Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Compatibility Characterization of Poly(lactic acid)/

Poly(propylene carbonate) Blends


XIAOFEI MA, JIUGAO YU, NING WANG
School of Science, Tianjin University, Tianjin 300072, China

Received 18 July 2005; revised 26 September 2005; accepted 26 September 2005


DOI: 10.1002/polb.20669
Published online in Wiley InterScience (www.interscience.wiley.com).

ABSTRACT: The compatibility of poly(lactic acid) (PLA)/poly(propylene carbonate) (PPC)


blends was investigated with Fourier transform infrared (FTIR) spectroscopy, differ-
ential scanning calorimetry (DSC), thermogravimetric analysis (TGA), and tensile
testing. The PLA/PPC blends were prepared over the whole composition range. FTIR
spectroscopy revealed that there were several specific interactions between the chains
of PLA and PPC: the interaction between C H and O¼ ¼C  and C¼ ¼OO¼¼C or
C¼¼OO C dipole–dipole interactions. Moreover, PLA and PPC were compatible.
DSC indicated that PLA and PPC were partially miscible but compatible to some
extent because of the similar chemical natures of the blend components. TGA showed
that the compatibility of PLA and PPC enhanced the thermal stability of PPC in the
blends. As calculated by the Horowitz–Metzger equation, the activation energy for
decomposition (Et) of PPC in PLA/PPC (70/30) was 200.6 kJ/mol, whereas Et of pure
PPC was only 56.0 kJ/mol. A study of the mechanical properties versus the composi-
tion and the strain versus the stress illustrated that there was good compatibility
between PLA and PPC, and the phase inversion of the PLA/PPC system occurred
between 70 and 60 wt % PLA in the PLA/PPC blends. The Pukanszky model gave
credit to very strong interfacial adhesion between PLA and PPC. V C 2005 Wiley Periodi-

cals, Inc. J Polym Sci Part B: Polym Phys 44: 94–101, 2006
Keywords: blends; compatibility; extrusion; poly(lactic acid); poly(propylene carbonate)

INTRODUCTION temperature (Tg) and melting temperature of about


55 and 175 8C, respectively. They require processing
Environmental concerns and a shortage of petro- temperatures in excess of 185–190 8C. At these
leum resources have driven efforts aimed at the temperatures, unzipping and chain-scission re-
bulk production of biodegradable materials.1 actions leading to a loss of molecular weight, as
Poly(lactic acid) (PLA) is a linear aliphatic poly- well as thermal degradations, are known to
ester produced from renewable resources and is occur. Consequently, PLA homopolymers have a
readily biodegradable. PLA is synthesized by the very narrow processing window. The most widely
ring-opening polymerization of lactides and lac- used method for improving PLA processability is
tic acid monomers, which are obtained from the based on melting point depression by the random
fermentation of sugar feed stocks.2 PLA is a incorporation of small amounts of lactide enan-
thermoplastic, high-strength, high-modulus poly- tiomers of the opposite configuration into the poly-
mer.3 PLA homopolymers have a glass-transition mer (i.e., adding a small amount of D-lactide to
L-lactide to obtain poly(D-lactide and L-lactide)
Correspondence to: J. G. Yu (E-mail: makong2001@hotmail. (PDLLA). The melting point depression is accom-
com)
panied by a significant decrease in the crystallin-
Journal of Polymer Science: Part B: Polymer Physics, Vol. 44, 94–101 (2006)
V
C 2005 Wiley Periodicals, Inc. ity and crystallization rates.3 However, the low
94
COMPATIBILITY CHARACTERIZATION 95

