You are on page 1of 14

Journal of Biomaterials

http://jba.sagepub.com/
Applications

Τhe effect of silica nanoparticles on the thermomechanical properties and degradation behavior of
polylactic acid
P Georgiopoulos, E Kontou, A Meristoudi, S Pispas and M Chatzinikolaidou
J Biomater Appl 2014 29: 662 originally published online 4 August 2014
DOI: 10.1177/0885328214545351

The online version of this article can be found at:


http://jba.sagepub.com/content/29/5/662

Published by:

http://www.sagepublications.com

Additional services and information for Journal of Biomaterials Applications can be found at:

Email Alerts: http://jba.sagepub.com/cgi/alerts

Subscriptions: http://jba.sagepub.com/subscriptions

Reprints: http://www.sagepub.com/journalsReprints.nav

Permissions: http://www.sagepub.com/journalsPermissions.nav

Citations: http://jba.sagepub.com/content/29/5/662.refs.html

>> Version of Record - Oct 3, 2014

OnlineFirst Version of Record - Aug 4, 2014

What is This?

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
Biomaterials Processing
Journal of Biomaterials Applications
2014, Vol. 29(5) 662–674
! The Author(s) 2014
She effect of silica nanoparticles on the Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav
thermomechanical properties and DOI: 10.1177/0885328214545351
jba.sagepub.com
degradation behavior of polylactic acid

P Georgiopoulos1, E Kontou1, A Meristoudi2, S Pispas2 and M Chatzinikolaidou3,4

Abstract
In this work a series of polylactic acid/SiO2 nanocomposites have been prepared by a melt mixing procedure. The
dispersion quality was examined by scanning electron microscopy. To study the degradation behavior of the polylactic
acid/nanocomposites prepared, the samples were immersed in a buffer solution at a temperature of 37 C with a pH of
7.4 for a time period of up to 23 weeks. These conditions simulate those in the human body, appropriate in medical
applications. In order to assess their suitability in biomedical applications, we investigated the biocompatibility of these
materials in terms of cell viability, growth, and morphology. A good initial cell adhesion has been detected, supporting
their potential use in bone tissue engineering applications. The hydrolytic degradation of polylactic acid, under the
prescribed conditions, was studied by the molecular weight reduction in terms of size exclusion chromatography,
whereas the progress of thermal stability of polylactic acid and polylactic acid/nanocomposites during aging was
tested by thermogravimetric analysis. The evolution of the materials’ thermomechanical properties during aging was
studied by differential scanning calorimetry, dynamic mechanical analysis, and tensile testing. The crystallization behavior
in polylactic acid and the way it is affected by the presence of nanofillers during degradation procedure has been studied
and values of 44% crystallinity increment have been found. At the specific aging conditions studied, silica nanoparticles
accelerate the degradability of polylactic acid, having a higher impact on Young’s modulus, under the specified aging
conditions, for 7 weeks and hereafter this acceleration is retarded, due to the crystallinity increment, as a result of the
molecular weight reduction.

Keywords
Biodegradation, cytotoxicity testing, nanocomposites, mechanical properties, crystallinity

Introduction polymers also create undesirable effects which must


Polylactic acid (PLA) is nowadays one of the most be overcome. For several applications the degradation
important polymers, due to its combined properties, rate of PLA is still low as compared to the waste accu-
such as biocompatibility, biodegradability, and ade- mulation rate. On the other hand, the fact that PLA
quate mechanical performance. PLA consists of an
alternative to petrochemical-based polymers, because
1
it can be produced from natural sources such as corn School of Applied Mathematical and Physical Sciences, Department of
Mechanics, National Technical University of Athens, Athens, Greece
and wheat and is therefore related to a wide range of 2
Theoretical and Physical Chemistry Institute, National Hellenic Research
medical, textile, and packaging applications. The Foundation, Athens, Greece
(bio)degradability of PLA though has both negative 3
Department of Materials Science and Technology, University of Crete,
and positive aspects. The positive aspect is the (bio)deg- Heraklio, Greece
4
radation to form nonhazardous products when PLA Institute of Electronic Structure & Laser, Foundation for Research &
Technology Hellas, Heraklio, Greece
polymers or articles are discarded or composted after
completing their useful life.1 The negative aspect is the Corresponding author:
E Kontou, School of Applied Mathematical and Physical Sciences,
degradation of PLA polymers during processing. Thus, Department of Mechanics, National Technical University of Athens, 5
the same properties that make PLA polymers desirable Heroes of Polytechnion, 15773 Athens, Greece.
as replacements for nondegradable petrochemical Email: ekontou@central.ntua.gr

