Vous êtes sur la page 1sur 13

Plasticized Polylactide/Clay Nanocomposites. I.

The Role
of Filler Content and its Surface Organo-Modication
on the Physico-Chemical Properties
MIROSLAW PLUTA,1 MARIE-AMELIE PAUL,2 MICHAEL ALEXANDRE,2 PHILIPPE DUBOIS2
1

Centre of Molecular and Macromolecular Studies, Polish Academy of Sciences, 90-363 Lodz, Poland

Laboratory of Polymeric and Composite Materials, University of Mons-Hainaut, 7000 Mons, Belgium

Received 29 July 2005; revised 15 October 2005; accepted 15 October 2005


DOI: 10.1002/polb.20694
Published online in Wiley InterScience (www.interscience.wiley.com).

Polylactide (PLA)-layered silicate nanocomposites plasticized with 20 wt % of


poly(ethylene glycol) 1000 were prepared by melt blending. Three kinds of organo-modied
montmorillonitesCloisite1 20A, Cloisite1 25A, and Cloisite1 30Bwere used as llers
at a concentration level varying from 110 wt %. Neat PLA and plasticized PLA with the
same thermomechanical history were considered for comparison. Nanocomposites based
on amorphous PLA were obtained via melt-quenching. The inuence of both plasticization
and nanoparticle lling on the physicochemical properties of the nanocomposites were
investigated. Characterization of the systems was achieved by size exclusion chromatography (SEC), thermogravimetric analysis (TGA), thermally modulated differential scanning
calorimetry (TMDSC), X-ray diffraction (XRD), and dynamic mechanical analysis (DMTA).
SEC revealed a decrease of the molecular weight of the PLA matrix with the ller content.
Thermal behavior on heating showed one cold crystallization process in the reference neat
PLA sample, while two cold crystallization processes in plasticized PLA and plasticized
nanocomposites. The thermal windows of these processes tend to increase with the ller
content. The crystalline form of PLA developed upon heating was affected neither by the
plasticization nor by the type and content of Cloisite used. It was found that the series of
organo-modied montmorillonites with decreasing afnity to PLA is Cloisite1 30B,
Cloisite1 20A, and Cloisite1 25A, respectively. The dynamic mechanical properties were
sensitive to the sample composition. Generally, the storage modulus increased with the
ller content. Glassy PEG, well dispersed within unlled PLA matrix, exhibited also a
reinforcing effect, since the storage modulus of this sample was higher than for unplasticized reference at temperature region below the glass transition of PEG. Moreover, loss
modulus of all plasticized samples revealed an additional maximum ascribed to the glass
transition of PEGrich dispersed phase, indicating partial miscibility of organic components of the systems investigated. The magnitude of this mechanical loss was correlated
with the ller content, and to some extent, also with the nanoller ability to be intercalated
C 2005 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 44: 299
by polymer components. V
ABSTRACT:

311, 2006

Keywords: montmorillonite; organoclay; physical properties; plasticized nanocomposites; polylactide

INTRODUCTION
Correspondence to: M. Pluta (E-mail: mpluta@bilbo.cbmm.
lodz.pl)
Journal of Polymer Science: Part B: Polymer Physics, Vol. 44, 299311 (2006)
C 2005 Wiley Periodicals, Inc.
V

Recently, an enormous interest has arisen from


nanocomposite materials, especially for those
prepared from polymer matrix and nanoscale299

300

PLUTA ET AL.

sized inorganic particles. Different polymer


matrices as well as various kinds of nanoparticles, with respect of geometry and properties,
have been consideredisotropic nanoparticles
such as Al2O3 or ZnO1 but also nanoparticles
characterized by higher aspect ratio such as
nanotubes2 or layered llers.3 More particularly,
polymer-layered silicate (clay) nanocomposites
represent an interesting class of materials due to
the variety of structural forms possible to obtain
(intercalated, exfoliated, mixed), leading to signicant improvement of the physicochemical
properties at low ller content (from 1 to 5 wt %)
in comparison with the neat polymer or microcomposite counterparts.4 Such an improvement
of the properties has motivated the academic and
industrial communities to develop nanocomposites, playing on the composition or on the preparation method. Three main preparation techniques can be distinguished5: blending the molten
polymer matrix with the layered silicate, mixing
the polymer with the layered silicate in a common solvent (followed by recovering of the nanocomposite material by solvent evaporation or precipitation in a nonsolvent), or in situ polymerization of the monomer in the presence of dispersed
nanoparticles.
This work deals with the preparation and
characterization of nanocomposites prepared
from plasticized polylactide (PLA) and different
types of organically surface-treated montmorillonites, that is 1-nm thick-layered aluminosilicates with high aspect ratios (501000). Plasticized PLA-based nanocomposites have a chance
to nd new application in the eld of exible
packaging material. Furthermore, Cloisites applied are available as commercial products. The
melt blending technique, in which shearing processes are involved to disperse the ller within the
polymer matrix, was used as a suitable method
for large-scale production. PLA was chosen as a
matrix because of its interesting mechanical
properties (comparable to atactic polystyrene6)
and good processability on standard processing
equipments. Beside, PLA can be produced from
annually renewable resources; it exhibits other
advantageous features, including biocompatibility and biodegradability in short time periods
(ecological aspect).6 Moreover, the increasing
availability of PLA, due to the development of
large scale production (for example, ref. 7),
makes this polymer convenient for the production of numerous environmental-friendly disposable plastic articles.

