Vous êtes sur la page 1sur 7

Composites Science and Technology 70 (2010) 815–821

Contents lists available at ScienceDirect

Composites Science and Technology


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

Functionalized cellulose nanocrystals as biobased nucleation agents


in poly(L-lactide) (PLLA) – Crystallization and mechanical property effects
Aihua Pei a, Qi Zhou a,b,*, Lars A. Berglund a,c
a
Department of Fiber and Polymer Technology, Royal Institute of Technology, SE-100 44 Stockholm, Sweden
b
School of Biotechnology, Royal Institute of Technology, AlbaNova University Center, SE-106 91 Stockholm, Sweden
c
Wallenberg Wood Science Center, Royal Institute of Technology, SE-100 44 Stockholm, Sweden

a r t i c l e i n f o a b s t r a c t

Article history: The important industrial problem of slow crystallization of poly(L-lactide) (PLLA) is addressed by the use
Received 9 November 2009 of cellulose nanocrystals as biobased nucleation reagents. Cellulose nanocrystals (CNC) were prepared by
Received in revised form 26 January 2010 acid hydrolysis of cotton and additionally functionalized by partial silylation through reactions with n-
Accepted 27 January 2010
dodecyldimethylchlorosilane in toluene. Such silylated cellulose nanocrystals (SCNC) were dispersible
Available online 1 February 2010
in tetrahydrofuran and chloroform, and formed stable suspensions. Nanocomposite films of PLLA and
CNC or SCNC were prepared by solution casting. The effects of surface silylation of cellulose nanocrystals
Keywords:
on morphology, non-isothermal and isothermal crystallization behavior, and mechanical properties of
Cellulose nanocrystals
A. Nano composites
these truly nanostructured composites were investigated. The unmodified CNC formed aggregates in
A. Particle-reinforced composites the composites, whereas the SCNC were well-dispersed and individualized in PLLA. As a result, the tensile
E. Casting modulus and tensile strength of the PLLA/SCNC nanocomposite films were more than 20% higher than for
B. Mechanical properties pure PLLA with only 1 wt.% SCNC, due to crystallinity effects and fine dispersion.
Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction nucleation density and thus to enhance the crystallization rate is


the addition of heterogeneous nucleating agents [9]. To broaden
Poly(lactic acid) (PLA) is a biodegradable thermoplastic with the utilization of PLA, environmentally friendly and biocompatible
high-strength (50–70 MPa), high-modulus (3 GPa), and relatively nucleating agents having excellent nucleation ability without any
good biocompatibility. PLA has attracted much attention from mate- limitations or drawbacks are desired.
rial researchers and industry due to its potential in biomedical appli- Cellulose fibers have been used as a nucleating and reinforcing
cations, and as an engineering polymer based on renewable material in the lignocellulosic/thermoplastic composites, in partic-
resources. Because of the inherent brittleness and lower impact ularly the cellulose/polypropylene systems [10–13]. Transcrystal-
resistance of PLA, considerable efforts have been made to improve line morphology was observed on the surface of the cellulose
its properties in order to achieve the compatibility with thermoplas- fibers. In addition, a transcrystalline layer of polypropylene at the
tics processing and manufacturing, and compete with commodity edge of films cast from cellulose nanocrystal suspensions has been
polymers. Composites of PLA reinforced with natural and man-made demonstrated by Gray [14]. Cellulose nanocrystals are highly crys-
cellulose fibers [1] including regenerated cellulose fibers [2], wood talline cellulose particles produced by acid hydrolysis through par-
fibers [3], recycled newspaper fibers [4], plant fibers from kenaf tially dissolving cellulose fibrils [15]. Using nanoscale cellulose
[5], and microfibrillated cellulose (MFC) [6–8] have been prepared crystallites as fillers for PLA has advantages over inorganic fillers
and exhibited improved biodegradability, thermal stability, and because of their exceptionally unique characteristics such as bio-
mechanical properties. Such biocomposites are of great interest as degradability and biocompatibility, high stiffness, high aspect ratio
replacement of petroleum-based polymers since they are com- (length/diameter), and low density. A specific challenge to achieve
pletely based on renewable resources. However, the slow crystalli- maximum performance of nanocomposites from PLA and cellulose
zation of PLLA is a major problem industrially, since practical melt- nanocrystals is the difficulty to obtain well-dispersed cellulose
processing conditions results in products where the PLLA often is nanocrystals in the PLA polymer matrix.
in a state of very low crystallinity. An efficient method to increase In our previous work [16], a high-strength elastomeric nanocom-
posite has successfully been prepared by dispersing microcrystalline
cellulose (MCC) in a polyurethane matrix with the addition of a small
* Corresponding author. Address: School of Biotechnology, Royal Institute of amount of lithium chloride (LiCl). Biodegradable nanocomposites
Technology, AlbaNova University Center, SE-106 91 Stockholm, Sweden. Tel.: +46 8
from PLA and MCC have been prepared previously [17,18]. MCC
55378383; fax: +46 8 55378468.
E-mail address: qi@kth.se (Q. Zhou). has higher capacity to act as a nucleating agents compared to cellu-