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
Georgiopoulos et al. 663

degrades slowly—over a period of several weeks up to The present work is complementary to our previous
about 1 year—is advantageous for some other applica- works1,19 on the degradation effects on crystalline and
tions such as biomedical, since it leads to a relatively thermomechanical properties of PLA/nanocomposites.
good shelf life. In the present work, the degradation ability of PLA
PLA degrades during thermal processing or under reinforced with nanosilica particles at three different
hydrolytic conditions, giving a reduction of molecular contents, 2, 3, and 5% wt, under specific environmental
weight that affects the final properties of the material, conditions, namely immersion in a buffer solution at
such as the mechanical strength.2,3 The degradation temperature of 37 C with a pH of 7.4 has been studied
behavior depends strongly on the molecular weight experimentally. These conditions simulate those in the
and the crystallinity of the PLA.4 Previous studies human body, appropriate in medical applications. The
have shown that degradation of PLA is enhanced by majority of prior art studies on the stability of PLA have
an increase in temperature and relative humidity (RH).5 been carried out at similar physiological conditions,4,6,23
The temperature, in presence of oxygen, triggers ther- but there are still some issues that need to be addressed.
mal oxidation, while the moisture promotes hydrolytic A matter of importance has been proved to be the crys-
degradation. PLA degradation driven by hydrolysis tallization behavior in PLA and the way it is affected by
needs higher temperature in order to take place the presence of nanofillers during degradation proced-
(T > 50 C). ure. As it is mentioned in Chen et al.18 the effect of silica
Chemical hydrolysis of the hydrolytically unstable on the hydrolytic degradation behavior of PLA is less
backbone is the primary mechanism for degradation researched. Apart from this, the role of nanofillers on the
of the PLA polymer. Degradation occurs first by evolution of thermomechanical properties of PLA/nano-
water penetrating the bulk of the polymer and hydro- composites at various stages of degradation has been less
lyzing the ester bonds, preferentially those in the examined, and this work is a contribution toward this
amorphous phase, and converting long chains into direction. The stability of PLA and PLA/nanocompo-
shorter water-soluble fragments.6 sites was monitored by a number of techniques, includ-
On the other hand, the thermomechanical and bar- ing size exclusion chromatography (SEC), scanning
rier properties of PLA can be enhanced by the incorp- electron microscopy (SEM), thermogravimetric analysis
oration either of biocompatible fibers7 or of a fairly (TGA), differential scanning calorimetry (DSC),
low amount of nanosized fillers (1–5 wt%).8–10 Many dynamic mechanical analysis (DMA), and tensile meas-
studies have been reported on the preparation and urements. In order to assess their suitability in biomed-
characterization of a wide range of PLA nanocompo- ical applications, we investigated the biocompatibility of
sites.10–12 A large variety of inorganic nanoparticles, the materials tested in terms of cell viability, growth, and
including silica,13 layered silicates,14 hydroxyapatite,15 morphology. We therefore cultured MC3T3-E1 preos-
nanotubes,16 have been investigated for this purpose. teoblastic cells on films made either from PLA or PLA
Among the various types of nanoparticles, layered sili- with 2% wt silica nanocomposite, visualized the morph-
cates, which are naturally abundant, economic, and ology of the adhered cells, and quantified the cell prolif-
benign to the environment, have been most extensively eration on them.
used for the preparation of PLA nanocomposites.17
Besides the thermomechanical properties, the hydro-
lytic degradation behavior of PLA is also greatly
Materials
affected by the presence of nanofillers. The manner in PLA was employed for the preparation of the nano-
which the thermomechanical properties of PLA and composites. The selected grade of PLA (supplied by
PLA/nanocomposites alter over time is still an issue Nature works) has a D content of 4.25%, a residual
of great importance. Depending on the applications, monomer content of 0.3%, and a density of 1.24 g/
acceleration of degradation (medical devices) is cm3. The material in pellets form was dried at 45 C
required, while in some other cases (packaging) sup- for a minimum of 8 h prior to use in a desiccating dryer.
pression of degradation is needed. The effect of various For the preparation of the PLA nanocomposites,
nanofiller types on the PLA degradation behavior has Silica Aerosil R972 (supplied by Degussa Chemicals)
been extensively studied18–22 and it has been reported has been employed, which is hydrophobic fumed
that the hydrolytic degradation of PLA takes place first silica after treated with dimethyldichlorosilane and
at the interface between PLA matrix and nanofillers based on hydrophilic fumed silica, with a specific sur-
rather than in the bulk polymer.18 A complex effect face area of 130 m2/g. The average primary size is
on the degradation behavior of PLA has been revealed 16 nm. A series of PLA nanocomposites with a filler
in Kontou et al.,19 where the effect of two different content of 2, 3, and 5 wt.% of silica (Si) have been
nanofiller types (silica and montmorillonite) and their prepared and designated as PLA/Si/2, PLA/Si/3, and
mixtures has been studied on the degradability of PLA. PLA/Si/5.

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
664 Journal of Biomaterials Applications 29(5)

SEM images were obtained by direct observation of


Experimental the topography of the sample’s surface with a NovaTM
For the preparation of the materials, melt mixing of NanoSEM 230 (FEI, Hillsboro, Oregon, USA) scan-
nanofillers with matrix was performed with a ning microscope operated at an acceleration voltage
Brabender mixer at a temperature of 180 C, while the of 5 or 15 kV. All the studied samples were coated
rotation speed of the screws was 40 r/min. The tempera- with gold to avoid charging under the electron beam.
ture for melt mixing was kept as low as possible in She studied surface of the samples was the fractured
order to minimize the effect of degradation. cross-section obtained from the tensile test.
Hereafter, the material was compression molded at
130 C, using a thermo-press and a special mold of
2 mm thickness. The material was then cooled slowly
Cell culture
down at ambient temperature. For each material type, Minimum essential Eagle’s medium (a-MEM), penicillin/
two samples were prepared in the same manner, which streptomycin, fetal bovine serum (FBS), and trypsin/ethy-
provided reproducible results. lenediaminetraacetic acid (EDTA) were purchased from
To study the effect of nanofillers on the degradability Sigma-Aldrich. PrestoBlueTM reagent and carboxyfluores-
of PLA, PLA/nanocomposites were exposed to specific cein diacetate succinimidyl ester (CFSE) cell viability assay
aging conditions, namely immersed in a buffer solution were purchased from Invitrogen Life Technologies. Cell
at a temperature of 37 C with a pH of 7.4, for a time culture plates were obtained from Corning.
period of up to 23 weeks. The effect of the above-men- Early passages of MC3T3-E1 preosteoblastic cells
tioned aging conditions has been studied at time inter- (DSZM, Germany) were cultured in alpha-MEM sup-
vals of this period. plemented with 10% FBS and 1% penicillin/strepto-
For the in vitro biocompatibility assessment, we mycin and maintained in an incubator (Thermo
employed thin films on round glass substrates of Scientific) at 37 C in a humidified atmosphere with
15 mm in diameter, by drop casting the materials’ 5% CO2. Cultures were medium changed every 2–3
PLA or PLA with 2% per weight nanosilica. The days and subcultured using trypsin/EDTA. Prior to
glass substrate surface was covered with the material seeding on the different substrates, confluent cells
and dried in an oven at 60 C. Samples were disinfected were harvested using trypsin/EDTA and counted in a
in 70% ethanol and air-dried in a laminar flow bench Neubauer chamber.
prior to cell seeding on them.
The evolution of molecular weight of the PLA at vari-
ous aging time intervals was recorded by SEC. All aged In vitro cytotoxicity testing/assessment
and nonaged samples were dissolved in tetrahydrofuran
(THF) without any insoluble residues being visible to the
Optical and fluorescence microscopy
naked eye, suggesting that no crosslinking took place. A suspension of 2  104 cells in a-MEM cell culture
The molecular weights and molecular weight distribu- medium was seeded on the PLA-based material films
tions of the PLA samples were determined by SEC and placed in the cell culture incubator at 37 C. The
using a Waters system, composed of a Waters 1515 iso- cells on the samples were examined daily and visualized
cratic pump, a set of three m-Styragel mixed bed col- by means of a Zeiss Axiovert 200 microscope. Images
umns, with a porosity range of 102–106 Å, a Waters were taken by a ProgResÕ CFscan Jenoptik camera
2414 refractive index detector (at 40 C) and operated/ (Jena, Germany) using the ProgResÕ CapturePro 2.0
controlled through Breeze software. THF, containing software and an objective lens for the 20-fold magnifi-
3% v/v triethylamine, was the mobile phase used at a cation. The visualization of the living cells by means of
flow rate of 1.0 ml/min at 30 C. The setup was calibrated the CFSE fluorescent dye was performed in the same
with linear polystyrene standards having weight average microscope using a fluorescence lamp and a filter for
molecular weights in the range 1250–900,000 g/mol. The emission at 520 nm. Images were taken with the above-
PLA samples were dissolved directly to the carrier solv- mentioned software with an exposure time set at
ent overnight at concentrations ca. 0.1% w/v and char- 152 ms.
acterized the next day.
TGA was performed by using a TGA Q500 V20.2
Build 27 instrument by TA Instruments (New Castle,
Cell viability and proliferation on the materials
DE) in an inert atmosphere of nitrogen. In a typical Cell viability and proliferation of MC3T3-E1 preosteo-
experiment, 10 mg of the material was placed in the blasts were assessed using the Presto BlueTM assay,
sample pan, and the temperature was equilibrated at which is a resazurin-based non-toxic metabolic indica-
25 C. The temperature was then increased to 600 C tor for viable cells.24 PLA-coated samples were disin-
at a rate of 20 C/min. fected with ethanol and transferred to an empty plate