deformation at break of PLA still limits its ap- EXPERIMENTAL


plication.2
One potential method of overcoming this diffi- Materials
culty is blending PLA with other thermoplastics.
PLA was obtained from Natureworks LLC (United
Numerous studies on the miscibility of polymer
States). The general molecular weight average was
blend systems have also been devoted to PLA/biode- about 160,000–220,000. The concentration of the D-
gradable polyester blends. Polycaprolactone (PCL) ()-isomer was 12.0 6 1.0%. PPC was obtained
is reactively blended with PLA.4 Mechanical prop- from the Inner Mongolia Meng Xi High-Tech Group
erty measurements indicate that the elongation of (China). The general molecular weight average was
the PLA/PCL blends is improved significantly, and about 70,000–110,000 (weight-average molecular
a synergism is observed for certain compositions weight/number-average molecular weight ¼ 4.3).
(PLA/PCL ¼ 80/20 or 20/80). Poly(3-hydroxybuty-
rate-co-3-hydroxyhexanoate) [P(3HB-co-3HHx); con- Blend Preparation
centration of 3-hydroxyhexanoate ¼ 13.4 mol %]
has the potential to increase the toughness of PLA The blends were mixed in an SJ-25(s) single-
dramatically, thereby enlarging the design space of screw plastic extruder (Axon Australia Pty, Ltd.,
sustainable materials achievable by the blending of Australia) with a screw diameter of 30 mm and
P(3HB-co-3HHx) with PLA. Infrared (IR) spectra a length-to-diameter ratio of 25:1. The tempera-
provide insights into the submolecular-level interac- ture profile along the extruder barrel was 110/
tions and miscibility of this polymer blend system.5 115/120 8C, and the temperature at the die was
Naoyuki and Yoshiharu6 revealed that the structure 90 8C. The screw speed was 20 rpm. The die
of poly(3-hydroxybutyric acid) (P3HB)/PLA blends was a round sheet with 3-mm-diameter holes.
is strongly dependent on the molecular weight of
FTIR Spectroscopy
the poly(L-lactide) (PLLA) component. Blends of
P3HB with PLA of high weight-average molecular FTIR spectra were obtained at a 2-cm1 resolu-
weight (Mw) values (>20,000) show two phases in tion with a Bio-Rad FTS3000 IR spectrum scan-
the melt at 200 8C, whereas blends of P3HB with ner. Typically, 64 scans were signal-averaged to
PLA of low Mw values (<18,000) are miscible in the reduce spectral noise. The extruded thermoplas-
melt over the whole composition range. PLA and tic starch (TPS) strips were pressured into trans-
poly(hydroxy ester ether) are also miscible.7 parent slices with a thickness of around 0.2 mm
Poly(propylene carbonate) (PPC) is also a bio- in a flat sulfuration machine tested by the trans-
degradable aliphatic polycarbonate. In the pres- mission method.10
ence of a heterogeneous catalyst system, propy-
lene oxide and carbon dioxide produce a polymer DSC
involving the regular alternating copolymer A Netzsch DSC 204 differential scanning calorime-
PPC.8,9 PPC is an amorphous polymer and has a ter was employed to measure Tg of the blends. The
chemical structure similar to that of PLA. The specimens were heated from 0 to 100 8C at a rate of
chemical structures of PLA and PPC are shown 10 8C/min. The midpoints of the transitions in the
in Figure 1. In this study, the compatibility of traces recorded in the heating scan were taken as
melt-mixed blends of PLA/PPC in the complete the values of Tg. The specimen weight was 8–10 mg.
composition range was characterized mainly
with Fourier transform infrared (FTIR) spectro- TGA
scopy, differential scanning calorimetry (DSC),
TGA of the blends were measured with a ZTY-
scanning electron microscopy (SEM), thermogra-
ZP-type thermal analyzer. The sample weight
vimetric analysis (TGA), and tensile testing.
varied from 10 to 15 mg. The samples were
heated from room temperature to 500 8C at a
heating rate of 15 8C/min.

Mechanical Testing
Samples [8 cm  3-mm diameter (/)] were cut from
the extruded strips. A Testometric AX M350-10KN
materials testing machine, operated at a crosshead
Figure 1. Chemical structures of PLA and PPC. speed of 50 mm/min, was used for tensile testing
96 MA, YU, AND WANG

Figure 2. FTIR spectra of PLA/PPC blends.