0266-3538/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compscitech.2010.01.018
816 A. Pei et al. / Composites Science and Technology 70 (2010) 815–821

lose fibers due to the surface topography of the fiber [18]. MCC cellu- temperature. The dissolved solution was poured onto a leveled
lose fibrils remained as aggregates in the composites and had poor glass Petri dish with a diameter of 5 cm, and then allowed to dry
adhesion with the PLA matrix, which resulted in poor mechanical for about 24 h at room temperature. For the preparation of the
properties [17]. In order to improve the dispersion of MCC in PLA, nanocomposite films, a predetermined amount of SCNC suspension
MCC was treated with N,N-dimethylacetamide (DMAc) containing in chloroform (20 mg/mL) was mixed with the previously prepared
LiCl to swell and separate the cellulose whiskers. The mechanical PLLA solution. The mixture solutions were stirred for 4 h before
properties of the PLA/MCC composite were improved but DMAc/LiCl they were cast onto the glass Petri dish and dried. The resultant
caused degradation of the composite at high temperature processing films with a thickness of ca. 70 lm were peeled from the glass Petri
[19]. Further, cellulose nanowhiskers (CNW) were produced by acid dish after drying. PLLA/SCNC nanocomposite films with 1 wt.% and
hydrolysis of MCC. An anionic surfactant [20], the acid phosphate es- 2 wt.% of SCNC were prepared and coded as PLLA–SCNC-1 and
ter of ethoxylated nonylphenol, was utilized to improve the disper- PLLA–SCNC-2, respectively. For comparison, PLLA/CNC nanocom-
sion of CNW in the PLA matrix [21,22]. However, increased posite films with 1 wt.% and 2 wt.% of CNC were prepared
surfactant content caused degradation of PLA. following the same procedures as described above and coded as
In this study, cellulose nanocrystals were partially surface sily- PLLA–CNC-1 and PLLA–CNC-2, respectively. The pure PLLA film
lated to improve their dispersion in PLLA. The PLLA/cellulose nano- was coded as PLLA. All films were dried in vacuum oven at room
crystals composite films containing 1 wt.% and 2 wt.% of cellulose temperature for 48 h to constant weight before further analysis.
were prepared by the solution casting method. The effect of surface
silylation of cellulose nanocrystals on morphology, non-isothermal
and isothermal crystallization behaviors, and mechanical proper- 2.4. Characterization
ties of the nanocomposites were investigated by field emission
scanning electron microscopy (FE-SEM), polarized optical micros- Fourier transform infrared spectroscopy (FTIR) were conducted
copy (POM), differential scanning calorimetry (DSC), and tensile on a Perkin–Elmer Spectrum 2000 FTIR equipped with a MKII Gold-
test. en Gate, single reflection attenuated total reflectance (ATR) system
(Specac Ltd., London, UK). The ATR crystal was a MKII heated dia-
2. Experimental mond 45° ATR top plate. To avoid the superposition of many layers
of nanocrystals, a drop of the diluted CNC or SCNC suspensions in
2.1. Materials THF (0.1% w/w) was mounted directly on a carbon-coated electron
microscopy grid. The specimens were observed using a Philips Tec-
Poly(L-lactide) (PLLA, BiomerÒ L9000; weight average molecular nai 10 electron microscope operated at 80 kV. Although this meth-
mass, Mw = 220 kDa, number average molecular mass, Mn = od can create some drying artifacts, it is commonly used for the
101 kDa) was obtained from Biomer Inc. (Krailling, Germany). PLLA observation of cellulose nanocrystals. A Hitachi S-4800 scanning
resins were dried in a vacuum oven at 60 °C for 24 h before use. electron microscope was used to capture secondary electron
The cellulose nanocrystal suspension was prepared from cotton images of the cross section of the nanocomposite films. The films
cellulose by sulfuric acid hydrolysis, following the recipe used by were frozen in liquid nitrogen and fractured. The specimens were
Cranston and Gray [23]. fixed on a metal stub using carbon tape and coated with a dou-
ble-layer coating (5 nm) consisting of graphite and gold–palladium
2.2. Partial silylation of the cellulose nanocrystals using Agar HR sputter coaters. Differential Scanning Calorimetry
(DSC) measurements were taken on a Mettler Toledo DSC 820
The cellulose nanocrystals were partially silylated following the using nitrogen as the purge gas. Samples of about 7.0 mg were en-
route developed by Gousse et al. [24]. CNC in aqueous suspension closed in aluminum pans. In the non-isothermal procedure, the
(100 mg/mL) were solvent exchanged to acetone and then to dry sample was first heated to 210 °C and held for 5 min, and then it
toluene in which they were precipitated because of the aggregates was cooled to 0 °C at a heating and cooling rates of 10 °C/min.
formation. The n-dodecyldimethylchlorosilane (DDMSiCl) (Aldrich) The melting temperature (Tm) and the enthalpy of cold crystalliza-
was then added with a molar ratio of 3:1 for DDMSiCl and anhy- tion and melting (DHcc and DHm, respectively) were calibrated
droglucose units (AGU) in cellulose. Imidazole used for trapping using indium and tin as standards. For isothermal crystallization,
the HCl released during the reaction was added in quantity equi- all samples were heated from room temperature to 210 °C and kept
molar to that of DDMSiCl. The graft reaction was conducted for for 5 min to completely eliminate any possible crystalline phase.
12 h at room temperature (22 °C) under vigorous stirring, and then Then the samples were quenched to a desired isothermal crystalli-
was terminated by adding a mixture of tetrahydrofuran (THF) and zation temperature (125 °C, 120 °C, 115 °C, and 110 °C) and held
methanol (80/20, v/v). The partially silylated cellulose nanocrystals for sufficient time to allow crystallization completely from the qui-
product was washed thoroughly with THF twice to remove the escent melt. The exothermic curves of heat flow as a function of
impurities (e.g. imidazolium chloride and disilyl ether) by centrifu- time were recorded. Polarized Optical Microscopy (POM) was car-
gation at 12,000g for 10 min at 4 °C and subsequent dispersion in ried out on a Leitz Ortholux POL BKII optical microscope equipped
fresh THF. Such partially silylated cellulose nanocrystals (SCNC) with a Mettler Toledo Hot Stage FP82 controlled by Mettler FP90
can be suspended in organic solvents such as THF and chloroform Central Processor. Samples were prepared by melting a sample be-
and form stable homogeneous suspensions after sonication (Bran- tween two cover slips on a hot plate and gently pressing them to-
son SonifierÒ Model 250 ultrasonic cell disruptor/homogenizer) at gether with tweezers. Isothermal crystallization was recorded by
80% output (with cooling in an ice bath) for 1 min, while unmodi- heating the sample to 210 °C and holding for 5 min, then quench-
fied CNC aggregated and precipitated from THF and chloroform ing (20 °C/min) to 125 °C. Pictures were taken at specified inter-
solution after sonication. vals. The mechanical properties of composites were measured
using a horizontal Miniature materials tester (MiniMat 2000) with
2.3. Preparation of PLLA/cellulose nanocrystals composites a 200 N load cell at a cross-head speed of 2 mm/min. The gauge
length is 10 mm. Samples were cut into rectangular pieces with
PLLA/cellulose nanocrystals nanocomposite films were pre- width of 3 mm. The results obtained from the Minimat 2000 can
pared by the solution casting method [25]. PLLA (0.2 g) was dis- only be used for comparison the samples in this study, because
solved in 2 mL chloroform with vigorous stirring at room the values are based on the rotational movement of the drive shaft.
A. Pei et al. / Composites Science and Technology 70 (2010) 815–821 817