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
Georgiopoulos et al. 665

before assaying. A total of 2  104 cells in complete min. This value corresponds to an effective strain rate
alpha-MEM culture medium were seeded on the sam- of 1.1  10–4 s–1. The deformation could be measured
ples and placed in the cell culture incubator at 37 C. On very accurately with an experimental procedure,
days 1, 3, and 7 postseeding cell viability and prolifer- which is based on a noncontact method with a laser
ation assay was performed with the PrestoBlueTM extensometer, described in detail in a previous work.26
reagent according to the manufacturer’s instructions. Tensile stress–strain curves were then obtained up to
The absorbance was measured in a spectrophotometer the breaking point. Five specimens were tested for
(Molecular Devices SpectraMax M2) and cell number each aging procedure, and the scatter between experi-
quantification was performed by means of a calibration mental data was lower than 8%.
curve. Error bars represent the average of tripli-
cates  standard deviation in two independent experi-
ments for each time point. Results and discussion
Examination of cell metabolic activity on the mater-
SEC results
ials was performed by staining with CFSE, which is
a nonfluorescent, cell-permeable dye that diffuses into The degradation of PLA matrix has been tested by
cells and the acetate groups are cleaved by intracellular examining the molecular weight distribution after
esterases to yield highly fluorescent carboxyfluorescein exposure to the prescribed aging conditions, for differ-
succinimidyl ester. After cleavage of the acetate ent time periods. The weight average molecular weight
groups, it has a peak excitation of 494 nm and peak M w versus aging time is given in Table 1. A substantial
emission of 521 nm. Cells (2  104) were cultured on molecular weight decrement is obtained after 163 days
PLA-containing samples for 7 days, then washed with of aging.
phosphate buffered saline (PBS), stained for 15 min with As can be seen in Table 1, after aging at 37 C for 65
20 mM CFSE dye in PBS and viewed by epifluorescence days, PLA underwent little change in its molecular
microscopy. weight. Significantly lower molecular weights were
DSCs were carried out using a Setaram DSC 141 measured for samples after 163 days (23 weeks) of
instrument, calibrated with an Indium standard. Each aging. This is exemplified in Figure 1, which shows
PLA sample (aged and nonaged) was heated at a con- the elution curves of PLA samples for different time
stant heating rate of 10 C/min from 20 C up to 170 C, intervals. As can be seen in Figure 1, the elution peak
and the thermogram was recorded. The degree of crys- shifted to lower molecular weight with increasing aging
tallinity was calculated by considering a melting time. Furthermore, the elution peaks of the aged sam-
enthalpy of 93.1 J/g for 100% crystalline PLA.25 ples display a higher contribution of low molecular
Three sheet specimens of 2 mm thickness were tested weight fractions, which is attributed mainly to chain
for every aging procedure. DSC samples were taken scission due to hydrolysis.
from both the center and the tip of every PLA sheet, The degradation rate was quantified in terms of the
and it was found that the DSC results were repeatable, average hydrolytic degradation rate constant (kt). The
with an average scatter lower than 5%. In our study, (kt) values were evaluated assuming an exponential
especially for DSC tests, the aging time period was 36 decrease of Mn using the following equation4
weeks.
The DSC samples were washed with distilled water lnMn ðt2 Þ¼ lnMn ðt1 Þkt t ð1Þ
two times and dried in the ambient for 2 h, before
testing. where Mn(t2) and Mn(t1) are the Mn values at the
DMA experiments were performed using the TA hydrolytic degradation times of t2 and t1, respectively.
Instruments DMA Q800 instrument. The mode of In Figure 2, the molecular weight changes are plotted
deformation applied was the single cantilever beam, logarithmically with varying the aging time. In the same
and the mean dimensions of sample plaques were
12.6 mm  2 mm  17.5 mm. In order to obtain a
direct measure of the storage modulus evolution
versus aging time, the test temperature was kept con- Table 1. PLA molecular weight variation with aging time.
stant equal to 20 C, so no further thermal effects could
take place. The experiments were performed at a con- Aging time (weeks) Molecular weight (g/mol)
stant frequency of 1 Hz. 0 194,873
Tensile measurements were performed with an 9.0 169,439.5
Instron 1121 type tester, at room temperature. The 23 89,609
dumbbell type specimens were of a gauge length of
48 69,600
30 mm, and the applied crosshead speed was 0.2 mm/

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
666 Journal of Biomaterials Applications 29(5)

4
results.2 Regarding the aging temperature effect, the
degradation process was very slow at temperatures
3
equal or below 40 C, but the speed was greatly
0 weeks enhanced at 50 C. The hydrolytic degradation rate at
the buffer solution 37 C appears to be almost the same
2 with the conditions of 40 C and 80% RH.
MV

