Вы находитесь на странице: 1из 45

CHAPTER 8

Biodegradable
Polymer/Layered Silicate
Nanocomposites: A Review
Masami Okamoto
Advanced Polymeric Nanostructured Materials Lab,
Graduate School of Engineering, Toyota Technological Institute,
Hisakata 2-12-1, Tempaku, Nagoya 468 8511, Japan

CONTENTS
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. History of PLS Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . 3
3. Structure of LS and its Modification with Surfactants . . . . . . . . . 4
4. Preparation Methods and Structure of
PLS Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4.1. Intercalation of Polymers or
Prepolymers from Solution . . . . . . . . . . . . . . . . . . . . . . . . 5
4.2. The In Situ Intercalative Polymerization Method . . . . . . . . 5
4.3. The Melt Intercalation Method . . . . . . . . . . . . . . . . . . . . . 6
4.4. Structure of PLS Nanocomposites . . . . . . . . . . . . . . . . . . . 6
5. Characterization of PLS Nanocomposites . . . . . . . . . . . . . . . . . 7
6. Biodegradable Polymer/LS Nanocomposites . . . . . . . . . . . . . . . 8
6.1. PCL/LS Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . 9
6.2. PVA/LS Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . 10
6.3. PLA/LS Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . 10
6.4. PBS/LS Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . 13
6.5. PHB/LS Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . 14
6.6. Starch/LS Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . 15
6.7. Plant Oil/LS Nanocomposites . . . . . . . . . . . . . . . . . . . . . 15
6.8. Chitosan/LS Nanocomposites . . . . . . . . . . . . . . . . . . . . . 17
7. Materials Properties of Biodegradable Polymer/LS
Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
7.1. Mechanical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

Handbook of Biodegradable Polymeric


ISBN: 1-58883-053-5/$00.00 Materials and Their Applications
Copyright © 2005 by American Scientific Publishers Edited by Surya Mallapragada and Balaji Narasimhan
All rights of reproduction in any form reserved. Volume 1: Pages (1–45)

1
2 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

8. Crystallization of Biodegradable Polymer/LS


Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
9. Melt Rheology of Biodegradable Polymer/LS
Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
9.1. Dynamic Oscillatory Shear Measurement . . . . . . . . . . . . . 34
9.2. Steady Shear Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
9.3. Elongational Flow and Strain-Induced Hardening . . . . . . . 38
10. Processing Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
11. Creating Porous Ceramic Materials Via PLA/LS
Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
12. Current Status and Future Prospects of
Biodegradable Nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . 43
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

1. INTRODUCTION
Advanced technology in petrochemical-based polymers has brought many benefits to
mankind. However, it has become clear that nondegradable plastic materials used for dispos-
able applications are significantly disturbing and damaging the Earth’s ecosystem. The envi-
ronmental impact of persistent plastic wastes is of increasing global concern, and alternative
disposal methods are limited. Incineration of these plastic wastes always produces a large
amount of carbon dioxide that contributes to global warming; in some cases, toxic gases are
also produced, which contributes to global pollution. On the other hand, satisfactory landfill
sites are also limited. Another disadvantage of nondegradable plastic materials is that the
Earth has finite resources in terms of fossil origin fuel.
For these reasons, there is an urgent need to develop renewable, source-based, environ-
mentally benign polymeric materials (biopolymers), especially for use in short-term packag-
ing and disposable applications. Such materials would not involve the use of toxic or noxious
components in their manufacture, and could allow for composting into naturally occurring
degradation products.
The ideal biopolymer is of renewable biological origin and biodegradable at the end of
its life. Biopolymers include polysaccharides such as cellulose and starch; carbohydorate
polymers produced by bacteria and fungi [1]; and animal protein-based biopolymers such
as wool, silk, gelatin, and collagen. On the other hand, poly(vinyl alcohol) (PVA), poly(-
caprolactone) (PCL), and poly(butylene succinate) (PBS) are examples of polymers that
have synthetic origin but are biodegradable.
In today’s commercial venues, biopolymers have proven to be relatively expensive and
available only in small quantities. This has led to limited applications to date. However,
there are signs that this is changing, with increasing environmental awareness and more
stringent legislation regarding recyclability and restrictions on waste disposal. Cargill Dow
has a polylactide (PLA) in production (Natureworks™), and Metabolix has been developing
polyhydroxyalkanoate (PHA) (Biopol™).
Thus, the increasing appreciation of the various intrinsic properties of biopolymers, cou-
pled with the knowledge of how such properties can be improved to achieve compatibility
with thermoplastics processing, manufacturing, and end-use requirements, has fueled tech-
nological and commercial interest in biopolymers.
Of particular interest is a recently developed nanocomposite technology consisting of a
polymer and layered silicate. This combination often exhibits remarkably improved mechan-
ical and various other properties [2] when compared to pure polymers or conventional
composites (both micro- and macrocomposites). These improvements can include high
moduli [3], increased strength and heat resistance [4], decreased gas permeability [5] and
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 3

flammability [6], and increased biodegradability of biodegradable polymers [7]. On the other
hand, these materials have also proven to be unique model systems for the study of the
structure and dynamics of polymers in confined environments [8].
The main reason for these improved properties is interfacial interaction between matrices
and organically modified layered silicates (OMLSs), as opposed to conventional composites.
Layered silicates (LSs) have layer thickness in the order of 1 nm and very high aspect ratios
(e.g., 10–1,000). A small weight percentage of OMLSs that are properly dispersed throughout
the matrix thus creates a much larger surface area for polymer-filler interfacial interactions
than do conventional composites [9]. Although the intercalation chemistry of polymers (when
mixed with appropriately modified layered silicate and synthetic layered silicates [10, 11])
has been understood for a long time, the study of polymer/LS (PLS) nanocomposites has
recently gained greater momentum.
This review is intended to highlight the major developments in this area during the last
decade. The different techniques used to prepare biodegradable PLS nanocomposites, their
physicochemical characterization, the improved materials properties that those materials can
display, and the processing and probable applications of biodegradable PLS nanocomposites
will be reported in detail.