3. Results and discussion

3.1. Surface modification of cellulose nanocrystals

Cellulose nanocrystals are very difficult to disperse in a hydro-


phobic polymeric matrix as they have a large surface area and pos-
sess large hydrogen forces among themselves. Cellulose
nanocrystals can be dispersed in a polar organic solvent, N,N-
dimethylformamide (DMF) [26]. In addition, stable suspensions
of cellulose nanocrystals dispersed in organic medium have been
achieved by using a surfactant [20], or by surface chemical modifi-
cation, e.g. partially silylation [24]. As inspired by the PLA/clay
nanocomposites [27,28], in which organically modified montmo-
rillonite (MMT) was often applied to increase the hydrophobicity
and enhance the dispersion, we employed the surface chemical
modification method to disperse cellulose nanocrystals in organic
solvents and the PLLA matrix.
The FTIR spectra of unmodified and partially silylated cellulose
nanocrystals are shown in Fig. 1. The band at around 1050 cm–1,
which is the most intensive in the spectrum of CNC (Fig. 1a), has
been identified in the SCNC sample (Fig. 1b) as corresponding to
the pyranose ring ether band of cellulose. The successful silylation
of cellulose nanocrystals was proved by the presence of the charac-
teristic bands at 779 cm–1 and 840 cm–1 that were attributed to Si–
CH3 stretching and bending, respectively [29]. In addition, new
bands at 2922 cm–1 (asymmetrical CH2 stretching vibration) and
2852 cm–1 (symmetrical CH2 stretching vibration) introduced by
the silylating agent n-dodecyldimethylchlorosilane were also Fig. 2. Transmission electron micrographs of (a) CNC and (b) SCNC from THF
detected. suspension (scale bar, 1000 nm).
The dispersion quality of SCNC in non-polar organic solvents
was confirmed by the visualization of the individualized nanocrys-
tals using transmission electron microscopy (TEM). The formation in the form of bundles is a typical feature observed previously
of aggregates was observed for unmodified cellulose nanocrystals for cotton microcrystals in aqueous suspensions [30].
as shown in Fig. 2a. After surface silylation, a good dispersion
was observed for the cellulose nanocrystals suspension in THF.
3.2. Morphology of nanocomposite films
Fig. 2b displays the TEM micrograph of SCNC dried from dilute
THF suspension. The nanocrystals were well individualized with
Fig. 3 shows FE-SEM images of the cryo-fractured surfaces of
typical dimensions ranging mostly from 200 to 300 nm in length
pure PLLA and cellulose nanocrystals nucleated PLLA composite
and around 15 nm in width. The level of dispersion of SCNC in
films. In the PLLA–CNC-2 composite with unmodified CNC
THF is better than the surfactant assisted dispersion of cotton cel-
(Fig. 3b), the cellulose nanocrystals were agglomerated in the PLLA
lulose nanocrystals in toluene, in which bundles of nanocrystal
matrix. Fig. 3c shows that the dispersion of cellulose nanocrystals
aggregates were observed [20]. Aggregates of a few nanocrystals
in PLLA was dramatically improved with surface partial silylation
on CNC. The surface of the PLLA–SCNC-2 composite sample was
smooth as the pure PLLA surface (Fig. 3a) and no visible micro-
100
scale aggregations of SCNC were observed. Such morphology is dif-
a ferent from the cryo-fractured surfaces of tunicin whiskers filled
Transmittance (%)