It has been shown that PLA-clay systems prepared by solution mixing and then casting8
exhibited improved mechanical properties, even
if the ller, arranged in the form of tactoids consisting of several silicate monolayers, did not
lead to the formation of a real nanocomposite
(neither complete intercalation nor exfoliation).
On the other side, the PLA-clay nanocomposites
prepared by blending of the molten polymer
with organo-modied clays (3 wt %) exhibited
intercalated structure and increased storage modulus compared with the unlled polymer
matrix.911 Intercalated morphologies have also
been found in PLA-clay nanocomposites prepared by extrusion.1216 In these latter works,
different properties, including mechanical, thermal, biodegradability, and permeability have
been investigated.
It is known that PLA is not exible enough
and breaks down at rather low deformation. Filling PLA with layered silicates leads to nanocomposites of even higher stiffness. This feature may
be undesired for some end-use applications of the
nanocomposites. A way for reducing the rigidity
of PLA-based materials consists in its plasticization that can be achieved simply by blending the
PLA matrix with low molecular weight additives
or another miscible polymer, characterized by a
low Tg. The best plasticizing effect in PLA is
obtained with the monomer lactide itself.6 Other
organic molecules have been tested as possible
plasticizers for PLA : oligo(e-caprolactone),12 citrate esters,17 poly(ethylene glycol) (PEG),11,18
glucose monoesters and partial fatty acid
esters,19 glycerol and oligomeric lactic acid,20 triacetine, acetyl tributyl citrate, and acetyl triethyl
citrate.21
In this study, we have considered PEG-plasticized PLA nanocomposites lled with layered silicates organo-modied with different ammonium
cations (namely, the commercially available Cloisite1 20A, Cloisite1 25A, and Cloisite1 30B, see
experimental). Owing to the organo-modication,
the clays are intercalated with alkylammonium
cations bearing long alkyl chains, which increase
the interlayer spacing and can improve the compatibility of layered silicates and the polymer
matrix. As a result, chains of polymer matrix can
intercalate the layered organo-clay and may further favor exfoliation, that is the dispersion of
individual platelets upon blending under shear.
In all the considered nanocomposites, PLA
matrix was plasticized by 20 wt % PEG with Mn
1000 (PEG1000). A simple blend of PLA and

PLASTICIZED PLA/CLAY NANOCOMPOSITES

301

Table 1. Characteristics of the Organo-Modied Montmorillonites

Montmorillonite
Type
Cloisite1 20A
Cloisite1 25A
Cloisite1 30B

Ammonium Cation
(Hydrogenated-C18-C16-C14)2-N(CH3)2
(Hydrogenated-C18-C16-C14)-N
(CH3)2[CH2-CH(C2H5)-C3H9]
(C18-C16-C14)-N(C2H4OH)2CH3

PEG1000 was also prepared as a reference composition. The attention was focused on three main
topics: (1) the effect of the ller concentration (0,
1, 3, 5, and 10 wt %), (2) the nature of its organomodication as well as (3) the effect of plasticization on the basic physicochemical and thermal
characteristics of the prepared nanocomposites.
Molecular parameters were determined by sizeexclusion chromatography (SEC). The type of
nanocomposites was established by X-ray diffraction (XRD). Viscoelastic measurements (DMTA)
were used to study the mechanical response and
relaxation processes in relation to the composition
and temperature. Thermal behavior on heating
was determined using thermally modulated differential scanning calorimetry (TMDSC). Finally, the
most efcient ller for the modication of plasticized PLA was dened via the magnitude of the
intercalation and the extent of the thermomechanical property changes.

EXPERIMENTAL
Materials
Poly(L,L-lactide) containing 100% of L,L-lactide
units (abbreviated PLA) (from Galactic S.A., Mn
81,800, Mw/Mn 1.9) was used as the matrix.
Poly(ethylene glycol) with a molecular weight of
1000 (PEG1000) (from SigmaAldrich, Fluka
div.) was selected as the plasticizer of the PLA.
Three organically treated clays supplied by
Southern Clay Products (Gonzales, TX) were
used in this study: Cloisite1 20A (modied with
dimethyl di(hydrogenated tallowalkyl) ammonium cations), Cloisite1 25A (modied with
dimethyl-2-ethylhexyl(hydrogenated tallowalkyl)
ammonium cations), and Cloisite1 30B (modied
with methyl-bis(2-hydroxyethyl) tallowalkyl ammonium cations). Description of the organo-modied montmorillonites used in this study is given
in Table 1.

Organic
Fraction
(wt %)

Interlayer
Spacing
(nm)

26.0
29.2

2.36
2.04

20.1

1.84

Nanocomposite Preparation
Prior to the preparation, PLA was dried at 60 8C
overnight under reduced pressure and stored
under vacuum in the presence of a humidity
absorbent. The clays were dried at 40 8C for 4 h
under reduced pressure. PEG1000 was used as
received. Melt blending of PLA with clay particles
and PEG1000 was carried out in the presence of
0.3 wt % of Ultranox 626 stabilizer (General Electric Co.). The components were loaded simultaneously into an internal mixer (Brabender OHG)
and blended at a rotation speed of 20 rpm during
4 min. and then at 60 rpm for 3 min. The processing temperature was set at 185 8C; however, it
increased to about 195 8C as a result of shearing.
Nanocomposites containing 1, 3, 5, and 10 wt %
(relative to the inorganic content) of organomodied montmorillonite Cloisite1 20A, Cloisite1
25A, and Cloisite1 30B were compounded, respectively. 20 wt % PEG1000-plasticized PLA
without nanoller was prepared as well. From
these materials, 0.5-mm thick samples were prepared by compression molding at 185 8C and then
quenching between two aluminum sheets at 0 8C.
All nanocomposites considered in this work are
specied in Table 2. In this Table, the abbreviation of pN10B30 for instance refers to plasticized
nanocomposites (pN) based on 10 wt % Cloisite1
30B (10B30) and having a composition PLA:
PEG1000:Cloisite1 30B in the weight ratio of
70:20:10. Melt-quenched neat PLA and PLA plasticized with 20 wt % of PEG1000 are denoted as
PLA and pPLA, respectively. In time-preceding
characterization, all samples were stored in
closed plastic bags at lowered temperature to
4 8C, that is below the melting temperature of
PEG component (Tm(PEG)) and well below the
glass-transition temperature of PLA (Tg(PLA)),
to inhibit structural reorganization ascribed to
the aging effects known to occur at ambient temperature.18,22,23

302

PLUTA ET AL.