2. HISTORY OF PLS NANOCOMPOSITES


The first successful PLS nanocomposite appeared about ten years ago, through a pioneering
effort of a research team at Toyota Central Research & Development Co., Inc. (TCRD).
This material was a nylon 6/LS hybrid [12, 13].
Earlier attempts at preparing polymer/clay composites are described in patent literature
of the mid-1990s [14, 15]. In such cases, incorporation of 40–50 wt% LS (clay mineral)
(bentonite, hectorite, etc.) into a polymer was attempted but produced unsatisfactory results:
The maximal modulus enhancement was only around 200%, although the clay loading was
as much as 50 wt%. The poor results were due to failure to achieve good dispersion of clay
particles in the matrix, in which clay minerals existed as agglomerated tactoids. Such poor
dispersion of the clay particles could improve the material rigidity, but certainly sacrificed
the strength, the elongation at break, and the toughness of the material [14, 15].
A primary reason why it is impossible to improve tactoid dispersion into well-dispersed
exfoliated monolayers of clay is due to the intrinsic incompatibility of hydrophilic layered
silicates with hydrophobic engineering plastics. One attempt at circumventing this difficulty
was made by Unitika Ltd. [16] about thirty years ago; those researchers prepared nylon 6/clay
composites (not nanocomposites) by in situ polymerization of -caprolactam with montmo-
rillonite, but the results were disappointing.
The first major breakthrough in this area was in 1987, when Fukushima and Inagaki of
TCRD, in their detailed study of polymer/layered silicate composites, persuasively demon-
strated that lipophilization caused by replacing inorganic cations in galleries of native clay
with alkylammonium surfactants successfully made them compatible with hydrophobic poly-
mer matrices [17]. The modified clay was thus called lipophilized clay, organophilic clay, or
simply organoclay (OMLS). Furthermore, they found that the lipophilization enabled expan-
sion of the clay galleries and exfoliation of the silicate layers into single layers 1 nm thick.
In 1993, Usuki, Fukushima, and their colleagues at TCRD successfully prepared, for the
first time, an exfoliated nylon 6/LS hybrid (NCH) by in situ polymerization of -caprolactam,
in which alkylammonium-modified organoclay was thoroughly dispersed in advance [12, 13].
The resulting composite (when only 4.2 wt% clay was loaded) possessed a doubled modulus;
strength enhanced by 50%; and an 80 C increase in heat distortion compared to neat nylon 6,
as shown in Table 1. This produced a new generation of engineering materials, which we
call “polymer/LS nanocomposites.”
Thus, along the stream of development in PLS nanocomposite technologies, many studies
have been devoted to PLS nanocomposites with intrinsically excellent polymer properties
that should have an attractive potential for continuous expansion of application versatility.
4 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

Table 1. Mechanical and thermal properties of a nylon 6/LS hybrid.

Nylon 6 Neat
Properties Nanocomposite Nylon 6

Clay content (wt%) 4.2 0


Specific gravity 1.15 1.14
Tensile strength (Mpa) 107 69
Tensile modulus (Gpa) 2.1 1.1
Impact (kJ/m2 ) 2.8 2.3
HDT ( C at 1.8 Mpa) 147 65

3. STRUCTURE OF LS AND ITS MODIFICATION


WITH SURFACTANTS
The clays commonly used for the preparation of PLS nanocomposites are in the same general
family of 2:1 layered or phyllosilicates (see Table 2). Their crystal structure consists of lay-
ers composed of two silica tetrahedrals fused to an edge-shared octahedral sheet of either
aluminium or magnesium hydroxide. The layer thickness is ∼1 nm and the lateral dimen-
sions of these layers may vary from 30 nm to several microns and even larger, depending
on the particular layered silicate. Stacking of the layers leads to a regular van der Waal’s
gap between the layers, called the interlayer or gallery. Isomorphic substitution within the
layers (for example, Al+3 replaced by Mg+2 or by Fe+2 , or Mg+2 replaced by Li+1 ) generates
negative charges that are counterbalanced by alkali and alkaline earth cations situated inside
the galleries, as shown in Figure 1.
The most commonly used layered silicates are montmorillonite (MMT, Mx (Al4−x Mgx )·
Si8 O20 (OH)4 ) hectorite (Mx (Mg6−x Lix )Si8 O20 (OH)4 ), and saponite (Mx (Si8−x Alx )Si8 O20 ·
(OH)4 (x = 0.3–1.3). This type of clay is characterized by a moderate surface charge (cation
exchange capacity: CEC of 80–120 mequiv/100 gm) and layer morphology. These clays
are only miscible with hydrophilic polymers, such as poly(ethylene oxide) (PEO) [18] and
poly(vinyl alcohol) (PVA) [17]. To improve miscibility with other polymer matrices, one must
convert the normally hydrophilic silicate surface to organophilic, which enables intercalation
of many engineering polymers. Generally, this can be done by ion-exchange reactions with
cationic surfactants, including primary, secondary, tertiary, and quaternary alkylammonium
or alkylphosphonium cations. The role of alkylammonium or alkylphosphonium cations in
the organosilicates is to lower the surface energy of the inorganic host and improve the
wetting characteristics with the polymer matrix; this results in larger interlayer spacing. One
can demonstrate that about 100 alkylammonium salt molecules are localized near the indi-
vidual silicate layers (∼8 × 10−15 m2 ) and active surface area (∼800 m2 /g). Additionally,
the alkylammonium or alkylphosphonium cations could provide functional groups that can
react with the polymer matrix or, in some cases, initiate the polymerization of monomers to
improve the strength of the interface between the inorganic and the polymer matrix [10, 19].
Vaia and Giannelis [20] have shown that alkyl chains can vary from liquid-like to solid-like,
with the liquid-like structure dominating as the interlayer density or chain length decreases
(see Fig. 2), or as the temperature increases by using Fourier transform infrared spectroscopy
(FTIR). This is understandable because of the relatively small energy differences between
the trans and gauche conformers; the idealized models described earlier assume all trans
conformations. In addition, for the longer chain length surfactants, the surfactants in the
layered silicate can show thermal transition akin to melting or liquid-crystalline to liquid-like
transitions upon heating.
Table 2. Chemical formulas and characteristic parameters of commonly used 2:1 phyllosilicates.