80 nanocomposites prepared from a DMF solution, where tunicin


whiskers appear like white dots corresponding to the transversal
60
sections of the cellulose whiskers [26]. This is due to the fact that
40 the length of tunicin cellulose whisker is in the range of microme-
ters, while it is only 200–300 nm for cellulose nanocrystals. In con-
20 clusion, the smooth fracture surface in Fig. 3c indicates the
4000 3500 3000 2500 2000 1500 1000 600
presence of well-dispersed SCNC particles and successful prepara-
100
tion of a truly nanostructured composite based on PLLA and SCNC.
Transmittance (%)

b
80
3.3. Crystallization and melting behaviors
60
The non-isothermal crystallization and melting behavior of the
40 PLLA/cellulose nanocrystal composites were investigated by DSC.
In the heating scans (Fig. 4a), all samples showed a distinct exo-
20
4000 3500 3000 2500 2000 1500 1000 600 thermic peak, attributing to the cold crystallization during heating
process. The melting temperatures (Tm) of PLLA in the nanocom-
Wavenumber (cm -1)
posites were similar to that of pure PLLA. For semicrystalline poly-
Fig. 1. FTIR spectra of (a) cellulose nanocrystals (CNC) and (b) cellulose nanocrys- mers, crystallinity may strongly influence mechanical properties.
tals that were partially silylated with n-dodecyldimethylchlorosilane (SCNC). The degree of crystallinity (Xc) estimated from DSC curves for PLLA
818 A. Pei et al. / Composites Science and Technology 70 (2010) 815–821

PLLA

PLLA-CNC-1

PLLA-CNC-2

Endo
PLLA-SCNC-1

PLLA-SCNC-2

0 50 100 150 200


Temperature (oC)

PLLA

PLLA-CNC-1
Endo PLLA-CNC-2

PLLA-SCNC1

PLLA-SCNC2

0 50 100 150 200


Temperature (oC)

Fig. 4. DSC (a) heating and (b) cooling scans of PLLA and PLLA/cellulose nanocrys-
tals nanocomposites.

Table 1
Fig. 3. FE-SEM micrographs of the cryo-fractured surfaces of (a) pure PLLA film, (b) Thermal parameters of PLLA and PLLA/cellulose nanocrystals composites derived
PLLA–CNC-2 and (c) PLLA–SCNC-2 nanocomposite films. from the heating DSC scans.

Tm (°C) DHm (J/g)a DHcc (J/g)a Xc (%)


PLLA 171.4 25.8 12.5 14.3
and PLLA/cellulose nanocrystals composites are summarized in Ta-
PLLA–CNC-1 171.4 28.2 12.9 16.4
ble 1. The following equation was used to calculate Xc, PLLA–CNC-2 171.8 30.0 16.0 15.0
PLLA–SCNC-1 171.8 32.3 4.0 30.4
PLLA–SCNC-2 171.4 29.4 10.1 20.7
DHm  DHcc
Xc ¼  100% ð1Þ
DH 1
m
a
Data corrected for the percentage of PLLA in the nanocomposites.