23 weeks
9 weeks
1 SEM results
The morphological changes incurred in the PLA/nano-
0
48 weeks composites after exposure to the prescribed aging con-
ditions were analyzed by SEM micrographs and shown
16 18 20 22
Minutes
24 26 28 30 representatively in Figure 3(a) to (d) for PLA/Si/3 and
PLA/Si/5 nanocomposites. From these photos, the
Figure 1. Elution curves for PLA at different aging times good quality of nanofiller’s dispersion is apparent,
obtained by size exclusion chromatography. given that the nanoparticles are visible and quite homo-
geneously dispersed. Nanoparticles of a few nano-
meters size coexist with larger aggregates, with the
12.4 aggregates being larger with increasing the nanofiller
12.2 content. For PLA/Si/3 after 23 weeks of aging, an
12.0 increasing surface roughness is observed, as well as
11.8
the existence of holes (dark regions) especially around
11.6
nanoparticles. Analogous trend is obtained for PLA/Si/
5 where, in some cases, holes connecting neighboring
Ln(Mw)

11.4
nanoparticles are also created.
11.2

11.0 80% RH/20°C


80% RH/40°C
80% RH/50°C
10.8
buf.soI/37°C
In vitro cytotoxicity testing/assessment
10.6

10.4
Initial cell adhesion on PLA and PLA þ SiO2 coatings
0 20 40 60 80 100 120 140 160 180 The optical microscopy images indicate a good initial
Time (Days) cell adhesion 2 h postseeding on both materials, the
PLA (Figure 4(b)) and PLA/Si/2 (Figure 4(c)), which
Figure 2. PLA molecular weight variation versus time at
is comparable to the tissue culture treated polystyrene
various aging conditions.
control polystyrene (PS) (Figure 4(a)). Cells extend elon-
gated protrusions, displaying a spindle-shape morph-
ology, and attach in a similar number and manner on
plot, analogous results for PLA, degraded under differ- both PLA-containing coatings as on the preferable PS
ent conditions1 are depicted for comparison. The control. These data depict a characteristic morphology
experimental procedure applied for aging PLA at of preosteoblastic cells seeded on a biocompatible mater-
80% RH and various temperatures is presented in ial surface without showing any adverse effects.
detail in Niaounakis et al.1 Following Figure 2, an aver-
age value kt for the hydrolytic degradation rate has
Cell proliferation on PLA and PLA þ SiO2 coatings
been calculated. The estimated kt values for the differ-
ent time intervals were 3.73  10–4, 2.96  10–3, and The graph in Figure 5 shows the number of living cells
20  10–3 days–1 at 20, 40, and 50 C, and 80% RH, on the tissue culture treated polystyrene control sur-
respectively. The sample immersed in the buffer face, PLA and PLA/Si/2 after 1, 3, and 7 days in cul-
solution at 37 C had an average degradation rate ture. Beginning with an initial number of 2  104 cells
equal to 4.9  10–3 days–1 for a similar time period of seeded on each sample type, we observe a significantly
aging with the other samples. When the time period is increasing cell proliferation after 3 and 7 days in culture
extended to 340 days, the value of kt is reduced to on both PLA and PLA/Si/2 materials, which is similar
3.185  10–3 days–1. The results showed that the degrad- to the PS control surface. Among the different material
ation rate was higher at the beginning (first 60 days) for surfaces at each experimental time point, we do not
every temperature and decreased significantly after 30 observe any statistically significant proliferation
days. This finding was in agreement with previous increase.

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
Georgiopoulos et al. 667

Figure 3. SEM micrographs of PLA/Si/3 after 23 weeks of aging. SEM micrographs of PLA/Si/5.

Figure 4. Initial preosteoblastic cell adhesion on PS (control) (a), PLA (b), and PLA/Si/2 (c) after 2 h in culture shown by optical
microscopy indicating a comparable number of cells with a similar spindle-shaped morphology attached on all surfaces, 20-fold
magnification. Scale bar represents 50 mm.

cells appear to have increasingly proliferated and


formed a confluent live cell layer on all three surfaces.
Cell metabolic activity
Additionally, these data corroborate with the quantita-
Figure 6 shows the epifluorescence of living preosteo- tive cell proliferation analysis shown in Figure 5.
blastic cells stained with the CFSE dye on either tissue
culture treated polystyrene control (a), or PLA (b), or
PLA/Si/2 (c) after 7 days in culture. We observe a com-
TGA results
parable number of living cells appearing in green, with The TGA results are summarized in Table 2, where the
a similar flattened morphology on both PLA and PLA/ 10 wt% loss temperature with varying aging time is pre-
Si/2 materials, as on the polystyrene control. The sented for all samples. From this table it is extracted
depicted high cell density and the polygonal cell morph- that the thermal stability of PLA and PLA/Si/5 nano-
ology are evident for well-attached living cells on the composite remains almost unaffected with aging time,
material coatings. After 7 days in culture, and com- while for PLA/Si/2 and PLA/Si/3 it is reduced and
pared to the initial adhesion stage shown in Figure 4, shifted toward lower temperatures. This effect denotes

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
668 Journal of Biomaterials Applications 29(5)

that the good quality dispersion of silica at those sam- restriction of silica nanoparticles on the segmental
ples, as postulated by SEM, increases the interfacial and long-range chain mobility of the matrix takes
region, which is more sensitive to the aging conditions. place, which would lead to a Tg increment of the nano-
The formation of large agglomerates in PLA/Si/5 does composites. On the other hand, no indication of poor
not have the same effect. adhesion between matrix and nanofillers exists, since in
that case a Tg reduction should be obtained.27 A simi-
lar effect, where the addition of 2% vol. of fumed silica
DSC results on PLA does not significantly affect the Tg of the
The DSC results of the PLA nanocomposites and their material has been found.28
evolution with aging time are summarized in Table 3. In the following paragraphs, the evolution of ther-
PLA nanocomposites exhibit three transitions, the mal properties with aging time of PLA/nanocomposites
glass transition temperature Tg, the cold crystallization is analytically presented. Following Table 3, the Tg of
temperature Tcc (for PLA matrix), and the melting PLA matrix is slightly decreasing during the first 12
regime Tm. weeks of aging, and hereafter starts increasing. The
The Tgs of the PLA/nanocomposites were not Tg of PLA/nanocomposites does not substantially
affected by the presence of silica, meaning that no change with aging time, exhibiting a slight increment
after 36 weeks of aging. Generally, the Tg decrement is
attributed to the molecular weight reduction of the
samples. In previous works it has been found that the
80000 Tg is either not significantly affected by aging20 or it is
PS decreased with exposure time.29 In our case, the Tg is
70000
PLA generally increased after a long time period of aging for
60000 PLA+SiO2 PLA and PLA/nanocomposites, with this increment
being more essential in the PLA matrix. The Tg incre-
50000
ment with aging has also been detected for similar
Cell number

40000
materials in Georgiopoulos et al.30
From Figures 7 to 10 where DSC curves are illu-
30000 strated for PLA and PLA/nanocomposites correspond-
ingly, it is shown that the Tg region is expressed by an
20000

10000

0 Table 2. TGA 10 wt% loss temperature in  C at various aging


Day 1 Day 3 Day7 times.
Figure 5. Cell proliferation on tissue culture treated polystyr- Aging time
ene (control), PLA and PLA/Si/2 after 1, 3, and 7 days in culture (weeks) PLA PLA/Si/2 PLA/Si/3 PLA/Si/5
by means of the PrestoBlueTM assay. Cells proliferate on both
PLA-containing materials similarly to the PS control surface. 0 335.8 342 332.7 333
Error bars represent the average of triplicates  STDV for each 14 333 332.5 331.5 331
time point. Optical density values are normalized to the cell 24 330.5 327.5 325 330.8
number according to a calibration curve.