2:1 Phyllosilicates Chemical Formulaa CEC (mequiv/100 gm) Particle Length (nm)

Montmorillonite Mx (Al4−x Mgx )Si8 O20 (OH)4 110 100–150


Hectorite Mx (Mg6−x Lix )Si8 O20 (OH)4 120 200–300
Saponite Mx Mg6 (Si8−x Alx )Si8 O20 (OH)4 86.6 50–60
a
M = monovalent cation; x = degree of isomorphous substitution (between 0.3 and 1.3).
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 5

Tetrahedral

Basal spacing
~1 nm
Octahedral

Tetrahedral

Exchangeable cations Al, Fe, Mg, Li

OH

Li, Na, Rb, Cs

Figure 1. Structure of 2:1 phyllosilicates.

4. PREPARATION METHODS AND STRUCTURE OF


PLS NANOCOMPOSITES
There is considerable literature available devoted to developing PLS nanocomposites with dif-
ferent combinations of OMLS and matrix polymers, employing somewhat different technolo-
gies appropriate to each. The technologies are broadly classified into three main categories.

4.1. Intercalation of Polymers or


Prepolymers from Solution
This technology is based on a solvent system in which polymers or prepolymers are soluble
and the silicate layers are swellable. The layered silicate is first swollen in a solvent such as
water, chloroform, or toluene, etc. When the polymer and layered silicate solutions are mixed,
the polymer chains intercalate and displace the solvent within the interlayer of the silicate.
Upon solvent removal, the intercalated structure remains, resulting in PLS nanocomposites.

4.2. The In Situ Intercalative Polymerization Method


In this method, the OMLS is swollen within the liquid monomer or a monomer solution
so the polymer formation can occur between the intercalated sheets. Polymerization can be
initiated either by heat or radiation, by the diffusion of a suitable initiator, or by an organic

(a) (b) (c)

Figure 2. Alkyl chain aggregation models: (a) short chain lengths, where the molecules are effectively isolated from
each other; (b) medium chain lengths, where quasi-discrete layers form with various degrees of in-plane disor-
der and interdigitation between the layers; and (c) long chain lengths, where interlayer order increases, leading
to a liquid-crystalline polymer environment. Open circles represent the CH2 segments while cationic head groups
are represented by filled circles. Adapted from Ref. 2, M. Okamoto, “Encyclopedia of Nanoscience and Nano-
technology,” with permission from American Scientific Publishers.
6 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

initiator or catalyst fixed through cation exchange inside the interlayer before the monomer
swelling step.

4.3. The Melt Intercalation Method


This method involves annealing, statically or under shear, a mixture of the polymer and
OMLS above the softening point of the polymer. This method has great advantages over
either in situ intercalative polymerization or polymer solution intercalation. First, this method
is environmentally benign due to the absence of organic solvents. Second, it is compatible
with current industrial processes such as extrusion and injection molding. The melt inter-
calation method allows the use of polymers which were previously not suitable for in situ
polymerization or the solution intercalation method.
Other possibilities are exfoliation-adsorption [21, 22] and template synthesis [23]. Nowa-
days, this solvent-free method is much preferred for practical industrial material production
due to its high efficiency and its possible avoidance of environmental hazards.

4.4. Structure of PLS Nanocomposites


As described by Kojima et al. [12] and in a 1950 U.S. patent [13], in situ polymerization was
employed for the first time in NCH production, and the melt intercalation was the direct
blending of OMLS into a modified polymer matrix such as used in polypropylene (PP)/LS
nanocomposites [24]. Since Giannelis and his colleagues [25] found that melt-compounding
of polymers with clay is possible without using organic solvents, nanocomposite preparations
using this method have been widely used, especially for polyolefin-based nanocomposites.
This process involves annealing, statically or under shear, a mixture of the polymer and
OMLS above the softening point of the polymer. During the annealing, the polymer chains
diffuse from the bulk polymer melt into the galleries between the silicate layers. Depending
on the degree of penetration of the matrix into the organically modified layered silicate gal-
leries, nanocomposites are obtained with structures ranging from intercalated to exfoliate.
Polymer penetration resulting in finite expansion of the silicate layers produces intercalated
nanocomposites consisting of well-ordered multilayers with alternating polymer/silicate lay-
ers and a repeat distance of few nanometers (intercalated, see Fig. 3) [24]. On the other

Al, Fe, Mg, Li


L OH
1 nm
O Tetrahedral
Li, Na, Ra, Cs
Octahedral
One Clay Platelet
L: 100 – 200 nm in case of MMT
Tetrahedral

Exchangeable cations
dclay
ξclay

Lclay

Form factors of dispersed clay The structure of 2:1 layered silicates

Intercalated Intercalated-and-flocculated Exfoliated

Figure 3. Schematic illustration of three different types of thermodynamically achievable polymer/clay nano-
composites.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 7

hand, extensive polymer penetration resulting in disordered and eventual delamination of


the silicate layers produces nearly exfoliated nanocomposites consisting of individual silicate
layers dispersed in the polymer matrix (exfoliated) [26]. Under some conditions, the inter-
calated nanocomposites exhibit flocculation because of the hydroxylated edge–edge interac-
tion of silicate layers (intercalated-and-flocculated). The length of the oriented collections
(in the range of 300–800 nm) is far larger than that of the original clay (mean diameter
∼150 nm) [27, 28]. Such flocculation presumably is governed by an interfacial energy between
the polymer matrix and OMLS and is controlled by ammonium cation-matrix polymer inter-
action. The polarity of the matrix polymer is of fundamental importance in controlling the
nanoscale structure.

