nanotubes and the formation of percolation networks in biodegradable polylactide/carbon nanotube
composites. J Polym Sci Part B Polym Phys 48:479–489
32. Yoon JT, Jeong YG, Lee SC, Min BG (2009) Influences of poly(lactic acid)-grafted carbon nanotube
on thermal, mechanical and electrical properties of poly(lactic acid). Polym Adv Technol
20:20631–20638
33. Ramontja J, Ray SS, Pillai SK, Luyt AS (2009) High-performance carbon nanotube-reinforced
bioplastic. Macromol Mater Eng 294:839–846

123
Polym. Bull.

34. Luo YB, Li WD, Wang XL, Xu DY, Wang YZ (2009) Preparation and properties of nanocomposites
based on poly(lactic acid) and functionalized TiO2. Acta Mater 57:3182–3191
35. Huang JW, Hung YC, Wen YL, Kang CC, Yeh MY (2009) Polylactide/nano- and micro-scale silica
composite films. I. Preparation and characterization. J Appl Polym Sci 112:1688–1694
36. Kim HS, Chae YS II, Kwon H, Yoon JS (2009) Thermal degradation behavior of multi-walled carbon
nanotube-reinforced poly(L-lactide) nanocomposites. Polym Int 58:826–831
37. Huang JW, Hung YC, Wen YL, Kang CC, Yeh MY (2009) Polylactide/nano- and micro-scale silica
composite films. II. Melting behavior and cold crystallization. J Appl Polym Sci 112:3149–3156
38. Huang TC, Yeh JM, Yang JC (2010) Effect of silica size on the thermal, mechanical and
biodegradable properties of polylactide/silica composite material prepared by melt blending. Adv
Mater Res 123–125:1215–1218
39. Yan SF, Yin JB, Yang JY, Chen XS (2007) Structural characteristics and thermal properties of
plasticized poly(L-lactide)-silica nanocomposites synthesized by sol-gel method. Mater Lett
61:2683–2686
40. Zhang J, Lou JZ, Ilias S, Krishnamachari P, Yan JZ (2008) Thermal properties of poly(lactic
acid)/fumed silica nanocomposites: experiments and molecular dynamics simulations. Polymer
49:2381–2386
41. Luo YB, Wang XL, Wang YZ (2012) Effect of TiO2 nanoparticles on the long-term hydrolytic
degradation behavior of PLA. Polym Degrad Stab 97:721–728
42. Qu M, Tu HL, Amarante M, Song YQ, Zhu SS (2014) Zinc oxide nanoparticles catalyze rapid
hydrolysis of poly(lactic acid) at low temperatures. J Appl Polym Sci 131:40287–40293
43. Chen HM, Wang YP, Chen J, Yang JH, Zhang N, Huang T, Wang Y (2013) Hydrolytic degradation
behavior of poly(L-lactide)/SiO2 composites. Polym Degrad Stab 98:2672–2679
44. Flahiff CM, Blackwell AS, Hollis JM, Feldman DS (1996) Analysis of a biodegradable composite for
bone healing. J Biomed Mater Res 32:419–424
45. He LH, Sun J, Wang XX, Fan XH, Zhao QL, Cai LF, Song R, Ma Z, Huang W (2012) Unzipped
multiwalled carbon nanotubes-incorporated poly(L-lactide) nanocomposites with enhanced interface
and hydrolytic degradation. Mater Chem Phys 134:1059–1066
46. Chen HM, Feng CX, Zhang WB, Yang JH, Huang T, Zhang N, Wang Y (2013) Hydrolytic degra-
dation behavior of poly(L-lactide)/carbon nanotubes nanocomposites. Polym Degrad Stab
98:198–208
47. de Paula EL, Mano V, Pereira FV (2011) Influence of cellulose nanowhiskers on the hydrolytic
degradation behavior of poly(D, L-lactide). Polym Degrad Stab 96:1631–1638
48. Paul MA, Delcourt C, Alexandre M, Degée Ph, Monteverde F, Dubois Ph (2005) Polylac-
tide/montmorillonite nanocomposites: study of the hydrolytic degradation. Polym Degrad Stab
87:535–542
49. Roy PK, Hakkarainen M, Albertsson AC (2010) Nanoclay effects on the degradation process and
product patterns of polylactide. Polym Degrad Stab 97:1254–1260
50. Chen HM, Chen JW, Chen J, Yang JH, Huang T, Zhang N, Wang Y (2012) Effect of organic
montmorillonite on cold crystallization and hydrolytic degradation of poly(L-lactide). Polym Degrad
Stab 97:2273–2283
51. Eili M, Shameli K, Ibrahim NA, Yunus WMZW (2012) Degradability enhancement of poly(lactic
acid) by stearate-Zn3Al LDH Nanolayers. Int J Mol Sci 13:7938–7951
52. Shi YY, Du XC, Yang JH, Huang T, Zhang N, Wang Y, Yuan GP, Zhang CL (2014) Super
toughened poly(L-lactide)/thermoplastic polyurethane blends achieved by adding dicumyl peroxide.
Polym Plast Technol Eng 53:1344–1353
53. Li YJ, Shimizu H (2009) Improvement in toughness of poly(L-lactide) (PLLA) through reactive
blending with acrylonitrile-butadiene-styrene (ABS): morphology and properties. Eur Polym J
42:738–746
54. Kumar M, Mohanty S, Nayak SK, Parvaiz MR (2010) Effect of glycidyl methacrylate (GMA) on the
thermal, mechanical and morphological property of biodegradable PLA/PBAT blend and its
nanocomposites. Bioresour Technol 101:8406–8415
55. Al-ltry Lamnawar K, Maazouz A (2012) Improvement of thermal stability, rheological and
mechanical properties of PLA, PBAT and their blends by reactive extrusion with functionalized
epoxy. Polym Degrad Stab 97:1898–1914
56. Boufarguine M, Guinault A, Miquelard-Garnier G, Sollogoub C (2013) PLA/PHBV films with
improved mechanical and gas barrier properties. Macromol Mater Eng 98:1065–1073