Figure 6. Epifluorescence of living cells stained with the CFSE dye on tissue culture treated polystyrene control (a), PLA
(b), and PLA/Si/2 (c) after 7 days in culture. A comparable number of living cells with a similar flattened morphology is observed on
both PLA-containing materials as on the PS control. Images are taken in a 20-fold magnification, exposure time is 152 ms. Scale bar
represents 50 mm.

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
Georgiopoulos et al. 669

Table 3. DSC properties of PLA/nanocomposites with varying aging time.

Material/aging Degree of Percentage change


time (weeks) Tg ( C) Har (J/g) Hcc (J/g) Ham (J/g) crystallinity XC (%) crystallinity (%) Tm ( C) Tm1/Tm2

PLA
0 58.42 3.49 5.31 22 18.0 0 146.1/154.4
3.0 56.8 5.99 8.16 22.49 15.4 14.4 146.25/154.52
12.0 57.1 7.86 6.9 23.77 18.1 0.55 146/155.2
18.0 58.4 8.32 9.05 24.66 16.8 6.6 146/155.8
23.0 60.24 7.9 9.36 22.45 14.0 22.2 146.8/157.8
36.0 63.7 8.35 9.33 27.72 19.8 10.0 146/156.0
PLA/Si/2
0 58.81 1.06 22.55 24.24 0 151/155.3
3.0 59.29 3.88 24.18 26.0 7.26 149.7/155.29
12.0 58.1 4.22 25.52 27.44 13.2 149.54/155.9
18.0 58.15 3.7 27.33 29.38 21.2 149.56/156.4
23.0 60.15 5.0 24.71 26.56 9.57 150/157.74
36.0 59.0 6.25 31.55 33.92 40.0 149.47/156.81
PLA/Si/3
0 58.22 1.15 21.89 23.53 0 150/155.3
3.0 61.0 3.13 24.64 26.49 12.6 150.38/154.8
12.0 58.2 4.11 25.62 27.54 17.0 149.7/155.9
18.0 58.82 4.51 26.39 28.37 20.6 150.1/156.5
23.0 59.6 4.75 27.34 29.39 25.0 149.5/156.3
36.0 62.4 6.0 30.1 32.36 37.52 150/156.54
PLA/Si/5
0 58.15 0.92 20.88 22.45 0 150.3/155.8
3.0 61.5 3.76 23.47 25.23 12.38 149.8/154.7
12.0 57.9 4.0 23.72 25.5 13.58 150.38/156.12
18.0 60.31 3.78 22.79 24.5 9.13 150.1/157.1
23.0 59.8 5.37 27.63 29.7 32.3 150.8/156.19
36.0 62.34 6.67 30.2 32.36 44.14 149.5/156.67
a
Enthalpy values are normalized with respect to pure polymer.

endothermic peak, corresponding to an enthalpy relax-


PLA
ation Hr. The existence of such an endothermic peak
is related to the melting of a mesophase, which is a
36 weeks
region with some kind of molecular ordering.31,32
As presented in Table 3, for the unaged materials, a
23 weeks
substantial Hr decrement with the addition of nano-
Heat Flow (mw)

18 weeks particles, which implies a confinement of a fraction of


exo PLA chains around nanofillers surface, exists, which
0.00 cannot contribute to a relaxation around Tg.33 In con-
12weeks
trast, Hr of all samples is increased with aging time,
3 weeks
and this increment is enlarged at PLA/nanocomposites
0 weeks compared to that of PLA matrix. This effect denotes
that after the sample has been hydrolyzed for a certain
120
period of time, the formation of this locally ordered
40 60 80 100 140 160 180
°Temperature (°C)
structure is enhanced.
As shown in Figure 7, PLA matrix exhibits the exo-
Figure 7. DSC curves of PLA at various aging times. thermic effect, attributed to the cold crystallization of

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
670 Journal of Biomaterials Applications 29(5)

PLA/Si2 PLA during the DSC heating process. It has been found
36 weeks that the cold crystallization temperature Tcc is decreas-
ing with aging time, while the corresponding enthalpy
23 weeks
Hcc is increasing. When cold crystallization occurs,
Heat Flow (mw)

18 weeks imperfect crystals are formed, which melt during the


0.0 exo DSC heating scan. In addition, the Hcc increment,
12 weeks related to the enlarged cold crystallization area of
PLA matrix after 36 weeks of aging (Figure 7), indi-
3 weeks cates the presence of more polymeric chains involved in
the cold crystallization process, thus strengthening the
0 weeks interpretation of morphological changes previously
drawn from the Hr assessment. Regarding crystallin-
40 60 80 100 120 140 160 180
ity, the PLA/nanocomposites, in the unaged stage,
Temperature (°C) exhibit a percentage change crystallinity, namely 34.7,
30.7, and 24.7% for PLA/Si/2, PLA/Si/3, and PLA/Si/5
Figure 8. DSC curves of PLA/Si/2 at various aging times. correspondingly. These values are calculated with
respect to the crystallinity content of PLA matrix,
and this effect shows that silica nanoparticles act as
nucleating agents.
PLA/Si3 As far as crystalline behavior of PLA and PLA/
nanocomposites under aging conditions is concerned,
36 weeks it must be noted that for a semicrystalline polymer,
23 weeks
the initial crystalline structure strongly affects the
hydrolytic degradation of PLA.18,34
Generally, hydrolytic degradation occurs first in the
Heat Flow (mw)