Table 2. Codes and Compositions of the Nanocomposites


Clay Nature/Sample Abbreviation
Composition by wt %
PLA/PEG/Clay

Cloisite1 20A

Cloisite1 25A

Cloisite1 30B

79:20:1
77:20:3
75:20:5
70:20:10

pN1A20
pN3A20
pN5A20
pN10A20

pN1A25
pN3A25
pN5A25
pN10A25

pN1B30
pN3B30
pN5B30
pN10B30

Characterization
The numberaverage molecular weight (Mn) and
polydispersity index (Mw/Mn) of the neat PLA and
PLA extracted from the nanocomposites were
determined by SEC. These experiments were
aimed at quantifying the effect of additives on
changes of Mn upon thermal processing. All the
samples were dissolved in chloroform and ltered
off to eliminate the clay when present. Residual
catalyst was removed by liquidliquid extraction
with a 0.1 M HCl aqueous solution, and PLA was
recovered by precipitation of the chloroform solution from cold methanol at 4 8C. SEC measurements were carried out in tetrahydrofuran (THF;
sample concentration: 2 wt %) with a Polymer
Laboratory (PL) liquid chromatograph equipped
with a PL-DG802 degazer, an isocratic HPLC
pump (LC1120, PL; ow rate 1 mL/min), a
Basic-Marathon autosampler from PL, a PL-RI refractive index detector, and four columns: a Plgel
10-lm guard column (50  7.5 mm2) and three
Plgel 10-lm mixed-B columns (300  7.5 mm2).
Molecular weights and molecular weight distributions were calculated by reference to a PS standard calibration curve, with the Khun-Mark-Houwink equation for poly(L-lactide) in THF: Mn(PLA)
0.4055  Mn(PS)1.0486.24 It should be noticed
that this procedure may lead to some loss of low
molecular weight fraction of extracted PLA.
The thermal behavior was measured using
Thermally Modulated DSC 2920 (TMDSC) from
TA Instruments. The TMDSC technique is an
enhancement to conventional DSC whereby the
total heat ow is separated into reversible events
(i.e. the glass transition, melting) and nonreversible components (contains kinetic events such as
cold crystallization, crystal perfection and reorganization, curing, and decomposition reactions).25
Measurements were performed with a heating
only prole at a ramp of 3 8C/min, with a modulation period of 40 s and a modulation temperature

amplitude of 0.318 8C according to the manufacturer principles25 and own experience with the
PLA samples investigated. For the PLA, the glasstransition temperature (Tg(PLA)) and the enthalpy
of transitions (DHc, DHm) connected with structural transformations of exothermic and melting
natures were determined. Conventional DSC characterization of the PEG1000, taken as the plasticizer of PLA, was performed using heating/cooling
run at a rate of 5 8C/min. Baseline calibration was
performed according to the experimental requirements.
The structure of nanocomposites as well as
development of the crystalline structure of the
PLA upon heating was investigated using hh
goniometer of Siemens Diffractometer D5000
with Ni-ltered Cu Ka radiation (k 0.154 nm).
The measurements were performed at the same
experimental conditions for all systems using
comparable sample thickness (i.e. comparable
scattering volume) for the purpose of comparison
and qualitative analysis of the recorded diffraction intensities.
Dynamic mechanical properties of all samples
were measured with an MkIII DMTA apparatus
(Rheometric Scientic, Inc.) in a dual-cantiliver
bending mode. The dynamic storage and loss
moduli (E0 and E@) were determined at a constant frequency of 1 Hz as a function of temperature from 90 to 150 8C at a heating rate of
3 8C/min.

RESULTS AND DISCUSSION


Size Exclusion Chromatography
PLA is reported to easily degrade upon melt processing and it can loose even up to 80% of its initial
molecular weight, depending on its grade and the
processing conditions.26 It is known that the
hydrolytic degradation mechanism is dominant

PLASTICIZED PLA/CLAY NANOCOMPOSITES

303

Table 3. Molecular Characteristics of PLA Chains Extracted From Nanocomposites


as Determined by SEC
Sample

Mn

Mw/Mn

Sample

Mn

Mw/Mn

Sample

Mn

Mw/Mn

pN1A20
pN3A20
pN5A20
pN10A20

52,700
38,400
37,400
28,400

1.6
1.7
1.8
1.7

pN1A25
pN3A25
pN5A25
pN10A25

56,400
62,900
n.d.
42,200

1.7
1.7
n.d.
1.7

pN1B30
pN3B30
pN5B30
pN10B30

60,300
39,400
n.d.
28,650

1.8
1.8
n.d.
1.9

n.d, not determined.