5. CHARACTERIZATION OF PLS
NANOCOMPOSITES
The structure of PLS nanocomposites has typically been established using a wide-angle
x-ray diffraction (WAXD) analysis and transmission electron microscope (TEM) observa-
tions. Due to its availability and ease of use, WAXD is most commonly used to probe PLS
nanocomposite structures and, sometimes, to study the kinetics of the polymer melt interca-
lation. By monitoring the position, shape, and intensity of the basal reflections from the dis-
tributed silicate layers, the PLS nanocomposite structure (either intercalated or exfoliated)
may be identified. For example, in exfoliated nanocomposites, the extensive layer separation
associated with the delamination of the original silicate layers in the polymer matrix results
in the eventual disappearance of any coherent x-ray diffraction from the distributed silicate
layers. On the other hand, for intercalated nanocomposites, the finite layer expansion asso-
ciated with the polymer intercalation results in the appearance of a new basal reflection
corresponding to the larger gallery height.
WAXD offers a convenient method to determine the interlayer spacing of the silicate lay-
ers in the original layered silicates and in the intercalated nanocomposites (within 1–4 nm),
but little can be said about the spatial distribution of the silicate layers or any structural
inhomogeneities in the PLS nanocomposites. Additionally, some layered silicates initially
do not exhibit well defined basal reflection. Thus, peak broadening and intensity decreases
are very difficult to study systematically. Therefore, only tentative conclusions can be drawn
concerning the mechanism of nanocomposite formation and their structure based solely on
WAXD patterns.
On the other hand, TEM allows a qualitative understanding of the internal structure,
spatial distribution of the various phases, and defect structure of nanocomposites through
direct visualization. However, special care must be exercised to ensure that a representative
cross section of the sample is evaluated. The WAXD patterns and corresponding TEM
images of three different types of nanocomposites are presented in Figure 4.
The solid-state nuclear magnetic resonance (NMR) method of quantifying the level of
clay exfoliation is also a very important facet of nanocomposite characterization. The main
objective in solid-state NMR measurement is to connect the measured longitudinal relax-
ation (TH 1 s) of protons and
13
C nuclei with the quality of clay dispersion. The extent of
and the homogeneity of the dispersion of the silicate layers within the polymer matrix are
very important for determining physical properties. The surfaces of naturally occurring lay-
ered silicates such as MMT are mainly made of silica tetrahedrals, while the central plane
of the layers contains octahedrally coordinated Al3+ (see Fig. 1 and Table 2) with frequent
nonstoichiometric substitutions, where an Al3+ is replaced by Mg2+ and, somewhat less fre-
quently, by Fe3+ . Typical concentrations of Fe3+ (spin = 5/2) in naturally occurring clays
produce nearest neighbor Fe–Fe distances of about 1.0–1.4 nm [29]. At such distances, the
spin exchange interaction between the unpaired electrons on different Fe atoms is expected
to produce magnetic fluctuations in the vicinity of the Larmor frequencies for protons or 13 C
nuclei [29]. The spectral density of these fluctuations is important because the TH 1 of protons
and 13 C nuclei within ∼1 nm of the clay surface can be directly shortened. For protons,
if that mechanism is efficient, relaxation will also propagate into the bulk of the polymer
by spin diffusion. Thus, this paramagnetically induced relaxation will influence the overall
8 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

2000
Original OMLS
1500

1000

500

0
Intercalated
1500 Intercalated 200 nm

1000
Intensity /A.U.

500

0
Intercalated-and-flocculated
1500

1000
Intercalated-and-flocculated 200 nm
500

0
Exfoliated
1500

1000

500

0 Exfoliated 200 nm
2 4 6 8 10
2Θ/degrees

Figure 4. (a) WAXD patterns, and (b) TEM images of three different types of nanocomposites.

measured TH 1 to an extent that will depend both on the Fe concentration in the clay layer
and, more importantly, on the average distances between clay layers. The latter dependence
suggests a potential relationship between measured TH 1 values and the quality of the clay
dispersion. If the clay particles are stacked and poorly dispersed in the polymer matrix, the
average distances between polymer-clay interfaces are greater, and the average paramagnetic
contribution to TH1 is weaker. VanderHart et al. [30] also employed the same arguments to
understand the stability of a particular OMLS under different processing conditions.