18 weeks
exo
12 weeks amorphous region. The degree of crystallinity Xc(%) of
0.0 PLA with increasing aging time exhibits a nonmono-
3 weeks
tonic behavior, but generally has a tendency to increase
up to a value of 10% (Table 3). PLA/Si nanocompo-
sites, on the other hand, appear as monotonic incre-
0 weeks
ment of the degree of crystallinity with aging time, up
60 80
to values of the order of 40–44% with respect to the
40 100 120 140 160 180
Temperature (°C) initial value before aging. This effect consists of a
strong indication that silica is an effective nucleating
Figure 9. DSC curves of PLA/Si/3 at various aging times. agent, promoting the crystallization of PLA matrix,
during the hydrolytic degradation.18,34 It has been
reported that the increased heat of fusion (and there-
fore degree of crystallinity increment) is attributed to
plasticization of PLA by water molecules as well as by
PLA/Si5 lactic oligomers that would give sufficient mobility to
the polymer chains to organize and further crystallize.22
36 weeks
This is the result of two competitive effects: The
silica content leads to the increase of the hydrophilicity
23 weeks
of the sample, which results in the damage of more
Heat Flow (mw)

exo crystalline structure during the hydrolytic degradation


process. On the other hand, due to degradation, we
0.0 18 weeks
have easier crystal formation, and the nucleation of
3 weeks
new crystalline regions, where smaller chains are
involved.
0 weeks Analyzing further the evolution of crystalline struc-
ture during aging, the samples obtained at various
40 60 80 100 120 140 160 180
aging time periods were examined by DSC. From
Temperature (°C) Figures 7 to 10, illustrating PLA and PLA/nanocom-
posites, as far as the melting region is concerned, two
Figure 10. DSC curves of PLA/Si/5 at various aging times. endothermic peaks appear in all material types.

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
Georgiopoulos et al. 671

The melting point generally is directly related to the Table 4. Tensile properties of PLA/nanocomposites with
quality of the crystalline region, and consequently the varying aging time.
thickness of the lamellae; the bigger the lamellae thick-
Material/aging Young’s Yield Strain
ness, the higher the melting peak temperature. time (weeks) modulus (GPa) stress (MPa) at break
Therefore, double melting peaks are generally asso-
ciated with the presence of two main crystal popula- PLA
tions, with different size and quality.35–37 Following 0 3.5 54.9 0.055
Table 3, the low temperature peak Tm1 for PLA/nano- 3.0 3.0 52.65 0.074
composites is higher than that of PLA matrix, whereas 7.0 3.03 54.4 0.033
the second melting peak Tm2 is almost the same for all 14.0 3.3 54.3 0.04
materials examined. The first peak Tm1 has been related 23.0 3.11 49.6 0.025
to the fusion of rather imperfect crystals, formed by PLA/Si/2
primary crystallization, whereas the second melting
0 4.87 56.2 0.055
peak Tm2 can be related to the melting of more perfect
3.0 3.6 53.32 0.062
crystals, which were formed during a reorganization
procedure upon heating. 7.0 3.6 54.3 0.048
With increasing aging time, it can be observed from 14.0 3.64 51.0 0.02
Figures 7 to 10 and Table 3 that Tm1 is almost constant 23.0 3.75 0.012
with varying aging time (with an exception for PLA/Si/ PLA/Si/3
2, where Tm1 slightly decreases), while Tm2 exhibits a 0 4.83 59.0 0.044
slight increment with aging time, for all material types. 3.0 3.67 53.2 0.061
It has been reported38 that when Tm decreases with 7.0 3.78 54.35 0.075
aging, crystalline regions are partially degraded, spe- 14.0 4.0 49.5 0.016
cially the interlamellar amorphous phases into the crys- 23.0 3.7 37.2 0.012
talline regions, even to a less extent than in the PLA/Si/5
amorphous regions.
0 4.85 58.7 0.07
In our case, following Figures 7 to 10, the Tm2 peak
3.0 3.58 55.9 0.032
becomes more intense with aging time for PLA/nano-
composites. It probably happens, due to a mechanism 7.0 3.62 53.62 0.06
of melting and recrystallization of less perfect crystal- 14.0 3.65 47.2 0.02
lites into thicker ones, which hereafter melt at a higher 23.0 3.87 39.3 0.012
temperature. Aging procedure seems to facilitate this
mechanism. This fact combined with the enhancement
of the total heat of fusion with aging, denotes that silica
can promote the crystallization of PLA matrix during 60

the hydrolytic degradation process for the specific aging 14 weeks 0 weeks
50
conditions applied. 7 weeks
3 weeks
23weeks
40
Stress (MPa)

Tensile testing results


30
The experimental tensile properties of PLA/Si nano-
PLA
composites are extensively analyzed in Kontou et al.19 20
and Georgiopoulos et al.30 The corresponding evolu-
tion of the tensile properties of pristine PLA and its 10
nanocomposites with aging at the specific conditions
examined, at various time periods are presented in 0
0.08
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Table 4 and in Figures 11 and 12 representatively for Strain
PLA and PLA/Si/3. From Table 4, PLA/Si nanocom-
posites exhibit a 39% Young’s modulus increment com- Figure 11. Tensile stress–strain curves for PLA at various aging
pared to the Young’s modulus of PLA matrix. From conditions.
Figure 11 of pristine PLA, it can be observed that in all
stages of aging, a yield stress and a subsequent strain PLA/nanocomposites exhibit a 26% reduction.
softening is exhibited. Following Table 4, it can be Hereafter, all studied samples exhibited almost the
noticed that after 3 weeks of aging, PLA matrix under- same trend in tensile properties during aging, i.e. their
goes a 14% reduction in the Young’s modulus, while Young’s modulus appears to be not further greatly

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
672 Journal of Biomaterials Applications 29(5)

Stain amplitude =0.2% PLA


60 4.0
0 weeks frequency=1 Hz PLA/Si/2
Temperature= 37°C PLA/Si/3
3 weeks 3.8
50 PLA/Si/5

14 weeks 7 weeks

Storage Modulus (GPa)


3.6
40
Strees (MPa)

23 weeks
3.4
30
PLA/Si/3 3.2
20
3.0

10
2.8

0
0.08 2.6
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07
0 5 10 15 20 25 30 35
Strain
weeks

Figure 12. Tensile stress–strain curves for PLA/Si/3 at various Figure 13. Storage modulus at various aging times for PLA and
aging conditions. PLA/nanocomposites.