123
Polym. Bull.

57. Nameroff TJ, Garant RJ, Albert MB (2004) Adoption of green chemistry: an analysis based on US
patents. Res Policy 33:959–974
58. Oyama HT, Tanaka Y, Kadosaka A (2009) Rapid controlled hydrolytic degradation of poly(L-lactic
acid) by blending with poly(aspartic acid-co-L-lactide). Polym Degrad Stab 94:1419–1426
59. Shirahase T, Komatsu Y, Marubayashi H, Tominaga Y, Asai S, Sumita M (2007) Miscibility and
hydrolytic degradation in alkaline solution of poly(L-lactide) and poly(p-vinyl phenol) blends. Polym
Degrad Stab 92:1626–1631
60. Tsuji H, Muramatsu H (2001) Blends of aliphatic polyesters: V Non-enzymatic and enzymatic
hydrolysis of blends from hydrophobic poly(L-lactide) and hydrophilic poly(vinyl alcohol). Polym
Degrad Stab 71:403–413
61. Huang Y, Ge FJ, Zhou YL, Jiang L, Dan Y (2014) Hydrolytic behavior of poly(lactic acid) films with
different architecture modified by poly(dodecafluorheptyl methylacrylate). Eur Polym J 59:189–199
62. Ojijo V, Ray SS, Sadiku R (2013) Toughening of Biodegradable polylactide/poly(butylene succinate-
co-adipate) blends via in situ reactive compatibilization. ACS Appl Mater Inter 5:4266–4276
63. Shibata M, Inoue Y, Miyoshi M (2006) Mechanical properties, morphology, and crystallization
behavior of blends of poly(L-lactide) with poly(butylene succinate-co-L-lactate) and poly (butylene
succinate). Polymer 47:3557–3564
64. Papageorgiou DG, Chrissafis K, Pavlidou E, Deliyanni EA, Papageorgiou GZ, Terzopoulou Z,
Bikiaris DN (2014) Effect of nanofiller’s size and shape on the solid state microstructure and thermal
properties of poly(butylene succinate) nanocomposites. Thermochim Acta 590:181–190
65. Aleman C, Lotz B, Puiggali J (2001) Crystal structure of the alpha-form of poly(L-lactide).
Macromolecules 34:4795–4801
66. Sasaki S, Asakura T (2003) Helix distortion and crystal structure of the alpha-form of poly(L-lactide).
Macromolecules 36:8385–8390
67. Wang RY, Wang SF, Zhang Y, Wan CY, Ma PM (2009) Toughening modification of PLLA/PBS
blends via In Situ compatibilization. Polym Eng Sci 49:26–33
68. Cho K, Lee J, Kwon K (2001) Hydrolytic degradation behavior of poly(butylene succinate)s with
different crystalline morphologies. J Appl Polym Sci 79:1025–1033
69. Tsuji H, Ikarashi K, Fukuda N (2004) Poly(L-lactide): XII. Formation, growth, and morphology of
crystalline residues as extended-chain crystallites through hydrolysis of poly(L-lactide) films in
phosphate-buffered solution. Polym Degrad Stab 84:515–523

123

You might also like