6. BIODEGRADABLE POLYMER/
LS NANOCOMPOSITES
Recently, some groups have undertaken the preparation and characterization of the materi-
als properties of various kinds of biodegradable polymer/LS nanocomposites having proper-
ties suitable for a wide range of applications. To date, reported biodegradable polymers for
the preparation of nanocomposites are:

• poly(-caprolactone) (PCL) [31–35]


• poly(vinyl alcohol) (PVA)[36, 37]
• poly(lactide) (PLA) [38–48],
• poly(butylene succinate) (PBS) [49–51],
• unsaturated polyester [52],
• poly(hydroxy butyrate) (PHB) [53, 54],
• aliphatic polyester [55–58],
• thermoplastic starch [59, 60]
• other renewable resources [61, 62]
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 9

6.1. PCL/LS Nanocomposites


For the preparation of poly(-caprolactone) (PCL)-based nanocomposites, Messersmith
and Giannelis [31] modified MMT with protonated aminolauric acid and dispersed the
modified MMT in liquid -caprolactone (CL) before polymerizing at high temperature.
The nanocomposites were prepared by mixing up to 30 wt% of the modified MMT with dried
and freshly distilled CL for a couple of hours, followed by ring-opening polymerization under
stirring at 170 C for 48 hours. The same researchers [32] have also reported on -caprolactone
polymerization inside a Cr+3 -exchanged fluorohectorite at 100 C for 48 hours.
Pantoustier and his colleagues [33] used this in situ intercalative polymerization method.
They used both pristine MMT and -amino dodecanoic acid-modified MMT for the com-
parison of prepared nanocomposite properties. For nanocomposite synthesis in a polymer-
ization tube, the desired amount of pristine MMT was first dried under vacuum at 70 C
for three hours. A given amount of CL was then added under nitrogen and the reaction
medium was stirred at room temperature for one hour. A solution of initiator (Sn(Oct)2 ) or
Bu2 Sn(Ome)2 ) in dry toluene was added to the mixture in order to reach a [monomer]/[Sn]
molar ratio equal to 300. The polymerization was then allowed to proceed for 24 hours
at room temperature. After polymerization, a reverse ion-exchange reaction was used to
isolate the PCL chains from the inorganic fraction of the nanocomposite. A colloidal sus-
pension was obtained by stirring 2 g of the nanocomposite in 30 mL of THF for two hours
at room temperature. Separately, a solution of 1 wt% of LiCl in THF was prepared. The
nanocomposite suspension was added to 50 mL of the LiCl solution and left to stir at room
temperature for 48 hours. The resulting solution was centrifuged at 3,000 rpm for 30 minutes.
The supernatant was then decanted and the remaining solid was washed by dispersing in
30 mL of THF, followed by centrifugation. The combined supernatant was concentrated and
precipitated from petroleum ether.
The polymerization of CL with pristine MMT gives PCL with a molar mass of 4,800 g/mol
and a narrow distribution. For comparison, the researchers also conducted the same exper-
iment without MMT, but there was no polymerization of CL. These results demonstrate
the ability of MMT to catalyze and control CL polymerization, at least in terms of molecu-
lar weight distribution that remains remarkably narrow. For the mechanism of polymeriza-
tion, the researchers assumed that the CL is activated through the interaction with acidic
site on the clay surface and the polymerization is likely to be proceeding via the activated
monomer mechanism by the cooperative function of Lewis acidic aluminum and Bronsted
acidic silanol functionalities on the initiator walls (see Fig. 5). On the other hand, in the poly-
merization of CL with the protonated -amino dodecanoic acid-modified MMT, the molar
mass was 7,800 g/mol with a monomer conversion of 92% and, again, a narrow molecular
weight distribution. The WAXD patterns of both nanocomposites indicate the formation
of intercalated structure. In another very recent publication [34, 35], same group prepared

100 nm

Figure 5. TEM image of a PCL nanocomposite containing 3 wt% MMT-Alk. Reprinted with permission from [66],
B. Lepoittevin et al., Polymer 43, 4017 (2002). © 2002, Elsevier Science Ltd.
10 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

PCL/MMT nanocomposites by using in situ ring-opening polymerization of CL using dibutyl


tin dimethoxide as an initiator/catalyst.

6.2. PVA/LS Nanocomposites


More recently, Strawhecker and Manias [36] used this method in attempts to produce
poly(vinyl alcohol) (PVA)/MMT nanocomposite films. PVA/MMT nanocomposite films were
cast from MMT/water suspensions where PVA was dissolved. Room temperature distilled
water was used to form a suspension of Na+ -MMT. The suspension was first stirred for
one hour and then sonicated for 30 minutes. Low viscosity, fully hydrolyzed atactic PVA was
then added to the stirring suspensions such that the total solid (silicate plus polymer) was
≤5 wt%. The mixtures were then heated to 90 C to dissolve the PVA, again sonicated for
30 minutes, and, finally, films were cast in a closed oven at 40 C for two days. The recov-
ered cast films were then characterized by both WAXD and TEM. Both the d-spacing and
their distribution decreased systematically with increasing MMT wt% in the nanocomposites.
The TEM photograph of 20 wt% clay containing nanocomposites reveals the coexistence of
silicate layers in the intercalated and the exfoliated states.

6.3. PLA/LS Nanocomposites


Okamoto and his colleagues [38, 40] first reported the preparation of intercalated PLA/LS
nanocomposites. For nanocomposite (PLACN) preparation, C18 -MMT and PLA were first
dry-mixed by shaking them in a bag. The mixture was then melt-extruded by using a twin-
screw extruder operated at 190 C to yield very light gray colored strands of PLACNs.
Nanocomposites loaded with very small amounts of oligo-PCL (Mw = 2,000 g/mol) as a
compatibilizer were also prepared in order to understand the effect of oligo-PCL on the
morphology and properties of PLACNs [38]. The compositions of various nanocomposites
of PLA with C18 -MMT are summarized in Table 3. WAXD patterns of a series of nanocom-
posites are shown in Figure 6. Figure 7 shows TEM photographs of nanocomposites corre-
sponding to the WAXD patterns. On the basis of WAXD analyses and TEM observations,
they calculated form factors (see Table 4) [i.e., average length (Lclay ), thickness (dclay )] of
the stacked intercalated silicate layers, and the correlation length (clay ) between them (see
Fig. 3). These data clearly established that silicate layers of the clay were intercalated and
randomly distributed in the PLA matrix. Incorporation of very small amounts of oligo-PCL
as a compatibilizer in the nanocomposites led to a better parallel stacking of the silicate
layers and also much stronger flocculation due to the hydroxylated edge–edge interaction of
the silicate layers. Due to the interaction between clay platelets and the PL-matrix in pres-
ence of very small amounts of oligo-PCL, the disk–disk interaction plays an important role

Table 3. Composition and characteristic parameters of various PLACNs based on PLA, oligo-PCL, and
C18 -MMT.