where DHm is the measured endothermic enthalpy of melting and The nucleation effect was remarkably enhanced when homoge-
DHcc is the exothermic enthalpy that is absorbed by the crystals neous cellulose nanocrystals dispersion in PLLA was achieved.
formed during the DSC heating scan. The theoretical melting enthal-
py of 100% crystalline PLLA was taken (DH1 m = 93.0 J/g) [28]. Com- 3.4. Polarized optical microscopy observations
pared with the pure PLLA sample, the Xc values of PLLA in the
PLLA/CNC nanocomposites were increased slightly, while the Xc val- The effect of the addition of unmodified and surface silylated cel-
ues of PLLA in the PLLA/SCNC nanocomposites were increased sig- lulose nanocrystals in the PLLA matrix as nucleation agents on the
nificantly by using surface silylated cellulose nanocrystals as the crystallization behavior were observed by polarized optical micro-
nucleating agents, e.g. from 14.3% to 30.4% for PLLA–SCNC-1. This scope. Optical micrographs were taken on the 0, 5th and 10th min
result is comparable with the composite of PLA containing 25% at 125 °C after quenched from melt at 210 °C for the PLLA, PLLA–
(by weight) MCC where an increase from 19.2% to 44.5% was re- CNC-1 and PLLA–SCNC-1 samples (Fig. 5). Both PLLA and PLLA–
ported [18]. In the cooling scans (Fig. 4b), the exothermic peaks SCNC-1 samples showed clean and uniform melt, while micro-scale
with very low intensity were observed for the pure PLLA and aggregates of cellulose nanocrystals were found in the PLLA–CNC-1
PLLA/CNC nanocomposite samples, indicating a rather low crystalli- sample before crystallization (0 min). This further indicated that
zation capability. As for the PLLA/SCNC nanocomposite samples, the surface silylated cellulose nanocrystals were better dispersed in
crystallization peaks had relatively higher intensity and started the PLLA matrix compared to the unmodified ones. The processes
from higher temperature as a result of enhanced crystallizability. of nucleation and growth control the crystallization kinetics. When
A. Pei et al. / Composites Science and Technology 70 (2010) 815–821 819

Fig. 5. Polarized optical microscope images of PLLA, PLLA–CNC-1, and PLLA–SCNC-1 acquired on the 0, 5th and 10th min at 125 °C after quenched from melt at 210 °C. Scale
bar, 200 lm.

the heterogeneous nucleating agent is added, an increase in the To describe the isothermal crystallization kinetics, the classical
overall crystallization rate occurs with a reduction of the nucleation Avrami equation was employed:
induction period and also an increase in the number of primary
nucleation sites [31]. With the addition of CNC, the nucleus density log½ lnð1  X t Þ ¼ log k þ n log t ð3Þ
of PLLA crystallites increased at 5 min and 10 min as compared to the
pure PLLA. Furthermore, the nucleus density increased significantly where k is the overall kinetic constant depending on the geometry
with the addition of SCNC, due to improved dispersion of cellulose of the growing crystalline phase, and n is the Avrami exponent cor-
nanocrystals in the PLLA matrix. As a result, more crystals were able relating with the nucleation mechanism and crystal growth dimen-
to nucleate and grow on the increased surface area of the interfaces sion. The values of log[ln(1  Xt)] were plotted versus log t, and
due to increasing numbers of nucleating particles. Similar phenom- the Avrami exponent n and the crystallization rate constant k were
ena were reported by Gurato et al. [32], who found that smaller talc calculated from the slope and intercept of the linear fit. The Avarmi
particles were more effective in increasing crystallization rate than equation rarely describes the whole crystallization process and is
larger ones for nylon 6. usually valid at the primary crystallization stage [33]. Thus, the ini-
tial linear portions of the plots were analyzed and the results are
3.5. Isothermal crystallization kinetics summarized in Table 2. The k values of the PLLA/SCNC samples were
higher compared to those of the PLLA and PLLA/CNC samples, in
In order to fully understand the effect of the addition of unmod- particular, the k values increased by almost one order of magnitude
ified and surface silylated cellulose nanocrystals on the bulk crys- at Tc of 125 °C. The Avrami exponent, n, relates to the nature of
tallization rate of PLLA, isothermal crystallization behaviors of the nucleation and the dimensionality of the growing crystals. The n
PLLA/cellulose nanocrystals nanocomposites were also studied. values of PLLA were around 2.8–3.3, suggesting a three-dimensional
The isothermal crystallization exotherms of (a) PLLA, (b) PLLA– crystallization growth and homogeneous nucleation mechanism
CNC-1, and (c) PLLA–SCNC-1 under different crystallization tem- [27,34]. The obtained n values were similar to those reported for
perature (Tc) are shown in Fig. 6. Within the same sample, the shift pure PLLA (2.3–3.2 at Tc = 90–130 °C by Tsuji et al. [35], and 2.8–
of exotherms to longer time with increasing Tc indicated a decrease 3.2 at Tc = 90–130 °C by Iannace and Nicolais [36]). The n values
of the crystallization rate. As for the same Tc, it is clear that the of PLA/SCNC biocomposites were around 3.5–4.9, indicating a crys-
crystallization time of PLLA–SCNC-1 (Fig. 6c) was much shorter tallization mode of heterogeneous nucleation [37,38]. Similar n val-
than that of PLLA and PLLA–CNC-1. These results indicated that ues were reported for PLA/nucleating agent (BaSO4, CaCO3, and
the crystallization rate of PLLA increased with surface silylated cel- TiO2) composites, where the Avrami exponent n is in the range of
lulose nanocrystals as nucleating agents. 4.0–5.2 [39].
The relative crystallinity (Xt) as a function of crystallization time Half-time of crystallization (t1/2) is defined as the time when Xt
(t) is calculated as follows: is equal to 0.5 and is calculated as follows [40]:
Rt  1=n
ðdH=dtÞdt ln 2
0
Xt ¼ R 1 ð2Þ t1=2 ¼ ð4Þ
ðdH=dtÞdt k
0
820 A. Pei et al. / Composites Science and Technology 70 (2010) 815–821