(ISO 1184-1983 standard). An average of five to preciable changes in the FTIR spectra of the
eight bars was recorded for every sample. blends with respect to the coaddition of each
component.11 However, if two polymers are com-
SEM patible, a distinct chemical interaction (a hydro-
gen-bonding or dipolar interaction) exists be-
The fracture surfaces of the samples were stud- tween the chains of one polymer and those of
ied with a Philips XL-3 scanning electron micro- the other, causing the IR spectra for the blend
scope operated at an acceleration voltage of to change (e.g., band shifts and broadening).12
20 kV. The samples were cooled in liquid nitro- As a result, FTIR can identify segment interac-
gen and then broken. The fracture surfaces were tions and provide information about the phase
vacuum-coated with gold for SEM. behavior of polymer blends.
Figure 2 shows the FTIR spectra of PLA/PPC
blends at room temperature in several specific
RESULTS AND DISCUSSION
stretching regions. The peak band wave num-
bers and assignments for PLA and PPC IR spec-
FTIR
tra are listed in Table 1.
The interaction of polymer blends can be identi- In the 3700–3100-cm1 region, the absorption
fied with FTIR spectra. If two polymers form of PPC and PLA is negligible because of the
completely immiscible blends, there are no ap- rareness of terminal OH groups. With the ad-
COMPATIBILITY CHARACTERIZATION 97

Table 1. Peak Band Wave Numbers (cm1) and Assignments for PLA and PPC IR Spectra


C¼¼O C
O C
O
Assignment 
CH
 Carbonyl 
O
C
O
 in 
CHO in 
OC¼
¼O

PLA (cm1) 2997, 2949 1749 — 1182 1127, 1082, 1044


PPC (cm1) 2986 1738 1223 1165 1124, 1065
Symbol A region B region i line C region D region

dition of PPC, the peak band of PLA in the A PLA and PPC indicate that there are several
region, ascribed to the CH stretching vibra- specific interactions between the chains of PLA
tion, changes greatly when the PPC concentra- and PPC, which are compatible.
tion varies from 0 to 100%; this suggests some
changes in the intermolecular interactions due
DSC
to blending. The electron attraction of ester in
PLA and carbonic ester in PPC lowers the den- Figure 3 shows the DSC traces of PLA/PPC
sity of the electron cloud of methine hydrogen, blends. A DSC analysis of the blends clearly in-
making an interaction possible between CH dicates two Tg values between PLA and PPC.
and OC or between C H and O¼ ¼C. The spectra are commonly observed for partially
Both PLA and PPC have a strong carbonyl miscible polymer pairs.
stretching absorption in the B region. With the There is a difference in the convergence of
addition of PPC, the C¼ ¼O peak of PLA at 1749 the main Tg between PLA and PPC. Tg of PLA
cm1 shifts to a lower wave number (by 11 cm1) changes slightly from 57 to 54 8C for the PLA/
when the PPC concentration increases from 0 to PPC (30/70) blend, being practically independent
100%. The interaction formation reduces the of the nominal composition, as shown in Table 2,
stretching vibration frequency of the carbonyl whereas Tg of PPC shifts greatly from 22 to
bond and gives rise to a shift to a lower frequency, 43 8C for the PLA/PPC 70/30 blend. For the 70/
as extensively studied for poly(vinylphenol)/PCL 30 PLA/PPC composition, the difference between
blends.13 This means that carbonyl groups also the Tg values of PLA and PPC is the lowest.
take part in the interaction between PLA and The DSC results support the view of a partially
PPC and result in the blueshift of the C¼ ¼O vibra- miscible blend in the temperature regime in
tion of PLA. One reason is the interaction which this method is applied. This is supported
between C H and O¼ ¼C , as mentioned previ- by the considerable component Tg convergence
ously. Another reason is C¼ ¼OO¼¼C or in the blend, particularly that of PPC. This indi-
C¼¼OOC dipole–dipole interactions, similar to cates that PLA/PPC blends are partially misci-
the dipole–dipole C¼ ¼OClC interaction sug- ble but compatible to some extent because of the
gested by Allard and Prudhomme14 for PCL/
chlorinated polypropylene blends.
According to Fei et al.,15 the  OC O
stretching vibration appears at 1223 cm1, which
deviates from the i line in Figure 2 with decreasing
PPC contents. The C and D regions in Figure 2,
separately related to the CO stretching
vibration in the  CHO  group and  O
C¼ ¼O group, show that the CO stretching
vibration in the CHO group, appearing at
17 cm1, blueshifts from pure PLA to PPC,
whereas the main peak of the CO stretching
vibration in the OC¼ ¼O group also has a 17-
cm1 blueshift with decreasing PPC contents. The
C¼¼OOC dipole–dipole interactions and the
interaction between CH and O C exist.
In the FTIR spectra of PLA/PPC blends, the
shifting and broadening of some listed bands for Figure 3. DSC traces of PLA/PPC blends.
98 MA, YU, AND WANG