Composition (wt%)

Sample PLA oligo-PCL C18 -MMTb Mw × 10−3 (g/mol) Mw /Mn Tg ( C) Tm ( C)


cc (%)

PLACN1 97 3 [2.0] 178 1.81 60.0 169 50.65


PLACN2 95 5 [3.0] 185 1.86 60.0 170 39.01
PLACN3 93 7 [4.8] 177 1.69 59.8 170 43.66
PLACN4 94.8 0.2 5 [3.3] 181 1.76 58.6 170 41.47
PLACN5 94.5 0.5 5 [3.3] 181 1.76 57.6 169 32.91
PLACN6 93 2.0 5 [2.8] 180 1.76 54.0 168 —
PLACN7 92 3.0 5 [2.4] 181 1.77 51.0 168 —
PLAa 100 187 1.76 60.0 168 36.24
PLA1 99.8 0.2 180 1.76 58.0 168.5 46.21
PLA2 99.5 0.5 180 1.76 57.0 168.8 52.51
PLA3 98 2.0 180 1.76 54.7 169 —
a
Mw and PDI of extruded PLA (at 190 C) are 180 × 103 (g/mol) and 1.6, respectively. b Value in the parentheses
indicates the amount of clay (inorganic part) content after burning. c The degree of crystallinity.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 11

(a)
2000 (001) C18-mmt
1500
(b)
1000 2000 PLACN2
500 1500
0 *
PLACN1 1000
2000
500
1500
0
Intensity/ A.U.

1000
PLACN4

Intensity /A.U.
* 2000
500
0 1500
PLACN2 *
2000 1000
1500 500
*
1000 0
500 PLACN5
2000
0
PLACN3 1500
2000 *
1000
1500
1000 * 500
500 0
0 2 4 6 8 10
0 2Θ/degrees
0 2 4 6 8 10
2Θ / degrees

Figure 6. WAXD patterns for C18 -MMT and various PLACNs: (a) without oligo-PCL, and (b) with oligo-PCL.
The dashed line in each figure indicates the location of the silicate (001) reflection of C18 -MMT. The asterisks
indicate the (001) peak for C18 -MMT dispersed in PLA matrices. Reprinted with permission from [28], S. Sinha Ray
et al., Macromolecules 35, 3104 (2002). © 2002, American Chemical Society.

in determining the stability of the clay particles and, hence, the enhancement of mechanical
properties of such nanocomposites.
In their further research [41, 42, 48], this group prepared a series of PLACNs with various
types of organoclay in order to investigate the effect of organoclay on the morphology, prop-
erties, and biodegradability of PLACNs. Four different types of pristine layered silicates were
used and each of them was modified with a different type of surfactant. Detailed specifications

(a) (b)

2 µm 2 µm

(c) (d)

500 nm 500 nm

Figure 7. TEM bright field images: (a) PLACN2 (×10,000); (b) PLACN4 (×10,000); (c) PLACN2 (×40,0000); and
(d) PLACN4 (×40,000). The dark entities are cross sections of intercalated organoclay, and the bright areas are
the matrices. Reprinted with permission from [28], S. Sinha Ray et al., Macromolecules 35, 3104 (2002). © 2002,
American Chemical Society.
12 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

Table 4. Comparison of form factors between PLACN2 and PLACN4


obtained from WAXD patterns and TEM observations.

Form factors PLACN2 PLACN4

WAXD
d001 (nm) 3.03 2.98
dclay (nm) 13 10
TEM
dclay (nm) 38 ± 17.25 30 ± 12.5
Lclay (nm) 448 ± 200 659 ± 145
Lclay /dclay 12 22
clay (nm) 255 ± 137 206 ± 92

of the various types of organoclay they used are presented in Table 5. On the basis of
WAXD analyses and TEM observations, the researchers successfully formed four different
types of PLACNs. Ordered intercalated-and-flocculated nanocomposites were obtained when
ODA was used as the organoclay; disordered intercalated structures resulted in the case of
PLA/SBE4 nanocomposites; PLA/SAP4 nanocomposites showed near to exfoliate nanocom-
posites; and the coexistence of stacked intercalated and exfoliated nanocomposite structures
was evident with PLA/MEE4 nanocomposites. Thus, the nature of OMLS has a strong effect
on the final morphology of PLA-based nanocomposites.
In a very recent work, Okamoto and Maiti [43] prepared a series of PLACNs with three
different types of pristine layered silicate such as saponite, MMT, and mica, and each of
them was modified with alkylphosphonium salts having various chain lengths. Their first
goal was to determine the effect of alkylphosphonium modifiers of different chain lengths
on the properties of organoclay and how the different clays behave differently having same
organic modifier. Second, they wanted to determine the effects of dispersion, intercalation,
and aspect ratio of clay on materials properties. From the resulting WAXD patterns, it was
clearly observed that the d-spacing (001) increases with increasing modifier chain length
and, for a fixed modifier, it increases with increasing lateral dimensions of the clay particles.
These researchers concluded that there are two reasons for this type of behavior: the CEC
value, and the lateral size of various pristine layered silicates. In both cases, layered silicates
followed the order mica > MMT > saponite. The CEC factor is more important than the
lateral size of the silicates to control the d-spacing/stacking of silicate layers. Since mica has
a large lateral size and also a high amount of surfactant molecules due to its high CEC value,
surfactant chains inside the integrally have restricted conformation due to physical jamming.
These researchers believe there is less physical jamming in saponite due to its lower CEC
and smaller lateral size. The results for OMLS, based on TEM and WAXD analyses, are
schematically illustrated in Figure 8.
Figure 9 compares the WAXD patterns of nanocomposites with different clay dimensions
having the same clay (n-hexadecyl tri-n-butyl phosphonium bromide (C16 )-modified) content
(3 wt%). For MMT-based nanocomposites, the peaks are sharp and the crystallite sizes are
slightly smaller than those of the corresponding organoclay, indicating an almost ordered

