You are on page 1of 9

Polymer Degradation and Stability 94 (2009) 13641372

Contents lists available at ScienceDirect

Polymer Degradation and Stability


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

Effect of microstructure on hydrolytic degradation studies of poly (L-lactic acid)


by FTIR spectroscopy and differential scanning calorimetry
Nadarajah Vasanthan*, Onah Ly
Department of Chemistry, Long Island University, One University Plaza, Brooklyn, NY 11201, USA

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 9 February 2009
Received in revised form
14 May 2009
Accepted 19 May 2009
Available online 28 May 2009

Structural changes during thermally induced crystallization and alkaline hydrolysis of Poly(L-lactic acid)
(PLLA) lms were investigated using differential scanning calorimetry (DSC), FTIR spectroscopy, weight
loss, HPLC and optical microscopy. It was shown that crystallinity (cc), glass transition temperature (Tg)
and melting temperature (Tm) were found to be strongly annealing temperature (Ta) dependent. The FTIR
study of PLLA lms suggested that the bands at 921 and 956 cm1 could be used to monitor the structural
changes of PLLA. An independent infrared spectroscopic method was developed for the rst time to
determine crystallinity of PLLA before degradation and it showed good qualitative correlation with DSC
crystallinity. The higher crystallinity values determined by FTIR were attributed to the intermediate
phase included in the IR crystallinity. Both the weight loss data and the percentage of lactic acid obtained
by HPLC showed that the alkaline hydrolysis of PLLA lms increased with increasing crystallinity. The
DSC observation showed an increase in Tg and no signicant change in Tm and heat of fusion, while IR
showed an increase in IR crystallinity with increasing hydrolysis time. The increase in IR crystallinity and
Tg with hydrolysis time suggested that degradation progressed from the edges of the crystalline lamellas
without decreasing lamellar thickness, but increased the intermediate phase and the short-range order.
2009 Elsevier Ltd. All rights reserved.

Keywords:
Hydrolytic degradation
FTIR spectroscopy
Crystallinity
Crystallization

1. Introduction
Polylactic acid (PLA) is a biodegradable polymer and it is widely
studied and used in the area of industrial packaging and biomedical
applications such as resorbable sutures, surgical implants, scaffolds
for tissue engineering and controlled drug-delivery devices [15].
However, it has been demonstrated that these systems are not
suitable for hard tissue regeneration due to its weak mechanical
properties [610]. PLA can exist as two stereoisomers, designated as
D and L, or as a racemic mixture, designated as DL. The D and L forms
are optically active while the DL form is optically inactive. Poly
(L-lactic acid) (PLLA) and poly(D-lactic acid) (PDLA) are semicrystalline while poly(DL-lactic acid) (PDLLA) is amorphous [2,11,12].
PLA has a glass transition temperature (Tg) of about 60  C and
crystalline melting temperature (Tm) in the range from 130 to 180  C.
[6] Tg and Tm depend on optical composition, thermal history and
molecular weight [1315]. Pure PDLA or PLLA has an equilibrium
crystalline melting temperature (Tmo) of 207  C [1619]. Baratian et al.
carried out the isothermal crystallization of a series of poly
(L-lactide-co-D-lactide) [20]. Kolstad [21] studied the crystallization

* Corresponding author. Tel.: 1 718 246 6328; fax: 1 718 488 1465.
E-mail address: nadarajah.vasanthan@liu.edu (N. Vasanthan).
0141-3910/$ see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.polymdegradstab.2009.05.015

kinetics of poly(L-co-meso-lactide) and found that the crystallization


half time increased approximately 40% for every 1 wt% increase in the
meso-lactide. PLLA crystallizes into three different crystal modications, a, b and g form, depending on the crystallization conditions.
However, the a and b forms are the common ones [22]. The a form is
formed either by melt or cold crystallization of PLLA while the b form
is produced during either drawing of PLLA lms at a high draw ratio
and high drawing temperature or spinning bers at a high spinning
speed. The g form is produced by epitaxial crystallization [22]. De
Santis and Kovacs showed that the a form of PLLA from solution-spun
bers has a pseudo-orthorhombic unit cell containing two lefthanded 103 polymeric helices which are arranged in an antiparallel
fashion. The unit cell dimensions of a crystal form are a 1.07 nm,
b 0.645 nm and c 2.78 nm. Hoogsteen et al. showed that the
b form of PLLA adopts an orthorhombic unit cell containing six
left-handed 31 polymeric helices. The unit cell dimensions of b form
are a 1.031 nm, b 1.821 nm and c 0.900 nm [22,23].
Polymer degradation means the changes in physical properties
due to the chemical bond scission reactions in macromolecules
[24,25]. Polymer degradation can be induced by thermal activation,
hydrolysis, biological activity (i.e. enzymes), oxidation or photolysis. Depending on the mode of initiation, polymer degradation
can be classied as thermal, mechanical, photochemical, biological
or chemical degradation [25]. In addition to the environmental