Table 2. Tg’s of PLA/PPC Blends TGA curves by the integral method proposed by
Horowitz and Metzger with eq 1:19
Tg (8C)
ln½lnð1  aÞ1  ¼ Et h=RT 2 max ð1Þ
PLA/PPC PLA PPC

100/0 57 — where a is the decomposed fraction, T is temper-


70/30 56 43 ature, h is T  Tmax, and R is the gas constant.
50/50 55 33 From the plots of ln[ln(1   a)1] versus h,
30/70 54 29 which are shown in Figure 6, Et can be deter-
0/100 — 22 mined from the slope of the straight line of the
plots. As demonstrated in Table 3, the Et values
of both PLA and PPC in blends are raised as
similar chemical natures of the blend compo- the PLA concentration increases from 0 to 70%.
nents; this is similar to the results for polycar- Et of PPC in PLA/PPC (70/30) is 200.6 kJ/mol,
bonate/copolyester blends.16 whereas Et of pure PPC is only 56.0 kJ/mol.
In view of Et, PLA (228.8 kJ/mol) in PLA/PPC
(70/30) is even more thermally stable than pure
TGA
PLA (213.9 kJ/mol). This is attributed to the
TGA has been performed for the blends, and the specific interactions between the chains of PLA
weight loss due to the volatilization of the deg- and PPC, which are similar to poly(vinyl chlor-
radation products has been monitored as a func- ide)/ethylene–vinyl acetate copolymer blends.20
tion of temperature, as shown in Figure 4. T5%, Although the degradation products of PPC
which is the temperature corresponding to 5% facilitate the thermal degradation of PLA at
weight loss determined from Figure 4, is sum- lower temperatures, the specific interactions
marized in Table 3. T5% of the PLA/PPC blends between the chains of PLA and PPC improve
increases from 202 to 351 8C as the concentra- the activation energy of PLA decomposition.
tion of PLA increases from 0 to 100 wt %. PPC
shows low thermal stability. Dong et al.17 re-
ported that PPC is easily decomposed to cyclic car- Mechanical Properties
bonate at a temperature of only about 180 8C.
Figure 7(A) illustrates the yield-stress/composi-
Meng et al.18 found that PPC pyrolysis obeys
tion relationship for the PLA/PPC blends. Some
a two-step pyrolysis mechanism: main-chain
studies have suggested that the yield behavior
random scission and unzipping. The addition of
of polymer blends is affected by the interfacial
PLA improves the thermal stability of PPC.
adhesion.21–23 Pukanszky and coworkers21–23
Derivative thermogravimetry (DTG) curves for
proposed upper and lower values for the yield
PLA/PPC blends in Figure 5 show the Tmax val-
stress in cases of extreme values of interfacial
ues of PPC and PLA in PLA/PPC blends, which
also are listed in Table 3. Here Tmax is the tem-
perature at the maximum rate of weight loss,
that is, the decomposition temperature. Tmax of
PPC and PLA in blends is located between those
of pure PPC and PLA. Tmax of PPC in the PLA/
PPC (70/30) blend reaches 551.55 K, in compari-
son with 528.85 K for pure PPC. This Tmax
improvement of PPC is ascribed to the compati-
bility of PLA and PPC in blends. However, Tmax
of PLA in PLA/PPC blends decreases with
increasing contents of PPC. Tmax of PLA de-
creases from 675.95 K for pure PLA to 570.35 K
for the PLA/PPC (30/70) blend. The degradation
products of PPC possibly facilitate the thermal
degradation of PLA.
The activation energy of decomposition, Et, of Figure 4. Thermogravimetric curves for PLA/PPC
the polymer blends can be calculated from the blends at a heating rate of 15 8C/min in air.
COMPATIBILITY CHARACTERIZATION 99

Table 3. Thermogravimetric Parameters of PPC and PLA Blends

PLA/PPC T5% PPC PLA Slope PPC Et PLA Et


(w/w) (8C) Tmax (K) Tmax (K) of Eq 1 (kJ/mol) (kJ/mol)