Table 2
a Avrami kinetic parameters for the isothermal crystallization of the PLLA and PLLA/
cellulose nanocrystals composites.

Sample Tc (°C) n k (minn) t1/2 (min)


PLLA 110 2.81 1.62  10–3 8.6
115 2.92 9.12  10–4 9.7
120 3.16 7.94  10–5 17.7
125 3.28 1.25  10–5 27.9
o
Endo

125 C
PLLA–CNC-1 110 3.47 3.16  10–4 9.2
115 3.34 3.09  10–4 10.1
120 3.32 1.99  10–4 11.7
o 125 3.24 6.02  10–5 17.9
120 C
o
o
115 C PLLA–CNC-2 110 3.30 3.82  10–4 9.7
110 C 115 3.15 3.78  10–4 10.9
120 3.24 1.78  10–4 12.8
125 3.31 4.07  10–5 19.0
10 20 30 40 50 60
PLLA–SCNC-1 110 4.07 2.09  10–3 4.2
Crystallization time (min) 115 4.13 1.02  10–3 4.9
120 4.26 4.68  10–4 5.5
125 3.61 4.17  10–4 7.8
b
PLLA–SCNC-2 110 4.91 1.14  10–3 5.9
115 4.35 7.76  10–4 4.8
120 4.08 4.78  10–4 6.0
125 3.82 2.18  10–4 8.3
Endo

o
125 C
Table 3
o
120 C Mechanical properties of PLLA and cellulose nanocrystals nucleated PLLA composite
films at room temperature.
o
115 C
Tensile Tensile Elongation
o modulus (GPa) strength (MPa) at break (%)
110 C
PLLA 1.1 ± 0.01 48.3 ± 2.9 31.1 ± 3.0
PLLA–CNC-1 1.0 ± 0.02 49.2 ± 0.4 10.5 ± 2.0
10 20 30 40 50 60 PLLA–CNC-2 1.2 ± 0.03 48.3 ± 0.4 12.2 ± 0.2
PLLA–SCNC-1 1.4 ± 0.08 58.6 ± 3.1 8.3 ± 0.6
Crystallization time (min) PLLA–SCNC-2 1.4 ± 0.04 53.8 ± 2.1 7.1 ± 1.8

c
the micro-scale agglomeration of CNC as shown in Fig. 2a, did
not show improvements in tensile modulus and tensile strength
compared to pure PLLA. Meanwhile, the addition of only 1 wt.%
of SCNC in PLLA resulted in a 27% increase in tensile modulus
o
and a 21% increase in the tensile strength compared to pure PLLA.
Endo

125 C
The main reason is increased degree of crystallinity (see Table 1),
and the associated truly nanostructural characteristics resulting
120 C
o from more efficient dispersion and alignment of the functionalize
115 C
o nanocrystals (SCNC) as shown in Fig. 2b. The elongation to break
110 C
o is affected by the volume fraction of the added reinforcement,
the dispersion of the reinforcement in the matrix, and the interac-
tion between the reinforcement and the matrix. The elongation at
10 20 30 40 50 60 break was reduced in both PLLA/CNC and PLLA/SCNC composite
Crystallization time (min) samples. The reason is that stiff reinforcements of cellulose nano-
crystals cause substantial local stress concentrations and failure at
reduced strain. In addition, although the partial surface silylation
Fig. 6. Isothermal crystallization exotherms of PLLA (a), PLLA–CNC-1 (b), and PLLA– of cellulose nanocrystals improved their degree of dispersion, it
SCNC-1 (c) under different crystallization temperatures (Tc). seems unlikely that the interfacial adhesion between PLLA and cel-
lulose nanocrystals was enhanced.
The results are also listed in Table 2. Generally, the t1/2 value in-
creased with increasing of Tc. The t1/2 values of PLLA/SCNC biocompos-
ites were lower than those of pure PLLA and PLLA/CNC biocomposites, 4. Conclusions
indicating rapid bulk crystallization rate by the addition of well-dis-
persed surface silylated cellulose nanocrystals in the PLLA matrix. The slow crystallization of PLLA was addressed by addition of a
biobased nucleating agent in the form of cellulose nanocrystals.
3.6. Mechanical properties The hypothesis of silane-treatment as a means to improve nucle-
ation efficiency was tested. Fine nanocrystals of 15 nm width and
The mechanical properties are summarized in Table 3. The 200–300 nm length were functionalized by partial silylation (SCNC
PLLA/CNC samples, which were actually microcomposites due to particles), in order to improve their dispersion in PLLA. Their fine dis-
A. Pei et al. / Composites Science and Technology 70 (2010) 815–821 821