Table 5. Specifications and designations of OMLS used for the preparation of PLACNs.

Particle length CEC Organic salt used for


Clay Codes Pristine Clay (nm) mequiv/100 gm the modification of clay Suppliers

ODA MMT 150–200 110 Octadecyl Nanocor Inc., USA


ammonium cation
SBE MMT 100–130 90 Trimethyloctadecyl Hojun yoko Co.,
ammonium cation Japan
MEE Synthetic 200–300 120 Dipolyoxyethylene CO-OP Chemicals,
F-mica alkyl(coco) methyl Japan
ammonium cation
SAP Saponite 50–60 86.6 Tributylhexadecyl CO-OP Chemicals,
phosphonium cation Japan
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 13

200 nm
70 nm
50 nm

1.87 nm

2.44 nm
2.13 nm
Smectite Montmorillonite Mica

Figure 8. Schematic representation of organoclays (OMLSs) with C16 ions. Reprinted with permission from [43],
P. Maiti et al., Chem. Mater. 14, 4654 (2002). © 2002, American Chemical Society.

structure of MMT in the nanocomposites. The peaks of the nanocomposites prepared with
mica clay are very sharp, similar to those of corresponding organoclay, and the slightly larger
crystallite sizes indicate that the number of stacked silicate layers is the same as that of the
original organoclay. However, some amount of PLA is intercalated between the galleries,
giving rise to a larger crystallite size. On the basis of WAXD patterns and crystallite size,
stacking of silicate layers in the organoclays and in various nanocomposites prepared with
three different organoclays is presented schematically in Figure 10.
More recently, Dubois et al. [44, 46] reported the preparation of plasticized PLA/MMT
nanocomposites. The OMLS they used was MMT modified with bis-(2-hydroxyethyl)methyl
(hydrogenated tallowalkyl) ammonium cations. WAXD analyses have confirmed the forma-
tion of intercalated nanocomposites (see Fig. 11).

6.4. PBS/LS Nanocomposites


Poly(butylene succinate) (PBS) is also an aliphatic thermoplastic polyester with many inter-
esting properties, including biodegradability, melt processability, and thermal and chemical
resistance. Although the above properties suggest potential applications of PBS, some of
its other properties such as softness, gas barrier properties, flexural properties, etc. are fre-
quently not adequate for a wide-range of applications.
K. Okamoto and M. Okamoto [49, 50] first reported the preparation of PBS /MMT
nanocomposites (PBSCNs) by simple melt extrusion of PBS and OMLS, having proper-
ties suitable for many applications. MMT modified with octadecylammonium chloride was
used as organoclay for nanocomposite preparation. In recent publications [50, 51], the same
researchers also reported the details of their studies on structure-property relationships
involving PBSCNs. Figure 12 represents the WAXD patterns of various PBSCNs. A TEM
image of the representative PBSCN is shown in Figure 13.

2000
C16

1500
Intensity /a.u

Mica

1000

MMT

500
Smectite

0
2 4 6 8 10
2Θ/ deg

Figure 9. WAXD patterns of smectite (SAP), MMT, and mica nanocomposites with C16 organoclay (OMLS) and
some clay content (3 wt%). Reprinted with permission from [43], P. Maiti et al., Chem. Mater. 14, 4654 (2002).
© 2002, American Chemical Society.
14 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

Smectite

MMT

Mica

organoclays nanocomposites

Figure 10. Schematic presentation of silicate layers in organoclay (OMLS) and in various nanocomposites.
Reprinted with permission from [43], P. Maiti et al., Chem. Mater. 14, 4654 (2002). © 2002 American Chemical
Society.

In 2002, Lee and his colleagues [55] reported the preparation of biodegradable aliphatic
polyester (APES)/organoclay nanocomposites using a melt intercalation method. Two kinds
of organoclays, Cloisite 30B and Cloisite 10A with different ammonium cations located in
the silicate galleries, were chosen for the nanocomposite preparation. The WAXD analyses
and TEM observations showed a higher degree of intercalation in the case of APES/Cloisite
30B nanocomposites as compared to that of APES/Cloisite 10A nanocomposites. According
to the researchers, this behavior may be due more hydrogen bonding interaction between
APES and the hydroxyl group in the galleries of Cloisite 30B nanocomposites than in the
APES/Cloisite 10A nanocomposites.

6.5. PHB/LS Nanocomposites


Poly(hydroxy butyrate) (PHB)/LS nanocomposites were successfully prepared by Maiti and
his colleagues [54]. PHB forms well-ordered intercalated nanocomposites with OMLSs.
Although the dispersion of clay particles was not so good (see Fig. 14), the nanocomposites
exhibited a higher storage modulus (an increase up to 40% when compared to pure PHB).
Better biodegradation behavior was also observed for these nanocomposites as compared to
the neat PHB.