up to 215 8C, and above, nonhydrolytic degradation becomes equally important.27 Therefore, it
was interesting to control the molecular characteristic (Mn and Mw/Mn) of PLA chains extracted
from the so-prepared nanocomposite materials.
The results are shown in Table 3 (with an accuracy of 10%). Unprocessed PLA is characterized
by a Mn equal to 81,800. As revealed by SEC, upon
melt processing, PLA plasticized by 20 wt %
PEG1000 degraded noticeably (Mn drops by
33%). In turn, the Mn of PLA extracted from
plasticized nanocomposites is even more decreased and shows dependence on the clay content;
the weight loss reaching up to 50% at maximal
lling (10 wt %). However, the PLA degradation,
according to the Mn decrease, seems to be relatively less dependent on the type of organo-modication of the clay surface. Several factors such as
mechanical blending and shearing processes during compounding of PLA with clay, also the clay
content, and the presence of hydroxy groups on
the clay surface (upon organo-modication) can
contribute to the Mn reduction. The chemical
purity of PLA matrix, for example, the presence of
residual catalysts can also play a role in the reduction of molecular weight upon melt processing.
Nanostructure by X-Ray Diffraction
The nanocomposite morphology is considered as
a specic form of dispersion of the layered silicates within the polymer matrix combined with
the intercalation and/or exfoliation of the ller.
These phenomena are generated by shearing
processes during melt blending and prove to be
dependent on compatibility of the polymer matrix
and organo-modied surface of the clay. To determine the nanostructure in the relation to the
ller type and its content, an XRD study has
been performed.
Figure 1(ac) presents diffractograms recorded
from 18 to 108 of 2h for plasticized nanocomposites with increasing content of the ller : Cloi-

site1 20A, Cloisite1 25A, and Cloisite1 30B,


respectively. At low 2h angle region, all investigated nanocomposites are featured by two distinct diffraction peaks, whose intensity increases
with the ller content. The larger diffraction
peak localized at lower value of 2h (slightly above
28) is indicative for the intercalation of the layered silicate with the matrix constituents. The
second diffraction peak seen at 2h  58 corresponds with the second registry (002). The
increase of the interlayer distance (Dd)dened
as a difference between interlayer spacing in the
nanocomposite and in the corresponding organomodied claydepends on the clay type and on
its content, as far as nanocomposites lled with
Cloisite1 20A and Cloisite1 25A are concerned
[Dd values are given in Fig. 1(ac)]. For these
two organo-clays, the increase of the interlayer
distance Dd becomes smaller with the ller content increase: below 1.60 nm for the materials
containing Cloisite1 20A and below 1.34 nm for
the Cloisite1 25A-based nanocomposites. In contrast, for the nanocomposites containing Cloisite1 30B, the increase of the interlayer distance
is independent on the ller content used and it is
the largest with Dd 1.96 nm. This indicates
that Cloisite1 30B is more prone to easily intercalate plasticized PLA matrix than Cloisite1 20A
and Cloisite1 25A. Probably, prolongation of
blending time will lead to disappearance of concentration dependence of the Dd value for the
two later Cloisites.
Concerning the competition between PLA
chains and PEG1000 molecules in the intercalation process of the layered silicate, the comparison of the gallery space increases for the plasticized nanocomposite pN3A25 of value Dd
1.25 nm, and for nonplasticized nanocomposite
containing the same amount of Cloisite1 25A, for
which Dd 1.10 nm,9 tends to indicate the possible cointercalation of PEG1000 with PLA in this
particular clay. Since, for Cloisite1 20A and
Cloisite1 25A, Dd is varying with the relative

304

PLUTA ET AL.

amount of clay, one can assess that, depending on


the nanocomposite composition, the amount of
intercalated PLA and PEG1000 may vary from

Figure 2. Conventional DSC thermograms during


cooling from the melt followed by a heating scan
recorded at a rate of 5 8C/min for PEG1000 taken as
received.

one sample to the other. Interestingly, this Dd


variation is not observed for Cloisite1 30B. This
might arise from a better polarity matching
between the PEG1000 and the bis(hydroxyl)
functionalized ammonium cation of this particular organo-clay, stabilizing the plasticizer within
the clay galleries.
Thermal Properties
The thermal properties were determined for
PEG1000 plasticizer, neat PLA, and pPLA-based
compositions. Figure 2 shows conventional DSC
thermograms obtained for PEG1000 on cooling
scan followed by the heating one at a rate of 5 8C/
min. No clear glass transition can be observed,
even though low temperature (down to 80 8C)
has been reached. This can be explained by the
high crystallinity of this sample. The value of the
glass-transition temperature for PEG1000
(Tg(PEG)) was determined to be 61.0 8C.28 As is
seen in Figure 2 PEG1000 crystallizes on cooling
from the melt between 20 and 30 8C (with the
peak of the exotherm at Tc 27.5 8C); then, during the heating run, it melts from about 20 to
40 8C (with the peaks at Tm1 28.6 8C and at
Figure 1. (a) X-ray diffractograms for plasticized
nanocomposites containing 1, 3, and 10 wt % of Cloisite1 20A, respectively, (b) X-ray diffractograms for
plasticized nanocomposites containing 1, 3, 5, and 10
wt % of Cloisite1 25A, respectively, and (c) X-ray diffractograms for plasticized nanocomposites containing
1, 3, 5, and 10 wt % of Cloisite1 30B, respectively.