N. Vasanthan, O. Ly / Polymer Degradation and Stability 94 (2009) 13641372

conditions such as moisture, temperature and microorganisms,


polymer degradation also depends on the structural properties
of the polymer. Chain orientation, stereochemical composition,
degree of crystallinity, molecular weight, molecular weight distribution and degree of crosslinking of polymers are among the
important ones [2628].
Hydrolysis belongs to the class of chemical degradation, which
occurs by scission of chemical bond in the main chain by reaction
with water. The mechanism associated with hydrolysis of ester
linkage in neutral or acidic media is different from the one in
alkaline media [29]. In neutral or acidic media (Scheme 1), the
hydrolysis is initiated by protonation [29] and is followed by the
addition of water and the cleavage of the ester linkage. On the other
hand, in alkaline media (Scheme 2) hydroxyl ions are attached to the
carbonyl carbons and followed by the breaking of the ester linkages.
Hydrolytic degradation of PLLA was reported to occur either by
surface erosion or bulk erosion mechanism. Surface erosion mechanism of polymer degradation occurs only at the polymerwater
interface. Bulk erosion mechanism is the uniform degradation
throughout the polymer [26,30]. The mechanism associated with
alkaline hydrolysis is still not well understood and therefore in the
present study, the effect of crystallinity and morphology on hydrolytic degradation and microstructure changes with degradation was
investigated using weight loss, HPLC, DSC, optical microscopy and
FTIR spectroscopy.
2. Experimental

1365

OH

OH
O

O
C

OH

HO

OH
O
C

Scheme 2. Hydrolysis reactions of ester linkage in alkaline media.

2.3. Hydrolysis
Alkaline hydrolysis was performed on both annealed and
as-prepared PLLA lms. The PLLA lms were annealed at 80  C and
110  C in the oven for 30 min. The alkaline hydrolysis of each PLLA
lm (1 in  1 in) was performed in a 100 mL beaker containing 50 mL
of 0.1 M NaOH solution at room temperature for 1, 2, 4, 6, 8, 10 and 12
days. After hydrolysis, the PLLA lms were removed and washed
with ice cold water to remove any traces of NaOH solution to
stop further hydrolysis. Then the PLLA lms were dried at room
temperature and kept in a desiccator at constant weight until further
use. These PLLA lms were used to study hydrolysis behavior. The
hydrolyzed solutions were used for HPLC studies.

2.1. Materials
Poly(L-lactic acid) (PLLA) with 6%D content was provided by
Professor Dennis Smith from Clemson University. The Mn and Mw of
this PLLA are 92,000 and 225,000 Da, respectively. Its polydispersity
index (PDI) is 2.44. NaOH solution of 0.1 M was obtained from Fisher
Scientic. Sodium lactate solution of 60% (w/w) with a density of
1.3 g/mL (C3H5O3Na) was purchased from SigmaAldrich. All of
these materials were used without further purication.
2.2. Preparation of PLLA lm
PLLA lms were prepared by melt-pressing technique with
a Carver press. The press was preheated to 200  C and the PLLA
pellets were placed between the two heated platens and allowed to
heat to 200  C for 5 min. Pressure of 20,000 pounds was applied on
the sandwich for about 4 min. The lm was then removed and
quenched quickly in ice cold water to prepare fully amorphous lm
(thickness is about 4050 m). The amorphous PLLA lm was dried at
room temperature and kept in a desiccator for further studies.
These amorphous PLLA lms were then annealed at different
temperatures from 80  C to 120  C with the increment of 10  C for
30 min for hydrolysis studies.

OH

O
C

OH
C

OH
O

OH
O

+ H2O

+ HO

Scheme 1. Hydrolysis reactions of ester linkage in neutral or acidic media.

2.4. Weight loss measurements


The percentage weight loss of hydrolyzed lms was calculated
from the weights of the dried PLLA lms before and after hydrolysis
according to the following equation:

.

Wbefore
Wloss % 100  Wbefore  Wafter
where Wloss(%) is the percentage weight loss of hydrolyzed PLLA
lm.
Wbefore is the weight of the dried PLLA lm before hydrolysis.
Wafter is the weight of the dried PLLA lm after hydrolysis.
2.5. High-performance liquid chromatography (HPLC) study
The chromatograms of all standard lactic acid and hydrolyzed
samples were obtained using the HewlettPackard 1050 HPLC
system consisting of diode array detector (DAD) and 250  4.6 mm
Phenomenex and Maxsil 5 RP 18 column. The 0.012 M HCl mobile
phase was run with the ow rate of 1.0 mL/min; the injection
volume was 20 mL. The hydrochloric acid solution with concentration of 0.012 M was used as the mobile phase for HPLC. This solution was prepared by diluting the 12.1 M concentrated hydrochloric
acid solution with E-pure water with the ratio of 1:1000 by volume.
Preparation of standard lactic acid solution and hydrolyzed
solutions were carried out by the following procedures. Lactic acid
solution of 13.9 mM stock was prepared by diluting 200 mL of
6.96 M sodium lactate with E-pure water, adjusted to pH 2.10 using
1 M hydrochloric acid (HCl) solution, and brought to the mark of
a 100 mL volumetric ask with E-pure water. A series of ve
standard lactic acid solutions was prepared by diluting 13.9 mM
stock lactic acid solution with E-pure water with the ratio of 1:2,
1:5, 1:10, 1:20 and 1:50 by volume for obtaining the concentration
of 6.950, 2.780, 1.390, 0.695 and 0.278 mM, respectively. Standard
solutions were placed in capped auto injector vials for HPLC