0/100 202 528.85 — 0.0241 56.0 —


30/70 229 534.85 570.35 0.0589 140.1 159.3
50/50 243 550.55 580.25 0.0598 150.7 167.4
70/30 256 551.55 589.05 0.0793 200.6 228.8
100/0 351 — 675.95 0.0563 — 213.9

adhesion. When the interfacial adhesion is facial adhesion. The dotted line has been plot-
strong enough for stress transfer to occur ted with eq 2 (strong interfacial adhesion),
between two phases, the yield stress obeys the whereas the dashed–dotted line has been plotted
law of mixtures (the upper value): with eq 3 (no interfacial adhesion). The PLA/
PPC blends have a significant positive devi-
rb ¼ r1 /1 þ r2 /2 ð2Þ ation above 30 wt % PPC, whereas the PLA/
PPC blends basically follow the additive rules
where b is the blend, r is the yield stress and predicted by eq 2 below 30 wt % PPC. The
subscripts 1 and 2 refer to component 1 (PLA) Pukanszky model gives credit to very strong
and component 2 (PPC), respectively. interfacial adhesion between PLA and PPC.
In the case of a lack of interfacial adhesion, Young’s modulus of polymer blends mainly
the dispersion of the minor component only re- depends on the modulus of each constitutive
sults in a reduction of the load-bearing cross component and blend composition. It is also
section of the matrix. The yield stress in then somewhat affected by the interracial interac-
calculated with eq 4 (the lower value): tions and changes in the phase morphology.21
Figure 7(B) shows that this general behavior is
1  /d confirmed by the PLA/PPC blends. When PPC is
r0b ¼ rm ð3Þ
1 þ 2:5/d the continuous phase at high PPC concentra-
tions (>40 wt %), PLA/PPC blends show a posi-
where superscript 0 denotes zero interfacial tive deviation in the modulus with respect to
adhesion, m is the matrix or continuous phase, the weight average values (dotted line). When
and d is the dispersed phase. Figure 7(A) com- the PPC concentration is less than 30 wt %,
pares the experimental data (full line and full PLA becomes the continuous phase in PLA/PPC
circles) with the predictions for extreme inter- blends, which show a negative deviation in the

Figure 6. Plots of ln[ln(1  a)1] versus h for the


Figure 5. DTG curves for PLA/PPC blends at a determination of Et. The straight line is the linear fit
heating rate of 15 8C/min in air. of the data points.
100 MA, YU, AND WANG

Figure 7. Mechanical properties versus the composition for PLA/PPC blends: (A)
yield stress, (B) Young’s modulus, (C) break stress, and (D) break energy.

modulus. The interfacial adhesion is compara- Both the number and height of the rubber pla-
tively strong when PPC is the continuous phase. teau vary with the PLA concentration. At a low
The phase inversion of the two-phase system PLA concentration (15 wt %), there is only one
occurs between 30 and 50 wt % PPC. plateau. When the PLA concentration reaches 30
As shown in Figure 7(C), the break stress of wt %, double plateaus appear. The first one is
PLA/PPC blends well follows the additive law related to the tensile strain of the PPC matrix
(dotted line) at high PPC contents. The beneficial with a low height, whereas the second one is re-
effect of the PLA/PPC interfacial adhesion could lated to the tensile strain of PLA with a high
also account for the negative deviation observed height. Because of the good compatibility, the PPC
in the dependence of the break stress on the PPC matrix induces PLA (dispersion phase) yielding.
content [Fig. 7(C)]. Indeed, the break stress of
PLA is much higher than that of PPC, so the con-
tinuous PPC phase is reinforced by PLA as effi-
ciently as the interfacial adhesion is high.
The break energy shows a remarkable posi-
tive deviation that corresponds to a synergism
over more than 85% of the composition range
[Fig. 7(D)]. The break energy is known to result
from a complex interplay of several experimen-
tal parameters, such as the phase morphology,
relative modulus of the phases, chain structure,
and interfacial adhesion. The continuous PPC
phase results in a lower yield stress and thus
favors the matrix yielding, which requires more
energy to break the materials.24
As shown in Figure 8, the stress–strain curves
of PLA/PPC blends have similar rubber plateaus. Figure 8. Strain–stress curves of PLA/PPC blends.
COMPATIBILITY CHARACTERIZATION 101

Figure 9. SEM micrographs of cryofractured PLA/PPC blends: (a) 70/30 PLA/PPC


and (b) 60/40 PLA/PPC.