PLASTICIZED PLA/CLAY NANOCOMPOSITES

305

Tm2 35.6 8C). These transitions, close to the


ambient temperature, justied storage of the
samples at 4 8C.
Figure 3(ac) presents TMDSC thermograms
obtained for plasticized PLA nanocomposites containing 1, 3, 5, and 10 wt % Cloisite1 25A,
respectively, as well as for plasticized PLA and
neat PLA for the sake of comparison. Since other
nanocomposites containing Cloisite1 20A or Cloisite1 30B exhibit very similar thermograms compared with Figure 3(ac), their TMDSC characterizations are not included. Figure 3(a) shows
total heat ow thermograms, equivalent to conventional DSC thermograms. Neat PLA is characterized by a glass transition (Tg(PLA)) close to
50 8C, a cold crystallization exotherm at Tc2
(87.0 8C), a premelting crystallization at Tc3
(158.6 8C), and nally, a melting endotherm at
Tm equal to 173.4 8C. For the plasticized sample
pPLA (without ller), the total thermogram is
more complex. Indeed, for this sample, the glass
transition appears at lower temperature (around
23.4 8C), and contrary to neat PLA, it is immediately followed by a weak exothermic peak at Tc1
(33.8 8C), at which, interestingly, the plasticizer
should already be molten. At a slightly higher
temperature, a very weak and broad endotherm
appears that partially compensates the low temperature shoulder of the cold crystallization peak
at Tc2 (around 88.3 8C). Therefore, the resultant
cold crystallization peak for pPLA sample appears at higher Tc2 than those observed for neat
PLA and for the plasticized nanocomposites as
discussed below. It is also seen that plasticization
decreases Tm by 5 8C. In turn, the total thermograms for the plasticized nanocomposites are featured by the absence of a weak endotherm
between Tc1 and Tc2 seen only for pPLA and by
the shift of the cold crystallization peak toward
higher Tc2 with the ller content.
Figure 3(b,c) shows separated TMDSC thermograms, the reversing signal (R), and nonreversing
signal (NR), respectively. The reversing signal
clearly shows jump characteristic for the glass
transition of the PLA matrix Tg(PLA). The values

Figure 3. TMDSC thermograms for plasticized


PLA-based nanocomposites containing 1, 3, 5, and 10
wt % Cloisite1 25A, respectively, and for plasticized
PLA and neat PLA: (a) total heat ow signals, (b)
reversing heat ow signals, and (c) nonreversing heat
ow signals. The thermograms are related to the total
mass of PLA systems.

306

PLUTA ET AL.

Table 4. Calorimetric Parameters Derived From the Total Heat Flow (Tc1, DHc1, Tc2, DHc2, Tm, and DHm) and
the Glass-Transition Temperature (Tg) Determined From the Reversing Signal for Nanocomposites, Plasticized
PLA and Neat PLA, as Determined by TMDSC (3 8C/min, Modulation Period 40 s, and Amplitude 0.318)
1st Cold Crystallization Peak

2nd Cold Crystallization Peak

Melting Peak

Sample

Tg (8C)

Tc1 (8C)

DHc1 (J/gPLA)

Tc2 (8C)

DHc2 (J/gPLA)

Tm (8C)

DHm (J/gPLA)

PLA
pPLA
pN1A20
pN1A25
pN1B30
pN3A20
pN3A25
pN3B30
pN5A25
pN5B30
pN10A20
pN10A25
pN10B30

49.7
23.4
28.1
28.3
25.5
28.2
25.8
27.2
28.5
26.3
27.6
28.5
28.2

33.8
45.2
45.3
45.7
45.7
44.8
45.1
45.7
47.5
47.2
55.1
49.6

9.0
9.8
10.6
9.5
11.3
11.6
9.4
11.3
10.7
12.9
12.1
16.3

87.0
88.3
82.9
83.7
83.1
85.0
85.5
84.2
84.5
84.2
85.8
84.8
86.8

34.3
19.9
17.4
20.5
23.8
21.3
23.4
22.3
22.6
18.7
18.1
18.7
23.3

173.4
168.5
170.0
170.0
169.8
169.2
170.3
168.9
169.4
168.5
168.5
170.1
167.2

47.5
58.3
56.4
58.0
58.3
55.8
58.2
56.1
57.5
56.6
55.6
55.5
61.5

Tg, glass-transition temperature determined from the reversing signal of the TMDSC; Tc1, Tc2, temperatures of the cold crystallization peaks; DHc1, DHc2, enthalpies of cold crystallization normalized to the unit mass of PLA matrix; Tm, temperature of
the melting peak; DHm, enthalpy of melting normalized to the unit mass of PLA matrix.

of Tg(PLA) determined from the reversing signal


are gathered in Table 4. In turn, the NR signal for
each plasticized PLA matrix reveals one additional
thermal event, in comparison to the corresponding
total thermogram, that is the premelting crystallization process at Tc3 from 130 to 150 8C [compare
Fig. 3(a) and (b)]. Insufcient deconvolution of the
R and NR signals at Tc2 region does not allow any
further interpretation on this transition.
In Table 4 are gathered calorimetric parameters
derived from the total thermogram (Tc1, Tc2, Tm,
DHc1, DHc2, and DHm) and from the reversible signal (Tg(PLA)) for all considered samples. Calorimetric parameters characterizing very weak thermal events occurring between Tc1 and Tc2 for pPLA
sample [Fig. 3(a)] and at Tc3 for all samples [Fig.
3(a,c)] are not accounted in Table 4 for the sake of
clarity. Plasticization reduces Tg(PLA) from 49.7 8C
for neat PLA to 23.4 8C for pPLA; however, this
decrease is slightly smaller (25 8C) for plasticized
nanocomposites without clear relation to the ller
type and related content. This stiffening action of
organo-clay particles toward plasticized PLA
matrix is also reected by a shift toward higher
temperature of the cold crystallization processes at
Tc1 and at Tc2 about 34 8C. This decrease in plasticization could be due to the inability of a part of
PEG1000 to interact with PLA because it is already
interacting with the organo-clays (either intercalated within or adsorb at their surface). Enthalpies