affected, with increasing aging time. Finally, after 23 present study, the degradation under immersion in a
weeks of aging, the Young’s modulus reduction for buffer solution at 37 C is enhanced, compared to
PLA is about 11%, while it is about 23% for PLA/ 40 C and 80% RH, which is also related to the
nanocomposites. Therefore, it can be extracted that higher hydrolytic degradation rate, calculated in the
silica enhances degradability of PLA, while this previous paragraph.
enhancement slows down with increasing aging time, It can be concluded that silica nanoparticles acceler-
probably due to the crystallinity increment for PLA/ ate the degradability of PLA, having a higher impact on
nanocomposites. Similar trend is obtained for the Young’s modulus, under the specified aging conditions,
yield stress variation. Initially, PLA/nanocomposites for 7 weeks and hereafter this acceleration is retarded,
have a slightly yield stress increment (of the order of due to the crystallinity increment, as a result of the
7%) compared to that of PLA matrix. With varying molecular weight reduction.
aging time, the yield stress of all samples is gradually
decreased, with the PLA/nanocomposites demonstrat-
ing higher decrement (up to 37% for PLA/Si/3). On the
DMA
other hand, after 18 weeks of aging, yield stress of In Figure 13 the storage modulus at a frequency of 1 Hz
the nanocomposites turns to be lower than that of the and 37 C is depicted with varying aging time. From this
matrix. It can be extracted that in this case, the entire plot it is revealed that the storage modulus is always
load is carried by the polymeric matrix and the load- higher for PLA/nanocomposites compared to that of
bearing cross-section is lowering with increasing PLA matrix, while the reinforcing effect is higher for
nanofiller content.28 This effect is an indication that PLA/Si/2, probably due to the finer dispersion of nano-
the reinforcing effect, observed initially, is no longer particles. The evolution of storage modulus with aging
active, revealing that degradation is more intense in time is monotonically decreasing, appearing to be dif-
the matrix–nanofiller interface. This effect is in agree- ferent than that of the Young’s modulus. This is due to
ment with SEM results, where the degradation appears the different type of deformation applied. The slope of
to be more intense around nanoparticles. In our previ- the storage modulus decrement is lower for PLA
ous study,19 with the same materials system, and differ- matrix, while PLA/Si/5 exhibits the higher slope of deg-
ent aging conditions, namely 40 C and 80% RH, yield radation compared to PLA/Si/2 and PLA/Si/3.
stress of PLA/Si nanocomposites was generally higher
than that of the PLA matrix. Comparing the two dif-
ferent aging conditions, it is revealed that at 40 C and
Conclusions
80% RH, the mechanical properties substantially A series of biobased polymer matrix nanocomposites
decrease in the first 3 weeks and hereafter they are based on PLA with varying nanosilica content have
recovering, while under immersion in a buffer solution been prepared and studied experimentally. A 39%
at 37 C, the same properties after the exhibited decre- Young’s modulus enhancement has been found for
ment in 3 weeks are slightly recovered and remain PLA/nanocomposites with respect to the pure PLA
almost stable for a long time period of aging. It must matrix. In addition, 33% crystallinity increment in
also be noted that PLA exhibits a lower degradability in average has been detected for PLA/nanocomposites,
all cases, compared to the PLA/nanocomposites. In the indicating that nanosilica fillers act as nucleating

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
Georgiopoulos et al. 673

agents. Hereafter, this enhancement does not change 3. Pillin I, Montrelay N, Bourmaud A, et al. Effect of
with varying nanofiller loading. thermo-mechanical cycles on the physic-chemical proper-
The cell culture investigation demonstrates that both ties of poly(lactic acid). Polym Degrad Stab 2008; 93:
materials PLA and the PLA/nanocomposites facilitate 321–328.
4. Saha SK and Tsuji H. Effects of molecular weight and
a good initial cell adhesion, increase preosteoblastic cell
small amounts of D-lactide units on hydrolytic degrad-
proliferation after 3 and 7 days in culture, and therefore ation of poly(L-lactic acid)s. Polym Degrad Stab 2006; 91:
support their potential use in bone tissue engineering 1665–1673.
applications. 5. Ho KLG, Pometto AL and Hinz PN. Effects of tempera-
A study of the change of the thermomechanical ture and relative humidity on polylactic acid plastic deg-
properties of PLA and PLA/nanocomposites with radation. Polym Environ 1999; 7: 83–92.
aging time after exposure at a buffer solution at a tem- 6. Middleton JC and Tipton AJ. Synthetic biodegradable
perature of 37 C with a pH of 7.4, simulating the polymers as orthopedic devices. Biomaterials 2000; 21:
human body, has been performed. Comparative study 2335–2346.
with various aging conditions, previously executed has 7. Felfel RM, Ahmed I, Parsons AJ, et al. Investigation of
also been made. Significantly lower molecular weights crystallinity, molecular weight change and mechanical
properties of PLA/PBG bioresorbable composites as
for PLA matrix were measured for samples after 23
bone fracture fixation plates. J Biomater Appl 2012; 26:
weeks of aging. The aged samples display a higher con-
765–789.
tribution of low molecular weight fractions, which is 8. Fukushima K, Abbate C, Tabuani D, et al.
attributed mainly to chain scission due to hydrolysis. Biodegradation of poly(lactic acid) and its nanocompo-
The crystallization behavior in PLA and the way it is sites. Polym Degrad Stab 2009; 94: 1646–1655.
affected by the presence of nanofillers during degrad- 9. Ray SS and Bousmina M. Biodegradable polymers and
ation procedure has been studied and values of 44% their layered silicate nanocomposites: In greening the 21st
crystallinity increment have been found. Double melt- century materials world. Prog Mater Sci 2005; 50:
ing endotherms at 146 and 156 C, associated with dif- 962–1079.
ferent morphologies have been detected for all material 10. MacNeill I and Leiper H. Degradation studies of some
types studied. The high temperature melting peak was polyesters and polycarbonates-2. Polylactide:
Degradation under isothermal conditions, thermal deg-
more intense with increasing aging time, and this effect
radation mechanism and photolysis of the polymer.
has been attributed to a recrystallization procedure fol-
Polym Degrad Stab 1985; 11: 309–326.
lowed by a subsequent melting. 11. Feijoo J, Cabedo L, Gimenez E, et al. Development of
At the particular aging conditions studied, silica amorphous PLA-montmorillonite nanocomposites.
nanoparticles accelerate the degradability of PLA, J Mater Sci 2005; 40: 1785–1788.
having a higher impact on Young’s modulus, under 12. Jungraithmayr W, Laube I, Hild N, et al. Bioactive nano-
the specified aging conditions, for 7 weeks and hereafter composite for chest-wall replacement: Cellular response
this acceleration is retarded, due to the crystallinity in amurine model. J Biomater Appl 2014; 29: 36–45.
increment, as a result of the molecular weight reduction. 13. Wen X, Lin Y, Han C, et al. Thermomechanical and
optical properties of biodegradable poly(L-lactide)/silica
Declaration of conflicting interests nanocomposites by melt compounding. J Appl Polym Sci
2009; 114: 3379–3388.
None declared. 14. Ray SS, Maiti P, Okamoto M, et al. New polylactide/
layered silicate nanocomposites. 1. Preparation,
Funding characterization, and properties. Macromolecules 2002;
This research has been co-financed by the European Union 35: 3104–3110.
(European Social Fund, ESF) and Greek national funds 15. Delabarde C, Plummer CJG, Bourban PE, et al.
through the Operational Program ‘‘Education and Lifelong Accelerated ageing and degradation in poly-L-lactide/
Learning’’, Research Funding Program Aristeia (E.K.), and hydroxyapatite nanocomposites. Polym Degrad Stab
Aristeia II ‘Osteobiomimesis’ (MC). 2011; 96: 595–607.
16. Zhao Y, Qiu Z and Yang W. Effect of functionalization
of multiwalled nanotubes on the crystallization and
References hydrolytic degradation of biodegradable poly(L-lactide).
J Phys Chem B 2008; 112: 16461–16468.
1. Niaounakis M, Kontou E and Xanthis M. Effects of aging 17. Ray SS, Yamada K, Okamoto M, et al. New polylactide-
on the thermomechanical properties of poly(lactic acid). layered silicate nanocomposites. 2. Concurrent improve-
J Appl Polym Sci 1966; 119: 472–481. ments of material properties, biodegradability and melt
2. Signori F, Coltelli M-B and Bronco S. Thermal
rheology. Polymer 2003; 44: 857–866.
degradation of poly(lactic acid) (PLA) and poly(butylene 18. Chen H, Wang Y, Chen J, et al. Hydrolytic degradation
adipate-co-terephthalate) (PBAT) and their blends upon behavior of poly(L-lactide)/SiO2 composite. Polym
melt processing. Polym Degrad Stab 2009; 94: 74–82.
Degrad Stab 2013; 98: 2672–2679.