1366

N. Vasanthan, O. Ly / Polymer Degradation and Stability 94 (2009) 13641372

analysis. Each of the hydrolyzed samples was prepared and placed


in capped auto injector vials for HPLC analysis. One-day and twoday samples were prepared by dilution with 0.1 M HCl solutions
with a ratio of 1:5 by volume. All the other samples (4d, 6d, 8d, 10d,
and 12d) were prepared by dilution with 0.1 M HCl solutions with
a ratio of 1:20 by volume. 1d, 2d, 4d, 6d, 8d, 10d and 12d samples
represent the hydrolyzed solutions after removing the PLLA lm,
which was immersed in 0.1 M NaOH solution for 1, 2, 4, 6, 8, 10 and
12 days, respectively.
2.6. Differential scanning calorimetry (DSC)

2.7. Fourier transform infrared spectroscopy (FTIR)


Infrared spectra of the annealed and hydrolyzed PLLA lms were
collected using a Nicolet Magna 760 spectrometer. The spectra were
collected using transmission mode with a resolution of 4 cm1
and 64 scans per sample in the mid-IR region of 4000500 cm1.
Grams software was used for peak tting. The linear baseline and
Lorentzian shape were used for all peak tting.
3. Results and discussion
3.1. Crystallization study
Fig. 1 shows DSC scans of melt-pressed PLLA lm and PLLA lms
annealed at 80  C and 110  C. All DSC scans show two transitions,
which are glass transition, Tg and melting temperature, Tm. It can be
seen that both Tg and Tm increase substantially with increasing Ta.
Double-melting endotherms named as low and high melting peaks
were observed for all annealed PLLA lms and they can be attributed
to either recrystallization of metastable crystals upon annealing or
crystallization from the amorphous region or a combination of both. It
should be noted from Fig. 1 that a strong endothermic transition
similar to melting transition occurred for Tg . Similar observation has
been made by number of groups and it was attributed to enthalpy of
stress relaxation [31]. Fig. 2 shows Tg and Tm of PLLA as a function of
annealing temperatures. Both Tg and Tm observed for annealed PLLA
lms increase with increasing Ta. Tg of melt-crystallized PLLA was
studied recently by Fitz et al. [32] and shown that a signicant
reduction in Tg occurs during crystallization of PLLA under constrained
conditions. Our PLLA samples were annealed under unconstrained
conditions and thus exhibit the conventional trend of increasing Tg
with increasing Ta. It should be pointed out that our PLLA samples have
lower melting points than previously reported values [33,34], this may
be attributable to the few D-lactic acid units incorporated in our PLLA
sample. Fig. 3 shows the variation in crystallinity of PLLA lms as
a function of Ta. The crystallinity values show an increase with
increasing Ta, which is consistent with previously reported data [35]. It
can be seen that the crystallinity increases slowly up to Ta 80  C and
increases rapidly for Ta from 80  C to 110  C. It is also apparent that

Fig. 1. DSC scans of PLLA at annealing temperature (Ta) of RT, 80  C and 110  C.

crystallinity increase becomes smaller when Ta exceeds 110  C. This


may be attributable to the slower crystallization rate at higher
temperatures that reduces the amount of nal crystallinity due to
insufcient crystallization within a given period of time.
In this study, crystallization of PLLA during annealing was
followed by FTIR spectroscopy. Fig. 4 shows the FTIR spectra of PLLA
lms annealed at different Ta from 80  C to 120  C in the regions of
13201280and 1000600 cm1. The structural changes were
investigated by comparing the spectral differences between semicrystalline and amorphous PLLA. It is obvious that the IR spectrum of
semicrystalline and IR spectrum of amorphous PLLA have distinct
differences. Amorphous spectra show a band at 1302 cm1 that splits
into two bands at 1302 and 1293 cm1 for the lms annealed at Ta
from 100 to 120  C. It appears that the band ratio of 1293 cm1/
1302 cm1 increases with increasing Ta. It should be noted that the
band at 956 cm1 decreases in intensity while the band at 921 cm1
increases in intensity with increasing Ta. The band at 921 cm1 was
attributed previously to the combination of CC backbone and CH3
rocking mode of PLLA a crystals [36,37]. The bands at 895 and
871 cm1 become weaker as Ta increases. The bands at 757 and
710 cm1 appear as a single band in the IR spectrum of melt-pressed
lm and both bands split into two bands as Ta exceeds 100  C. It
is also apparent that the bands at 739 and 697 cm1 increase in

200
Tg
Tm
Linear (Tm)
Linear (Tg)

180
160

Tg/Tm (C)

Tg, Tm and DH of annealed PLLA lms before and after hydrolysis


were determined using the Perkin Elmer DSC 7 differential scanning
calorimeter. The instrument was calibrated for temperature and heat
of fusion using standard indium (Tm 156.6  C and DH 28.5 J/g).
All experiments were performed under nitrogen atmosphere with
the ow rate of 20 mL/min. The samples were prepared by sealing in
aluminum pan with the weight of 46 mg. DSC measurements were
performed using the following procedure: The samples were held
for 5 min at 25  C and subsequently heated from 25  C to 200  C at
10  C/min. Heat of fusion is directly proportional to crystallinity (cc),
and it was calculated according to the following equation, cc DHs/
DHo, where DHo is the theoretical heat of fusion for 100% crystalline
PLLA and it was taken as 93 J/g [30].