for these two cold crystallization processes are


somewhat higher (with exception of pN1A20) than
those calculated for the unlled pPLA sample (9.0
and 19.9 J/gPLA for DHc1and DHc2, respectively).
Consequently, Tm of plasticized nanocomposites is
slightly higher (with exception of pN10B30) than
in case of unlled pPLA. It is worth to note that the
enthalpy of the cold crystallization of neat PLA
(DHc2 34.3 J/gPLA) slightly exceeds the sum of
the cold crystallization enthalpies (DHc1 and DHc2)
of the PLA-based nanocomposites (with exception
of pN10B30), while for the melting enthalpy of the
samples compared above, the relation is opposite,
that is DHm for all plasticized systems is higher
than for neat PLA. This observation is more likely
explained by the fact that the aforementioned values were determined from the total heat ow,
which therefore do not take into account the contribution from the premelting crystallization effect
(Tc3), as revealed by the nonreversing signal [Fig.
3(c)]. Combination of all these observations clearly
indicates that the presence of organo-clays tends to
induce a better ability for the plasticized PLA to
crystallize upon heating.
Structural Reorganization
The study of structural reorganization of PLA
accompanying thermal transitions during TMDSC
measurements were completed by XRD experi-

PLASTICIZED PLA/CLAY NANOCOMPOSITES

307

of the cold crystallization at Tc2 and then because of


the premelting crystallization at Tc3 is the same.
Furthermore, no contribution from the crystalline
phase of the PEG component is observed (if measured below Tm(PEG)), the strongest diffraction
peaks of crystalline PEG1000 being found at 2h of
19.38, 23.58, and 26.58.

Viscoelastic Properties
The temperature dependencies of the storage
modulus E0 and loss modulus E@ for the nanocomposites containing Cloisite1 20A and for pPLA
and neat PLA samples are shown in Figure 5(a,b),
Figure 4. XRD diffractograms for nanocomposite
pN10A25 and reference samples, that is neat PLA
and pPLA, recorded at different temperatures as indicated within the gure.

ments performed for all samples at selected temperatures [indicated by arrows in Fig. 3(a)]. Figure 4
presents representative diffractograms for nanocomposite pN10A25 recorded at 20, 60, 130, and
156 8C as well as for the comparative samples, that
is pPLA and neat PLA, recorded at 20 8C (thus
below both Tg(PLA) and Tm(PEG)). The diffractogram of neat PLA shows (at 20 8C) a broad maximum around 2h  178 what conrms the amorphous structure of the starting sample. However,
for the pPLA sample, a sharpening of the peak
around 168 is observed at 20 8C. This indicates that
plasticized polylactide is susceptible to some structural ordering immediately after melt-quenching
procedure (even if the sample is hold below the
ambient temperature). Consequently, this poorly
ordered phase undergoes melting upon heating
between Tc1 and Tc2 [cf. Fig. 3(a)]. However, this
structural ordering is considerably reduced in plasticized nanocomposites as the diffractogram for the
pN10A25 exhibits (at 20 8C) amorphous spectrum
with peaks ascribed only to the organo-clay particles
(at 2.938, 5.698, 8.698, and 19.988 in Fig. 4). In this
sample, a poorly ordered phase develops at Tc1 [Fig.
3(a)] and its presence is reected by small peak at
2h  168 seen on diffractogram taken at 60 8C (Fig.
4). Diffractograms recorded at higher temperatures
for plasticized nanocomposites, at 130 (above Tc2)
and at 156 8C (close to Tc3), are similar and show
crystalline peaks at 2h of 14.78, 16.58, 18.88, and
22.28 attesting for the pseudo-orthorombic crystalline form of PLA matrix.29,30 This observation
shows that the crystalline form developed because

Figure 5. (a) E0 versus temperature for nanocomposites containing Cloisite1 20A (1, 3, and 10 wt %) and
for pPLA and neat PLA samples. (b) E@ versus temperature for nanocomposites containing Cloisite1 20A
(1, 3, and 10 wt %) and for pPLA and neat PLA samples.

308

PLUTA ET AL.

from the cold crystallization process detected by


the TMDSC and XRD. Above 100 8C, the E0 drops
again because of softening and melting of the PLA
matrix. The inuence of the plasticizer on E0 curve
depends on the temperature region. At very low
temperature, glassy PEG reinforces PLA matrix
as demonstrated by the E0 value displayed by the
pPLA sample (this value is even higher than for
neat PLA). Above the glass transition of PEG1000
(Tg(PEG) 61.0 8C), E0 of pPLA sample rapidly
decreases reaching minimal values located at
lower temperature range (5080 8C) as compared
with neat PLA. Then E0 of pPLA increases because
of the cold crystallization process in the tempera-

Figure 6. (a) E0 versus temperature for nanocomposites containing Cloisite1 25A (1, 3, 5, and 10 wt %)
and for pPLA and neat PLA samples. (b) E@ versus
temperature for nanocomposites containing Cloisite1
25A (1, 3, 5, and 10 wt %) and for pPLA and neat
PLA samples.

respectively. Analogous dependencies for the systems containing Cloisite1 25A and Cloisite1 30B,
including samples pPLA and PLA, are shown in
Figures 6(a,b) and 7(a,b), respectively.
Storage Modulus
Neat PLA shows a gradual decrease of E0 with the
temperature increase from 95 to 50 8C; then, it
rapidly drops because of the glass transition and
reaches a minimal value around 75 8C. By further
increasing the temperature, E0 increases achieving a maximal value around 100 8C. This reects
an enhancement of the sample rigidity resulting

Figure 7. (a) E0 versus temperature for nanocomposites containing Cloisite1 30B (1, 3, 5, and 10 wt %)
and for pPLA and neat PLA samples. (b) E@ versus
temperature for nanocomposites containing Cloisite1
30B (1, 3, 5, and 10 wt %) and for PLAp and PLA
samples.