persion in organic solvents was confirmed experimentally. The crys- [14] Gray DG. Transcrystallization of polypropylene at cellulose nanocrystal
surfaces. Cellulose 2008;15(2):297–301.
tallization rate was strongly increased by only 1% SCNC addition,
[15] Lima MMD, Borsali R. Rodlike cellulose microcrystals: structure, properties,
whereas CNC particles did not have as strong effects. Compared with and applications. Macromol Rapid Commun 2004;25(7):771–87.
PLLA/CNC, the PLLA/SCNC composites were truly nanostructured [16] Wu QJ, Henriksson M, Liu X, Berglund LA. A high strength nanocomposite
with a highly dispersed nanocrystal phase and the associated larger based on microcrystalline cellulose and polyurethane. Biomacromolecules
2007;8(12):3687–92.
specific surface area for crystallite nucleation. This observation is in [17] Mathew A, Oksman K, Sain M. Mechanical properties of biodegradable
support of the presented hypothesis of silane-functionalized cellu- composites from poly lactic acid (PLA) and microcrystalline cellulose (MCC).
lose nanocrystals. Furthermore, the increased degree of crystallinity J Appl Polym Sci 2005;97(5):2014–25.
[18] Mathew AP, Oksman K, Sain M. The effect of morphology and chemical
significantly improved tensile modulus and strength of the nano- characteristics of cellulose reinforcements on the crystallinity of polylactic
composite. The strain-to-failure decreased with SCNC and CNC addi- acid. J Appl Polym Sci 2006;101(1):300–10.
tion, and possibly the interfacial PLLA–nanocrystals adhesion could [19] Oksman K, Mathew AP, Bondeson D, Kvien I. Manufacturing process of
cellulose whiskers/polylactic acid nanocomposites. Comp Sci Tech
be improved in order to address this problem. Reducing the length of 2006;66(15):2776–84.
cellulose nanocrystals (i.e. to obtain nanospheres instead of rods) is [20] Heux L, Chauve G, Bonini C. Nonflocculating and chiral-nematic self-ordering
expected to increase the nucleation efficiency, although this may de- of cellulose microcrystals suspensions in nonpolar solvents. Langmuir
2000;16(21):8210–2.
crease the potential reinforcement effect associated with the high [21] Bondeson D, Oksman K. Dispersion and characteristics of surfactant modified
aspect ratio of nanocrystals. cellulose whiskers nanocomposites. Compos Interf 2007;14(7–9):617–30.
[22] Petersson L, Kvien I, Oksman K. Structure and thermal properties of poly(lactic
acid)/cellulose whiskers nanocomposite materials. Comp Sci Tech
Acknowledgements
2007;67(11–12):2535–44.
[23] Cranston ED, Gray DG. Morphological and optical characterization of
This work was performed at the Swedish Center for Biomimetic polyelectrolyte multilayers incorporating nanocrystalline cellulose.
Fiber Engineering (Biomime; http://www.biomime.org/) funded by Biomacromolecules 2006;7(9):2522–30.
[24] Gousse C, Chanzy H, Excoffier G, Soubeyrand L, Fleury E. Stable suspensions of
the Swedish Foundation for Strategic Research (SSF). For A. Pei and partially silylated cellulose whiskers dispersed in organic solvents. Polymer
L.A. Berglund, part of the funding was from Wallenberg Wood Sci- 2002;43(9):2645–51.
ence Center. [25] Tang Z, Black R, Curran J, Hunt J, Rhodes N, Williams D. Surface properties and
biocompatibility of solvent-cast poly[epsilon-caprolactone] films. Biomaterials
2004;25(19):4741–8.
References [26] Azizi Samir MAS, Alloin F, Sanchez J-Y, El Kissi N, Dufresne A. Preparation of
cellulose whiskers reinforced nanocomposites from an organic medium
[1] Graupner N, Herrmann AS, Mussig J. Natural and man-made cellulose fibre- suspension. Macromolecules 2004;37(4):1386–93.
reinforced poly(lactic acid) (PLA) composites: an overview about mechanical [27] Krikorian V, Pochan DJ. Unusual crystallization behavior of organoclay
characteristics and application areas. Compos Part A – Appl Sci Manuf reinforced poly(L-lactic acid) nanocomposites. Macromolecules
2009;40(6–7):810–21. 2004;37(17):6480–91.
[2] Shibata M, Oyamada S, Kobayashi S, Yaginuma D. Mechanical composites and [28] Nam JY, Ray SS, Okamoto M. Crystallization behavior and morphology of