140
120
100
80
60
40
20

20

40

60

80

100

120

Ta (C)
Fig. 2. Tg and Tm of PLLA lms as a function of annealing temperature.

140

N. Vasanthan, O. Ly / Polymer Degradation and Stability 94 (2009) 13641372

35

1367

Crystallinity (%)

30
25
20
15
10
5

20

40

60

80

100

120

140

Annealing Temperature (C)

intensity with increasing Ta. In summary, increases in absorbance of


bands at 697, 739, 921 and 1293 cm1 are observed while decreases
in absorbance of bands at 710, 757, 895, 956 and 1302 cm1 are
observed with increasing Ta. It appeared that there is no signicant
change in the absorbance of the CH stretching region. It is well
known that annealing of PLLA lms produce a signicant amount of
crystallinity and therefore the bands showing an increase in absorbance can be attributed to the crystalline phase of PLLA while the
bands showing a decrease in absorbance can be attributed to the
amorphous phase of PLLA.
In order to conrm the reference band, curve tting was carried
out to separate different components in the CH stretching region
from 3060 to2840 cm1. Five Lorentzian peaks were found to t the
region, shown in 5a. Five peaks at 2996, 2961, 2945, 2901 and
2880 cm1 were obtained. Fig. 5b shows the peak areas of three
bands (2996, 2961 and 2945 cm1) of PLLA lms as a function of Ta.
No signicant changes in peak area for all PLLA lms as a function of
Ta were observed and therefore in the present study, absorbance of
the bands at 2996, 2961 and 2945 cm1 were used as reference
bands to monitor structural changes of PLLA during annealing. If we

50

40

Band Area

Fig. 3. Crystallinity of PLLA lms obtained by DSC as a function of annealing


temperature.

30

20

10

70

80

90

100

110

120

130

140

Temperature (C)
Fig. 5. a. Curve tting in the CH stretching region. b. Peak areas of three bands at
2996, 2961 and 2945 cm1versus annealing temperature.

assume PLLA satises a two-phase model, these bands can be used


to measure the crystallinity by calibration against another method
such as DSC measurement. Since we know the bands are associated
with the crystalline and amorphous phases, respectively, the band

0.8
0.7

Band Area Ratio

0.6
0.5
0.4
0.3
0.2
0.1
0.0
0

10

20

30

40

50

DSC Crystallinity (%)


Fig. 4. FTIR spectra of PLLA lms at various annealing temperatures in the region of
1000600 cm1: (a) 80  C, (b) 90  C, (c) 100  C, (d) 110  C and (e) 120  C.

Fig. 6. Plot of absorbance ratio (peak area) of crystalline and amorphous bands (A921/
A956) versus DSC crystallinity (%) of the same PLLA lm.

1368

N. Vasanthan, O. Ly / Polymer Degradation and Stability 94 (2009) 13641372

50

440

FTIR crystallinity (%)

400

ARef/A956

360
320
280
240

40

30

20

10

200
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

10

20

30

40

DSC Crystallinity (%)

A921/A956
Fig. 7. Plot of ARef/A956 versus A921/A956 of PLLA lms at different annealing temperatures where ARef is the absorbance of reference band.

Fig. 8. FTIR crystallinity (%) versus DSC crystallinity (%).

ratios of 921 and 956 cm1 can be correlated with the crystallinity
obtained from DSC in Fig. 6. It is apparent from Fig. 6 that (A921/
A956) increases with increasing DSC crystallinity. Crystallinity of an
unknown sample may then be obtained from this calibration plot.
This plot will be used to estimate the crystallinity change with
hydrolytic degradation; this will be discussed later.
It would be very advantageous to develop a new method of
determining crystallinity using FTIR that does not involve any
calibration. FTIR spectroscopy was used for the rst time to quantify
the amount of crystalline and amorphous fraction of PLLA with
increasing Ta and it was correlated with the crystalline fraction
values obtained by DSC measurement. The bands at 921 and
956 cm1 are used to represent the crystalline and amorphous
phase, respectively. If we assume PLLA satises a two-phase model, these bands can be used to measure the distribution of the
crystalline and amorphous components. We assume a two-phase
crystalline and amorphous model for PLLA as follows:

different Ta and DSC crystallinity. The results show the sum of


crystalline and amorphous fraction obtained by IR to be 1.00  0.05
for all PLLA lms at different Ta. Crystallinity obtained by FTIR versus
crystallinity obtained by DSC of PLLA lms at different Ta is plotted in
Fig. 8; the extrapolation of the plot to zero IR crystallinity yielded the
intercept at 6.5% instead of 0% due to the differences of crystallinity
obtained from FTIR and DSC studies. It is not surprising to see the
crystalline fraction obtained by FTIR is higher than crystalline fraction
obtained by DSC. It is known for sometime that FTIR always provides
higher crystallinity values than DSC because DSC measures the
meltable portion of the crystals while FTIR measures the short-range
order that includes crystalline phase as well as the intermediate
phase.