PLASTICIZED PLA/CLAY NANOCOMPOSITES

Table 5. Values of the Storage Modulus Determined


at Selected Temperatures by DMTA (1 Hz, 3 8C/min)
Value of the Storage Modulus E0
(MPa) Determined at Indicated
Temperature
Sample

75 (8C)

0 (8C)

75 (8C)

100 (8C)

PLA
pPLA
pN1A20
pN3A20
pN10A20
pN1A25
pN3A25
pN5A25
pN10A25
pN1B30
pN3B30
pN5B30
pN10B30

1055
1261
1751
2867
2666
1544
1850
1864
1732
1020
1654
2087
1782

771
567
870
1060
876
567
836
800
856
440
760
858
511

9
10
9
27
29
10
15
27
37
10
12
21
31

63
83
48
145
70
102
103
121
56
108
119
150
110

ture region characteristic for the neat PLA. Above


100 8C, it decreases nally. Plasticized nanocomposites exhibit a similar behavior. However, their
storage modulus increases with the ller content
and seems to be inuenced by the molecular
weight of the polymer matrix. The stiffening effect
is especially observed at lower temperature and it
is preserved in the positive temperature range
where the PLA matrix undergoes recrystallization. In Table 5 are compared the E0 values for all
samples, determined at selected temperatures,
that is at 75 8C (below Tg(PEG)), at 0 8C
(between Tg(PEG) and Tg(PLA)), at 75 8C (at
which the material exhibits the highest susceptibility for deformation), and at 100 8C (above the
cold crystallization process at Tc2). From these
data, it can be observed that, at 75 and 0 8C, E0
systematically increases with the ller content at
least up to 5 wt %. This increase is the most pronounced for the nanocomposites lled with Cloisite1 20A. At 10 wt % loading, the E0 increase
appears lower than at 5 wt %, most probably due
to the reduced molecular weight of the PLA
matrix of the formed nanocomposite (Table 3).
Loss Modulus
Figures 5(b) and 7(b) show the temperature dependencies of the mechanical loss (E@) for the
same set of samples as considered earlier. The
neat PLA sample does not reveal any loss maxi-

309

mum in temperature region below its Tg(PLA).


Actually, the Tg(PLA) value is determined in
DMTA at a maximum of the E@ localized at
51 8C, that is at nearly the same temperature as
those determined from the reversing signal of
the TMDSC measurement (49.7 8C, Table 4).
Consequently, the loss maximum (E@) connected
with the glass transition of all plasticized samples is shifted toward lower temperature. However, the plasticized polylactide (pPLA sample)
exhibits additional enhancement of the E@ in a
broad temperature range, displaying a maximum around 25 8C, that is well above Tg(PEG).
This maximum in E@ is ascribed to the glass
transition of the PEG-rich dispersed phase and
indicates that the polymer system is partially
miscible. Interestingly, the magnitude of the discussed mechanical losses is increased within
plasticized nanocomposites. Generally, higher
ller content leads to higher mechanical loss
(note that this relation is not fullled for samples
pN1B30, pN10B30, and pN10A25). Moreover,
the E@ maximum at Tg(PEG) region is better
developed for these nanocomposites in which the
intercalation effect proved to be more pronounced, that is for those containing Cloisite1
30B and Cloisite1 20A. This response actually
conrms that the plasticizer molecules readily
cointercalate the organo-modied layered silicate, together with PLA. In other words, higher
ller contents allow for entrapping larger
amounts of PEG molecules within the silicate
galleries. This contributes, respectively, to the
enhancement of the plastic response of the intercalated regions to the deformation in the temperature range around Tg(PEG). Contrary, under
absence of plasticizer, the mechanical loss of
PLA-based nanocomposites is lower than that of
unlled PLA15 in the considered temperature
range. The latter observation supports the interpretation done earlier.

CONCLUSIONS
PLA-based nanocomposites plasticized with 20
wt % PEG1000 were prepared by melt blending.
Three types of plasticized nanocomposites containing from 1 to 10 wt % of different organomodied layered silicates, that is Cloisite1 20A,
Cloisite1 25A, and Cloisite1 30B, were considered and compared with plasticized PLA and
neat PLA as references. Plasticizing was applied
as a method for reducing the brittleness of the

310

PLUTA ET AL.

PLA matrix. For the same reason, the meltquenching was used to promote the formation of
amorphous PLA matrices in all samples.
It is shown that plasticization reduces the
Tg (PLA) by about DTg 26.3 8C in case of unlled PLA, while for the PLA-based nanocomposites, the reduction was somewhat smaller, DTg
 21 8C. Both PEG and PLA seem to intercalate
within the layered silicate galleries (semiexfoliation can not be excluded). The way PLA and
PEG1000 intercalate seems to be dependent of
the organo-modication of the selected clay.
TMDSC and XRD analyses have demonstrated
that the physical organization of the investigated
samples was thermally unstable. Indeed, the
structure of neat PLA undergoes transformation
through a cold crystallization process into the
crystalline form classied as the a modication.
Plasticized PLA and plasticized nanocomposites
revealed the occurrence of some additional structural transition at lower temperature, slightly
above Tg(PLA), connected with semiordering of
the structure. This process is shifted toward
higher temperature with the ller content increase and it is completed by the cold crystallization process responsible for the formation of the
crystalline PLA phase. The presence of the plasticizer as well as lling with different type and content of clays does not modify the crystalline evolution of the PLA matrix.
Viscoelastic measurements revealed for the
neat PLA one strong maximum of the mechanical
loss (E@) ascribed to the glass transition. The
glass transition of PLA matrix determined from
the E@ maximum and from the reversing signal
in TMDSC were comparable. In the case of plasticized PLA, an additional E@maximum assigned to
the glass transition of the PEG-rich dispersed
phase was found (negative temperature region).
This mechanical loss also occurred within plasticized nanocomposites and exhibited a tendency
to increase with the ller content. This increase
was somewhat correlated with the intercalation
magnitude, giving an evidence on the ability of
the PEG molecules to penetrate the silicate gallery. The storage modulus of the nanocomposites
exceeds those for the neat PLA, while dispersed
PEG acts as a reinforcing agent for PLA in the
low-temperature region, below Tg(PEG).
The structural evolution of the plasticized
PLA-based nanocomposites as already pointed
out in this study has been the object of a more
detailed investigation actually carried out on a
much longer time period, that is over three years.