biodegradability of green composites based on biodegradable polyesters and biodegradable polylactide/layered silicate nanocomposite. Macromolecules
lyocell fabric. J Appl Polym Sci 2004;92(6):3857–63. 2003;36(19):7126–31.
[3] Huda MS, Drzal LT, Misra M, Mohanty AK. Wood-fiber-reinforced poly(lactic [29] Gardella JA, Chen JS, Magill JH, Hercules DM. Surface spectroscopic studies of
acid) composites: evaluation of the physicomechanical and morphological polymer surfaces and interfaces – poly(tetramethyl-para-
properties. J Appl Polym Sci 2006;102(5):4856–69. silphenylenesiloxane). J Am Chem Soc 1983;105(14):4536–41.
[4] Huda MS, Drzal LT, Misra M, Mohanty AK, Williams K, Mielewski DF. A study [30] Araki J, Wada M, Kuga S. Steric stabilization of a cellulose microcrystal
on biocomposites from recycled newspaper fiber and poly(lactic acid). Ind Eng suspension by poly(ethylene glycol) grafting. Langmuir 2001;17(1):21–7.
Chem Res 2005;44(15):5593–601. [31] Schmidt S, Hillmyer M. Polylactide stereocomplex crystallites as nucleating
[5] Maurizio A, Gordana B-G, Aleksandra B, Maria Emanuela E, Gennaro G, Anita G. agents for isotactic polylactide. J Polym Sci, Part B: Polym Phys
Poly(lactic acid)-based biocomposites reinforced with kenaf fibers. J Appl 2001;39(3):300–13.
Polym Sci 2008;108(6):3542–51. [32] Gurato G, Gaidano D, Zannetti R. Influence of nucleating-agents on
[6] Iwatake A, Nogi M, Yano H. Cellulose nanofiber-reinforced polylactic acid. crystallization of 6-polyamide. Macromol Chem Phys 1978;179(1):231–45.
Comp Sci Technol 2008;68(9):2103–6. [33] Zhang L, Xiong C, Deng X. Miscibility, crystallization and morphology of
[7] Nakagaito AN, Fujimura A, Sakai T, Hama Y, Yano H. Production of poly(beta-hydroxybutyrate)/poly(D,L-lactide) blends. Polymer
microfibrillated cellulose (MFC)-reinforced polylactic acid (PLA) 1996;37(2):235–41.
nanocomposites from sheets obtained by a papermaking-like process. Comp [34] Lin Y, Zhang K-Y, Dong Z-M, Dong L-S, Li Y-S. Study of hydrogen-bonded blend
Sci Tech 2009;69(7–8):1293–7. of polylactide with biodegradable hyperbranched poly(ester amide).
[8] Suryanegara L, Nakagaito AN, Yano H. The effect of crystallization of PLA on the Macromolecules 2007;40(17):6257–67.
thermal and mechanical properties of microfibrillated cellulose-reinforced PLA [35] Tsuji H, Takai H, Saha SK. Isothermal and non-isothermal crystallization
composites. Comp Sci Tech 2009;69(7–8):1187–92. behavior of poly(L-lactic acid): effects of stereocomplex as nucleating agent.
[9] Drumright RE, Gruber PR, Henton DE. Polylactic acid technology. Adv Mater Polymer 2006;47(11):3826–37.
2000;12(23):1841–6. [36] Iannace S, Nicolais L. Isothermal crystallization and chain mobility of poly(L-
[10] Amash A, Zugenmaier P. Study on cellulose and xylan filled polypropylene lactide). J Appl Polym Sci 1997;64(5):911–9.
composites. Polym Bull 1998;40(2–3):251–8. [37] Fornes TD, Paul DR. Crystallization behavior of nylon 6 nanocomposites.
[11] Qiu W, Endo T, Hirotsu T. Interfacial interaction, morphology, and tensile Polymer 2003;44(14):3945–61.
properties of a composite of highly crystalline cellulose and maleated [38] Kolstad J. Crystallization kinetics of poly(L-lactide-co-meso-lactide). J Appl
polypropylene. J Appl Polym Sci 2006;102(4):3830–41. Polym Sci 1996;62(7):1079–91.
[12] Quillin D, Caulfield D, Koutsky J. Crystallinity in the polypropylene/cellulose [39] Liao R, Yang B, Yu W, Zhou C. Isothermal cold crystallization kinetics of
system. 1. Nucleation and crystalline morphology. J Appl Polym Sci polylactide/nucleating agents. J Appl Polym Sci 2007;104(1):310–7.
1993;50(7):1187–94. [40] Zhang J, Tsuji H, Noda I, Ozaki Y. Structural changes and crystallization
[13] Son S, Lee Y, Im S. Transcrystalline morphology and mechanical properties in dynamics of poly(L-lactide) during the cold-crystallization process
polypropylene composites containing cellulose treated with sodium hydroxide investigated by infrared and two-dimensional infrared correlation
and cellulase. J Mater Sci 2000;35(22):5767–78. spectroscopy. Macromolecules 2004;37(17):6433–9.

Vous aimerez peut-être aussi