P1 A921 =ARef



P2 A956 =ARef 1

or ARef =A956 P1 A921 =A956 P2

(1)

where P1 and P2 are constants associated with crystalline and amorphous bands, respectively. The band ratio of reference and amorphous
area (ARef/A956) versus band ratio of crystalline and amorphous area
(A921/A956) of PLLA lms at different Ta is plotted in Fig. 7 and shows
a good linear t. P1 and P2 were found to be 272.9 and 217.5, respectively. Crystalline fraction (P1(A921/ARef)) and amorphous fraction
(P2(A956/ARef)) were obtained for PLLA lms at different Ta; the results
are shown in Table 1. Table 1 also includes the sum of crystalline
fraction and amorphous fraction for all PLLA lms annealed at

Table 1
Crystalline fraction, amorphous fraction, sum of crystalline fraction and amorphous
fraction of PLLA lms obtained by FTIR along with crystalline fraction obtained by
DSC at different Ta. Standard deviation is / 0.05 based on at least four
measurements.
Annealing
temperature
( C)

Crystalline
fraction
by DSC

Crystalline
fraction
by FTIR

Amorphous
fraction
by FTIR

Sum of crystalline
and amorphous
fraction

80
90
100
110
120

0.12
0.16
0.20
0.32
0.33

0.07
0.15
0.31
0.40
0.47

0.88
0.91
0.70
0.56
0.55

0.95
1.06
1.01
0.96
1.03

The percentage of weight loss of PLLA lms annealed at different


Ta was calculated from dried PLLA lms before and after the
hydrolysis as described in the Experimental section. A PLLA lm
was placed in distilled water for a week and no signicant change in
the mass of PLLA lm was shown. Annealing of PLLA lms were
carried out to obtain lms with varying crystallinity. All PLLA lms
were placed in water for a week and no change in crystallinity was

100

80

Weight Loss (%)

3.2. Hydrolytic degradation

60

40

20

0
0

10

12

Hydrolysis Time (days)


Fig. 9. Percentage weight loss as a function of hydrolysis time for PLLA lms annealed
at different temperatures (C) PLLA lm as-prepared; (B) PLLA lm annealed at 80  C;
(:) PLLA lm annealed at 110  C.

N. Vasanthan, O. Ly / Polymer Degradation and Stability 94 (2009) 13641372

much constant while the weight loss of PLLA lms with crystallinity
of 12.4% and 31.5% increased up to 10 days during hydrolytic
degradation. Alkaline and enzymatic hydrolysis of PLA copolymers
was reported recently30 and it has been shown that alkaline
hydrolysis of copolymers depends on initial crystallinity of the lm
and not the amount of the co-monomer unit. On the other hand,
enzymatic degradation depends on both initial crystallinity and the
co-monomer unit. It has also been shown that both alkaline and

1369

600
500

Peak Area

400
300
200
100
0

Concentration (mM)
Fig. 10. a. Chromatogram of 2.780 mM standard lactic acid solution, Fig. 10b.
Calibration curve of standard lactic acid solutions.

observed. This conrms that there is no additional crystallization


occurred from the free amorphous region during hydrolysis. The
percentage of weight loss as a function of hydrolysis time for all
PLLA lms with different crystallinity is shown in Fig. 9. It appears
that the percentage of weight loss increases with hydrolysis time
for all PLLA lms studied and increases with initial crystallinity of
the PLLA lm. Fig. 9 shows that weight loss of PLLA lms with
crystallinity of 9.8%, increased up to 6 days and then stayed pretty
80

Weight % Lactic Acid

70
60
50
40
30
20
10
0
0

10

12

14

Time (days)
Fig. 11. Weight percentage of lactic acid as a function of hydrolysis time for PLLA lms
annealed at different temperatures (C) PLLA lm as-prepared; (B) PLLA lm annealed
at 80  C; (:) PLLA lm annealed at 110  C.

Fig. 12. a. DSC scans of PLLA lms as prepared before and after hydrolysis, Fig. 12b. DSC
scans of PLLA lms annealed at 80  C before and after hydrolysis, Fig. 12c. DSC scans of
PLLA lms annealed at 110  C before and after hydrolysis.

1370

N. Vasanthan, O. Ly / Polymer Degradation and Stability 94 (2009) 13641372

Table 2
The values of DH, Tg and Tm of PLLA lms before and after hydrolysis for the sample annealed at RT, 80  C and 110  C.
Hydrolysis time (day)

0
2
6
10

80  C

RTa

DH DH b (J/g)

Tg c ( C)

Tm

9.1
7.2
9.6
10.3

59.1
60.5
61.2
61.2

140.0
141.0
140.9
140.8

d 

( C)

110  C

DH (J/g)

Tg ( C)

Tm ( C)

DH (J/g)

Tg ( C)

Tm ( C)

11.6
9.2
9.2
9.9

61.6
63.4
63.7
65.3

144.6
145.1
144.9
145.0

29.3
28.6
26.6
27.2

62.3
63.2
64.6
65.2

150.9
150.1
149.4
149.1

Condence limits for temperatures (/ 1  C) and DH (/0.5 J/g) based on three replicates.
a
Room temperature about 25  C.
b
Heat of fusion.
c
Glass transition temperature.
d
Melting temperature.