With increasing PLA concentration, the height 4. Wang, L.; Ma, W.; Gross, R. A.; McCarthy, S. P.
of the plateaus increases. When the PLA concen- Polym Degrad Stab 1998, 59, 161–168.
tration reaches 60 wt %, the second plateau is 5. Marcott, C.; Dowrey, A. E.; Van Poppel, J.; Noda,
obviously shortened. When the PLA concentra- I. Vib Spectrosc 2004, 36, 221–225.
6. Naoyuki, K.; Yoshiharu, D. Polymer 1997, 38,
tion reaches 70 wt %, only one plateau exists,
1589–1593.
and the strain descends rapidly to 35%. Here
7. Cao, X.; Mohamed, A.; Gordon, S. H.; Willett, J. L.;
phase inversion of the two-phase system occurs, Sessa, D. J. Thermochim Acta 2003, 406, 115–127.
as shown in Figure 9. The PLA matrix is 8. Chisholm, M. H.; Navarro-Llobet, D.; Zhou, Z. P.
smoother than the PPC matrix. The PLA matrix Macromolecules 2002, 35, 6494–6504.
has a much lower tensile strain but stronger 9. Chisholm, M. H.; Zhou, Z. P. J Am Chem Soc
stress than the PPC matrix. 2004, 126, 11030–11039.
From the results of the mechanical properties 10. Pawlak, A.; Mucha, M. Thermochim Acta 2003,
versus the composition and the strain versus the 396, 153–166.
stress, it can be concluded that there is good com- 11. Zhang, G. B.; Zhang, J. M.; Zhou, X. S.; Shen,
patibility between PLA and PPC, and phase in- D. Y. J Appl Polym Sci 2003, 88, 973–979.
12. Peng, S. W.; Wang, X. Y.; Dong, L. S. Polym Com-
version of the PLA/PPC system occurs between
pos 2005, 26, 37–41.
70 and 60 wt % PLA in PLA/PPC blends. 13. Kuo, S. W.; Huang, C. F.; Chang, F. C. J Polym
Sci Part B: Polym Phys 2001, 39, 1348–1359.
14. Allard, D.; Prudhomme, R. E. J Appl Polym Sci
CONCLUSIONS 1982, 27, 559–568.
15. Fei, B.; Chen, C.; Peng, S. W.; Zhao, X. J.; Wang,
FTIR, DSC, TGA, and mechanical properties X. H.; Dong, L. S. Polym Int 2004, 53, 2092–2098.
illustrate that PLA and PPC are partially misci- 16. Samios, C. K.; Kalfoglou, N. K. Polymer 2000, 41,
ble but are compatible to some extent because of 5759–5767.
the similar chemical natures of the blend com- 17. Peng, S. W.; An, Y. X.; Chen, C.; Fei, B.; Zhuang,
ponents. The compatibility of PLA and PPC Y. G.; Dong, L. S. Polym Degrad Stab 2003, 80,
enhances Tg and Tmax of PPC in blends. There 141–147.
are several specific interactions between the 18. Lia, X. H.; Meng, Y. Z.; Zhu, Q.; Tjong, S. C.
chains of PLA and PPC. Between 70/30 and 60/ Polym Degrad Stab 2003, 81, 157–165.
19. Horowitz, H. H.; Metzger, G. Anal Chem 1963,
40 PLA/PPC compositions, the phase inversion
35, 1464–1468.
of the PLA/PPC system occurs. 20. Elisabeth, E. C. M.; Clelio, T. Compos Sci Technol
1997, 57, 1159–1165.
21. Pukanszky, B.; Tudos, F. Makromol Chem 1990,
REFERENCES AND NOTES 38, 221–231.
22. Pukanszky, B. Makromol Chem 1993, 70, 213–223.
1. Zhang, J. F.; Sun, X. Z. Biomacromolecules 2004, 23. Kolarik, J.; Lednicky, F.; Pukanszky, B.; Pegoraro,
5, 1446–1451. M. Polym Eng Sci 1992, 32, 886–893.
2. Martin, O.; Avérous, L. Polymer 2001, 42, 6209–6219. 24. Liu, Z. H.; Marechal, P.; Jerome, R. Polymer
3. Garlotta, D. J Polym Environ 2001, 9, 63–84. 1998, 39, 1779–1785.

You might also like