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014
674 Journal of Biomaterials Applications 29(5)

19. Kontou E, Georgiopoulos P and Niaounakis M. The role FTIR spectroscopy and differential scanning calorimetry.
of nanofillers on the degradation behavior of polylactic Polym Degrad Stab 2009; 94: 1364–1372.
acid. Polym Compos 2012; 33: 282–294. 30. Georgiopoulos P, Kontou E and Niaounakis M.
20. Chen H, Chen J, Chen J, et al. Effect of organic mont- Thermomechanical properties and rheological behavior
morillonite on cold crystallization and hydrolytic degrad- of biodegradable composites. Polym Compos 2014; 35:
ation of poly(L-lactide). Polym Degrad Stab 2012; 97: 1140–1149.
2273–2283. 31. Chen HM, Feng CX, Zhang WB, et al. Hydrolytic deg-
21. Paul MA, Delcourt C, Alexandre M, et al. Polylactide/ radation behavior of poly(L-lactide)/carbon nanotubes
montmorillonite nanocomposites: Study of the hydrolytic nanocomposites. Polym Degrad Stab 2013; 98: 198–208.
degradation. Polym Degrad Stab 2005; 87: 535–542. 32. Zhang TP, Hu J, Duan YX, et al. Physical aging enhanced
22. Fukushima K, Tabuani D, Dottori M, et al. Effect of mesomorphic structure in melt-quenched poly(L-lactic
temperature and nanoparticle type on hydrolytic degrad- acid). J Phys Chem B 2011; 115: 13835–13841.
ation of poly(lactic acid) nanocomposites. Polym Degrad 33. Pluta M, Murariu M, Alexandre M, et al. Polylactide
Stab 2011; 96: 2120–2129. compositions. The influence of ageing on the structure,
23. Zhang X, Espiritu M, Bilyk A, et al. Morphological thermal and viscoelastic properties of PLA/calcium
behavior of poly(lactic acid) during hydrolytic degrad- sulfate composites. Polym Degrad Stab 2008; 93:
ation. Polym Degrad Stab 2008; 93: 1964–1970. 925–931.
24. Terzaki K, Kissamitaki M, Skarmoutsou A, et al. 34. Tsuji H, Ikarashi K and Fukuda N. Poly(L-lactide): XII.
Pre-osteoblastic cell response on three-dimensional, Formation, growth, and morphology of crystalline resi-
organic-inorganic hybrid material scaffolds for bone dues as extended-chain crystallites through hydrolysis of
tissue engineering. J Biomed Mater Res Part A 2013; poly(L-lactide) films in phosphate-buffered solution.
101A: 2283–2294. Polym Degrad Stab 2004; 84: 515–523.
25. Fischer EW, Sterzel HJ and Wegner G. Investigation of 35. Huang JW, Hung YC, Wen YL, et al. Polylactide/nano-
the structure of solution grown crystals of lactide copoly- and microscale silica composite films. II. Melting behav-
mers by means of chemical reactions. Colloid Polym Sci ior and cold crystallization. J Appl Polym Sci 2009; 112:
1973; 251: 980–990. 3149–3156.
26. Kontou E and Farasoglou P. Determination of the true 36. Shieh YT and Liu GL. Effects of carbon nanotubes on
stress-strain behaviour of polypropylene. J Mater Sci crystallization and melting behavior of poly(L-lactide) via
1998; 33: 147–153. DSC and TMDSC studies. J Polym Sci Part B Polym
27. Dacko P, Kowalczuk M, Janeczek H, et al. Physical Phys 2007; 45: 1870–1881.
properties of the biodegradable polymer compositions 37. Chen HM, Zhang WB, Du XC, et al. Crystallization kin-
containing natural polyesters and their synthetic ana- etics and melting behaviors of poly(L-lactide)-graphene
logues. Macromol Symp 2006; 239: 209–216. oxides composites. Thermochim Acta 2013; 566: 57–70.
28. Dorigato A, Sebastiani M, Pegoretti A, et al. Effect of 38. Ray SS, Yamada K, Okamoto M, et al. Polylactide-
silica nanoparticles on the mechanical performance of layered silicate nanocomposite: A novel biodegradable
poly(lactic acid). J Polym Environ 2012; 20: 713–725. material. Nano Lett 2002; 2: 1093–1096.
29. Vasanthan N and Ly O. Effect of microstructure on
hydrolytic degradation studies of poly(L-lactic acid) by

Downloaded from jba.sagepub.com at UCSF LIBRARY & CKM on November 28, 2014

You might also like