enzymatic hydrolyzibility of the PLA lm decreased with increasing


crystallinity, which is the opposite of what we have observed
during alkaline hydrolysis. The crystallinity range of the PLA
copolymer lms investigated was very small (2622%) while the
crystallinity of PLA lms used in our investigation was in the range
of 930%. That the maximum weight loss appeared to increase
with increasing crystallinity in alkaline hydrolysis suggests that the
mechanism associated with alkaline hydrolysis is different from the
enzymatic degradation reported previously.
In order to determine the amount of lactic acid in the hydrolyzed solutions, the calibration curve of standard lactic acid
solutions was constructed based on the result obtained from the
chromatograms of a series of ve standard lactic acid solutions
with different concentrations. Chromatogram observed for
a standard lactic acid solution with a concentration of 2.780 mM
shows only one peak with a retention time of 4.403 min, shown
in Fig. 10a. Therefore the retention time of 4.403 min was
attributed to lactic acid. Fig. 10 shows the calibration curve of
standard lactic acid solutions obtained by integrating the area
under the lactic acid peak of all standard lactic acid solutions.
Retention time for all standard lactic acid solutions was about
4.4 min. The calibration curve of the standard lactic acid solutions
in Fig. 10b revealed that the peak area of lactic acid linearly
increases with lactic acid concentration. The peak area of lactic
acid was obtained by integrating its area under the curve for all
hydrolyzed samples of PLLA annealed at RT, 80  C and 110  C.
Lactic acid concentrations in hydrolyzed samples were calculated
using the calibration curve of standard lactic acid solutions and
the peak areas of lactic acid peaks in hydrolyzed samples. Then
the percentages of lactic acid in hydrolyzed samples were calculated by dividing the weight of the lactic acid contained in the
hydrolyzed sample to the weight of PLLA lm before hydrolysis
and multiplied by 100. The percentage of lactic acid in hydrolyzed
samples obtained by HPLC as a function of hydrolysis time for the
PLLA samples with varying crystallinity is plotted in Fig. 11. It can
be seen that the percentage of lactic acid increases as hydrolysis
time increases. The percentage of lactic acid in the hydrolyzed
samples was found to decrease as the following for most of the
time: PLLA-110  C (Xc 31.5%) > PLLA-80  C (Xc 12.4%) > PLLART (Xc 9.8%). These results were in agreement with the results
obtained in our weight loss study.
3.3. Microstructure changes after degradation
Fig. 12ac show the DSC scans of PLLA lms annealed at RT, 80  C
and 110  C, respectively, before and after hydrolysis. The values of
DH, Tg and Tm for PLLA lms annealed at RT, 80  C and 110  C before
and after hydrolysis are tabulated in Table 2. It is clear from Table 2
that there was no signicant change in heat of fusion of PLLA lms
before and after hydrolysis. Table 2 also shows Tg of PLLA lms

before and after hydrolysis with respect to hydrolysis time. It is


apparent that Tg increases with increasing hydrolysis time for all
PLLA lms annealed at RT, 80  C and 110  C. It should also be noted
that there is no signicant change in Tm as a function of hydrolysis
time. Tm was found to be 140141  C for PLLA lms at RT,
144145  C for PLLA lms annealed at 80  C and 149151  C for
PLLA lms annealed at 110  C. The lack of signicant change in heat
of fusion as well as melting temperature suggests that the crystalline region is not affected by alkaline hydrolysis. The increase in
Tg may be attributed to increase in intermediate phase of PLLA lm
with degradation time that reduces the mobility of PLLA chain
which is supported by FTIR observation (it will be discussed later)
but not by DSC observation.
The optical micrographs of PLLA lms annealed at 110  C before
and after alkaline hydrolysis are given in Fig. 13. PLLA lm before
hydrolysis (Fig. 13a) shows spherulitic morphology. Fig. 13a illustrates typical regions in the PLLA lm and shows larger spherulites
surrounded by smaller spherulites. Three regions are apparent in
the optical micrographs: 1. crystalline region; 2. amorphous region
within the spherulite; and 3. free amorphous region between
the spherulites. Rotation of sample results in extinction patterns
suggests the presence of birefringence. Fig. 13b shows the
morphology of PLLA lm degraded for six days and, in Fig. 13b,
spherulitic boundaries are clearly visible. Close observation indicates some of the materials may have been etched away. Fig. 13c
illustrates the morphology of PLLA lm degraded for 10 days. It is
clear from Fig. 13c that spherulites are not impinging with one
another. Loss of spherulitic boundaries suggests that material
between spherulites is etched away. Therefore it was concluded
that alkaline hydrolysis progressed through the edge of the
lamellas without changing lamellar thickness; the lack of change in
Tm as a function of degradation time supports this conclusion.
FTIR spectra of PLLA lms before and after degradation were
taken and they were very similar to the semicrystalline spectrum
shown in Fig. 4. FTIR spectra suggest that PLLA lms before and
after degradation consists of a a crystalline phase and amorphous
phase. In order to see how crystallinity changes as a function of
hydrolysis for all PLLA lms degraded in NaOH solution, the plot
of the band ratio of crystalline and amorphous area (A921/A956)
versus hydrolysis time was constructed. The band ratio of crystalline and amorphous bands was obtained; the results of A921/
A956 for all PLLA lms are tabulated in Table 3. Table 3 represents
A921/A956 of PLLA lms before and after hydrolysis with respect to
hydrolysis time at RT, 80  C and 110  C. It is apparent that A921/
A956 increased as hydrolysis time increased. However, there was
a small increase in A921/A956 as hydrolysis time increased for all
PLLA lms at RT, 80  C and 110  C. This suggests that FTIR
crystallinity increased with increasing hydrolysis time of all PLLA
lms annealed at RT, 80  C and 110  C. Crystallinity change with
degradation was estimated using the Fig. 6 and it was about 9%