Such an aging study is reported in the second


part of this series.23
Contract grant sponsor: Poland State Committee for
Scientic Research; contract grant number: 7 T08E
027 19. XRD measurements were carried out by M.P.
at the Max-Planck-Institut fur Polymerforschung,
Mainz (Germany) in Prof. T. Pakula lab and were
made possible by the grant 7T0 8E 027 19. M.-A. Paul
thanks the F.R.I.A. (Fonds pour la Recherche Industrielle et Agricole) for her Ph.D. grant. M.A. and Ph.D.
are very grateful for the nancial support from
Region Wallonne and European Community (FEDER,
FSE) in the frame of Pole dExcellence Materia Nova.
LMPC thanks the Belgian Federal Government Ofce
of Science Policy (STC-PAI 5/3). CMMS and LMPC
thank the Polish Academy of Science and the Fonds
National de la Recherche Scientique (FNRS, Belgium)
for a grant allowing traveling between the research
laboratories.

REFERENCES AND NOTES


1. Pawlak, A.; Morawiec, J.; Piorkowska, E.; Galeski,
A. Solid State Phenomena 2003, 94, 335339. In
Interfacial Effects and Novel Properties of Nanomaterials; Lojkowski, W.; Blizzard, J. R., Eds.
2. Thostenson, E. T.; Ren, Z.; Chou, T.-W. Compos
Sci Technol 2001, 61, 18991912.
3. www.nanoclay.com (accessed 30 December 2004).
4. Alexandre, M.; Dubois, Ph. Mater Sci Eng 2000,
R28, 163.
5. Oriakhi, C. Chem Br 1998, 34, 5965.
6. Sinclair, R. G. J Macromol Sci Pure Appl Chem
1996, 33, 585597.
7. www.cargilldow.com (accessed 30 December 2004).
8. Ogata, N.; Jimenez, G.; Kawai, H.; Ogihara, T.
J Polym Sci Part B: Polym Phys 1997, 35, 389
396.
9. Pluta, M.; Paul, M.-A.; Alexandre, M.; Dubois,
Ph.; Galeski, A. J Appl Polym Sci 2002, 86, 1497
1506.
10. Paul, M.-A.; Alexandre, M.; Degee, Ph.; Henrist,
C.; Rulmont, A.; Dubois, Ph. Polymer 2003, 44,
443450.
11. Pluta, M. Polymer 2004, 45, 82398251.
12. Ray, S. S.; Maiti, P.; Okamoto, M.; Yamada, K.;
Ueda, K. Macromolecules 2002, 35, 31043110.
13. Ray, S. S.; Yamada, K.; Ogami, A.; Okamoto, M.;
Ueda, K. Macromol Rapid Commun 2002, 23, 943
947.
14. Ray, S. S.; Yamada, K.; Okamoto, M.; Ueda, K.
Nano Lett 2002, 2, 10931096.
15. Ray, S. S.; Yamada, K.; Okamoto, M.; Ueda, K.
Polymer 2003, 44, 857866.
16. Ray, S. S.; Yamada, K.; Okamoto, M.; Ogami, A.;
Ueda, K. Chem Mater 2003, 15, 14561465.

PLASTICIZED PLA/CLAY NANOCOMPOSITES

17. Labrecque, L. V.; Kumar, R. A.; Dave, V.; Gross,


R. A.; McCarthy, S. P. J Appl Polym Sci 1997, 66,
15071513.
18. Hu, Y.; Hu, Y. S.; Topolkaraev, V.; Hiltner, A.;
Baer, E. Polymer 2003, 44, 57115720.
19. Jacobsen, S.; Fritz, H. G. Polym Eng Sci 1999, 39,
13031310.
20. Martin, O.; Averous, L. Polymer 2001, 42, 6209
6219.
21. Ljungberg, N.; Wesslen, B. J. Appl Polym Sci
2002, 86, 12271234.
22. Celli, A.; Scandola, M. Polymer 1992, 33, 2699
2703.
23. Pluta, M.; Paul, M.-A.; Alexandre, M.; Dubois, Ph.
J Polym Sci Part B: Polym Phys, in press.

311

24. Degee, Ph.; Dubois, Ph.; Jerome, R. Macromol


Symp 1997, 123, 6784.
25. Thermal Analysis and Rheology, TA Instruments,
TN-45B.
26. Degee, Ph.; Dubois, Ph.; Jerome, R. Macromol
Symp 1997, 198, 19851995.
27. Cicero, J. A.; Dorgan, J. R.; Dec, S. F.; Knauss, D. M.
Polym Degrad Stab 2002, 78, 95105.
28. Brandrup, J.; Immergut, E. H.; Grulke, E. A. In
Polymer Handbook, 4th ed.; Wiley: New York,
1999.
29. DeSantis, P.; Kovacs, A. Biopolymers 1968, 6,
299306.
30. Grijpma, D. W.; Zondervan, G. J.; Pennings, A.
J Polym Bull 1991, 25, 327333.

Vous aimerez peut-être aussi