N. Vasanthan, O. Ly / Polymer Degradation and Stability 94 (2009) 13641372

1371

Table 3
Band ratio of crystalline and amorphous area (A921/A956) for PLLA lms before and
after hydrolysis with respect to hydrolysis time at RT, 80  C and 110  C.
Hydrolysis
time (day)

A921/A956
at RTa

A921/A956
at 80  C

A921/A956
at 110  C

0
2
6
10

0.0305
0.0417
0.0431
0.0581

0.0434
0.0480
0.0533
0.0633

0.5622
0.6154
0.6384
0.6337

Room temperature about 25  C.

4. Conclusions
Thermally induced crystallization of PLLA lms was studied
using FTIR spectroscopy and DSC. Tg and Tm showed an increase
with increasing Ta. The crystallinity also increased as a function of
Ta, as expected. An increase in absorbance of bands at 697, 739, 921
and 1293 cm1 and a decrease in absorbance of bands at 710, 757,
956 and 1302 cm1 were observed with increasing Ta. Infrared
bands showing an increase in absorbance were attributed to the
crystalline phase while the bands showing a decrease in absorbance were assigned to the amorphous phase. It was shown
that absorbance of the bands at 2996, 2961 and 2945 cm1 do not
change signicantly with increasing Ta and therefore these bands
were used as the thickness band for normalization. In the present
study the bands at 921and 956 cm1 were used to characterize the
crystalline and amorphous phases, respectively. The crystallinity of
PLLA lms at different Ta was obtained by an independent IR
method and it was in good agreement with the crystallinity
obtained by DSC. The degradation of PLLA lms in alkaline solution
obtained by weight loss study showed an increase with increasing
crystallinity, which was in agreement with the percentage of lactic
acid in hydrolyzed samples obtained by HPLC study. The DSC
showed an increase in Tg and no signicant change in Tm and heat of
fusion while IR showed an increase in IR crystallinity with hydrolysis. The increase in IR crystallinity and Tg with revealed that
degradation progressed from the edges of the crystalline lamellas
without decreasing lamellar thickness, but increases the intermediate phase and the short-range order.
Acknowledgement
The authors thank Professor Glen Lawrence of Long Island
University for his technical help of HPLC analysis. This work has
been partially supported by Department of Commerce and National
Textile Center. The authors also thank reviewers for their helpful
comments to improve the quality of this paper.
References

Fig. 13. Optical micrograph of PLLA lm a) annealed at 110  C; b) annealed at 110  C,


degraded for 6 days; c) annealed at 110  C, degraded for 12 days.

increase for PLLA as prepared, 6% for PLLA annealed at 80  C and


3% for PLLA annealed at 110  C after 10 days of degradation. Since
there was no change in DSC crystallinity, it is reasonable to
assume there is an increase in intermediate phase with hydrolytic degradation. The increased intermediate phase is probably
due to hydrolysis of restricted amorphous region that provide
more intermediate phase.

[1] Nostrum CV, Veldhuis FJ, Bos GW, Hennink WE. Hydrolytic degradation of
oligo(lactic acid): a kinetic and mechanistic study. Polymer 2004;45(20):
677987.
[2] Jain RJ. The Manufacturing techniques of various drug loaded biodegradable
poly(lactide-co-glycolide)(PLGA) devices. Biomaterials 2000;21(23):247590.
[3] Iwata T, Doi Y. Morphology and enzymatic degradation of poly (L lactic acid)
single crystals. Macromolecules 1998;31(8):2461.
[4] Fujita M, Doi Y. Annealing and melting behavior of poly (L-lactic acid) single
crystals as revealed by in situ atomic force microscopy. Biomacromolecules
2003;4(5):13017.
[5] Lee JH, Park TG, Park HS, Lee DS, Lee YK, Yoon SC, et al. Thermal mechanical
characteristics of poly(L-lactic acid). Biomaterials 2003;24(16):27738.
[6] Nam YS, Park TG. Biodegradable polymeric microcellular foams by modied
thermally induced phase separation method. Biomaterials 1999;20(19):
178390.
[7] Okuzaki H, Kubota I,, Kunugi T. Mechanical properties and structure of the
zone-drawn poly (L-lactic acid) bers. J Polym Sci Phys 1999;37(10):9916.

1372

N. Vasanthan, O. Ly / Polymer Degradation and Stability 94 (2009) 13641372

[8] Kulkarni RK, Moore EG, Hegyeli AF, Leonard F. Biodegradable poly(lactic acid)
polymers. J Biomed Mater Res 1971;5(3):16981.
[9] Tsuji H, Nakahara K. Poly(L-lactide). IX. Hydrolysis in acid media. J Appl Polym
Sci 2002;86(1):18694.
[10] Ajioka M, Suizu H, Higuchi C, Kashima T. Aliphatic polyesters and their
copolymers synthesized through direct condensation polymerization. Polym
Degrad Stab 1998;59(13):13743.
[11] Urayama H, Kanamori T, Fukushima K, Kimura Y. Controlled crystal nucleation
in the melt-crystallization of poly(L-lactide) and poly(L-lactide)/poly(D-lactide)
stereocomplex. Polymer 2003;44(19):563541.
[12] Lunt J. Large-scale production, properties and commercial applications of
polylactic acid polymers. Polym Degrad Stab 1998;59(13):14552. 1998.
[13] Jamshidi K, Hyon SH, Ikada Y. Thermal characterization of polylactides. Polymer 1988;29(12):222934.
[14] Migliaresi C, Cohn D, De Lollis A, Fambri L. Dynamic mechanical and calorimetric analysis of compression-molded poly(L-lactic acid) (PLLA) of different
molecular weights: effect of thermal treatments. J Appl Polym Sci 1991;43(1):
8395.
[15] Migliaresi C, De Lollis A, Famibri L, Cohn D. The effect of thermal history on the
crystallinity of different molecular weight PLLA biodegradable polymers. Clin
Mater 1991;8(12):1118.
[16] Gilding DK, Reed AM. Biodegradable polymers for use in surgery: polyglycolic/
poly(lactic acid) homo- and copolymers. Polymer 1979;20(12):145964.
[17] Kricheldorf HR, Kreiser-Saunders I, Boettcher C. Polylactones. 31. Sn(II)octoate-initiated polymerization of L-lactide: a mechanistic study. Polymer 1995;
36(6):12539.
[18] Vasanthakumari R, Pennings AJ. Crystallization kinetics of poly(L-lactic acid).
Polymer 1983;24(2):1758.
[19] Kalb B, Pennings AJ. Gneral crystallization behavior of poly (L-lactic acid).
Polymer 1980;21(6):60712.
[20] Baratian S, Hall ES, Lin JS, Xu R, Runt J. Crystallization and solid-state structure
of random polylactide copolymers: poly(L-lactide-co-D-lactide)s. Macromolecules 2001;34(14):485764.
[21] Kolstad JJ. Crystallization kinetics of poly(L-lactide-co-meso-lactide). J Appl
Polym Sci 1996;62(7):107991.
[22] Hoogsteen W, Postema AR, Pennings AJ, Ten Brinke Gerrit, Zugenmaier P. Crystal
structure, conformation and morphology of solution-spun poly(L-lactide) bers.
Macromolecules 1990;23(2):63442.

[23] DeSantis P, Kovacs AJ. Molecular conformation of poly(S-lactic acid).


Biopolymers 1968;6(3):299306.
[24]. Li S, Vert M. Biodegradable aliphatic polyesters, in degradable polymers,
principle and applications. London: Chapman & Hall; 1995.
[25]. Albertsson AC, editor. Degradable aliphatic polyesters; advances in polymer
science, vol.157. Berlin, Germany: Springer; 2002.
[26] Weir NA, Buchanan FJ, Orr JF, Farrar DF, Boyol A. Processing, annealing and
sterilisation of poly-L-lactide. Biomaterials 2004;25(18):393949.
[27] Ikada Y, Tsuji H. Biodegradable polyesters for medical and ecological applications. Macromol Rapid Commun 2000;21(3):11732.
[28] Scott G. Antioxidant control of polymer biodegradation. Macromol. Symp
1999;144:11325.
[29] Schnabel W. Polymer degradation principles and practical applications. USA:
Oxford University Press; 1988. pp. 178181.
[30] Tsuji H, Tezuka Y, Yamada K. Alkaline and enzymatic degradation of L-lactide
copolymers. II. Crystallized lms of poly(L-lactide-co-D-lactide) and poly
(L-lactide) with similar crystallinities. J Polym Sci Part B Polym Phys 2005;
43(9):106475.
[31] Cam D, Hyon S, Ikada Y. Degradation of high molecular weight poly(L-lactide)
in alkaline medium. Biomaterials 1995;16(11):83343.
[32] Fitz BD, Jamiolkowski DD, Andjelic S. Tg depression in poly(L(-)-lactide) crystallized under partially constrained conditions. Macromolecules 2002;35(15):
586972.
[33] Tsuji H, Ikada Y. Properties and morphology of poly(L-lactide):1 annealing
conditions effects on properties and morphology of poly(L-lactide). Polymer
1995;36(14):270916.
[34] Yasuniwa M, Iura K, Dan Y. Melting behavior of poly(L-lactic acid): effects of
crystallization temperature and time. Polymer 2007;48(18):5398407.
[35] Innance S, Maffezzoli A, Leo G, Nicolais L. Inuence of crystal and amorphous
phase morphology of hydrolytic degradation of PLLA subjected to different
processing conditions. Polymer 2001;42(8):3799807.
[36]. Zhang J, Duan Y, Sato H, Tsuji H, Noda I, Yan S, et al. Crystal modications and
thermal behavior of poly(L-lactic acid) revealed by infrared spectroscopy.
Macromolecules 2005;38(19):801221.
[37] Zhang JM, Tsuji H, Noda I, Ozaki Y. Structural changes and crystallization
dynamics of poly(L-lactide) during the cold-crystallization process investigated by infrared and two-dimensional infrared correlation spectroscopy.
Macromolecules 2004;37(17):64339.

You might also like