(a) (b)

100 nm 50 nm

Figure 11. TEM images of a fully exfoliated Cloisite 30B-based nanocomposite, showing (a) fine distribution of the
clay in the matrix, and (b) delamination of silicate layers. Reprinted with permission from [46], M.-A. Paul et al.,
Macromol. Rapid Commun. 24, 561 (2003). © 2003, Wiley-VCH Verlag GmbH & Co.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 15

(a) 2000
(001) C18-mmt
1500 (b)
1000 4000
(001) qC16-sap
500 3000

0 2000

1500 PBSCN1 1000

1000 0
800 PBSCN5
500
600
0
400
Intensity/A.U

PBSCN2

Intensity/A.U
1500 200
1000 0
PBSCN6
500 800
600
0
PBSCN3 400
1500
200
1000 0
800 PBS
500
600
0
PBSCN4 400
1500
200
1000 0
2 4 6 8 10
500 2Θ/ degrees
0
2 4 6 8 10
2Θ / degrees

Figure 12. (a) WAXD patterns for pure C18 -MMT powder and corresponding PBSCNs. The dashed line indicates
the location of the silicate (001) reflection of C18 -MMT. The asterisks indicate the (001) peak for C18 -MMT dis-
persed in the PBS matrix. (b) WAXD patterns for pure qC16 -MMT powder and corresponding PBSCNs. The dashed
line indicates the location of the silicate (001) reflection of qC16 -SAP. The asterisks indicate the (001) peak for
qC16 -MMT dispersed in the PBS matrix. Reprinted with permission from [50], S. Sinha Ray et al., Macromolecules
36, 2355 (2003). © 2003, American Chemical Society.

6.6. Starch/LS Nanocomposites


Park and his colleagues [59, 60] reported on their efforts to develop environmentally friendly
polymer hybrids. Biodegradable thermoplastic starch/LS nanocomposites were prepared by
the melt intercalation method. Natural montmorillonite (Na+ MMT; Cloisite Na+ ) and one
organically modified MMT with methyl tallow bis-2-hydroxyethyl ammonium cations located
in the silicategallery (Cloisite 30B) were chosen for the nanocomposite preparation. Starch
was prepared from natural potato starch by gelatinizing and plasticizing it with water and
glycerol. The dispersion of the silicate layers in the sarch hybrids was characterized by
using WAXD and TEM. They observed that the starch/Cloisite Na+ nanocomposites showed
higher tensile strength and thermal stability, and better barrier properties to water vapor as
compared to starch /Cloisite 30B nanocomposites and the pristine starch, due to the for-
mation of the intercalated nanostructure (see Fig. 15). The effect of clay contents on the
tensile, dynamic mechanical, and thermal properties as well as on the barrier properties of
the nanocomposites were investigated.

6.7. Plant Oil/LS Nanocomposites


Uyama et al. [61] investigated “green” nanocomposites prepared by an acid-catalyzed curing
of epoxidized plant oils in the presence of OMLS. Nanocomposites with a homogeneous
structure of organic and inorganic components were obtained, in which the clay was
16 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

(a) (b) (e) (f)

200 nm 100 nm 200 nm 100 nm


PBSCN1 PBSCN4

(c) (d) (g) (h)

200 nm 200 nm 200 nm 100 nm


PBSCN3 PBSCN6

Figure 13. TEM bright field images of PBSCNs: (a) PBSCN1 (×100,000), (b) PBSCN1 (×200,000), (c) PBSCN3
(×40,000), (d) PBSCN3 (×100,000), (e) PBSCN4 (×100,000), (f) PBSCN4 (×200,000), (g) PBSCN6 (×100,000),
and (h) PBSCN6 (×200,000) in which the dark entities are the cross sections of the intercalated or exfoliated
silicate layers. Reprinted with permission from [50], S. Sinha Ray et al., Macromolecules 36, 2355 (2003). © 2003,
American Chemical Society.

intercalated and randomly distributed in the polymer matrix (see Fig. 16). The reinforce-
ment effect of the addition of the clay was confirmed by dynamic viscoelasticity analysis.
Furthermore, the nanocomposites exhibited flexible properties. These researchers also found
good biodegradability of the cured polymer from epoxidized soybean oil. These nanocom-
posites are anticipated to become a new class of coating materials derived from inexpensive
renewable resources, which will contribute to global sustainability.

500 nm

Figure 14. TEM bright field images of PHB nanocomposites. Reprinted with permission from [54], P. Maiti et al.,
Polm. Mater. Sci. Eng. 88, 58 (2003). © 2003, Polymeric Materials Science & Engineering Division of the American
Chemical Society.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 17

(a ) (b)

(c) (d)

Figure 15. TEM images of starch/clay nanocomposites of different types of OMLS. (a) starch 95/Cloisite Na+ ,
(b) starch 95/Cloisit e6A, (c) starch 95/Cloisite10A, (d) starch 95/Cloisite 30B. Reprinted with permission from [59],
H. M. Park et al., Macromol. Mater. Eng. 287, 8, 553 (2002). © 2002, Wiley-VCH Verlag GmbH & Co.

6.8. Chitosan/LS Nanocomposites


Ruiz-Hitzky and his colleagues [62] reported on the intercalation of the cationic biopolymer
chitosan in Na+ -MMT, providing compact and robust three-dimensional nanocomposites
with interesting functional properties (see Fig. 17).
These researchers used CHN chemical analysis, x-ray diffraction, Fourier transform
infrared spectroscopy, scanning transmission electron microscopy, energy-dispersion x-ray
analysis, and thermal analysis to characterize the nanocomposites, confirming the adsorption
in mono- or bilayers of chitosan chains (depending on the relative amount of chitosan with
respect to the cationic exchange capacity of the clay). The first chitosan layer is adsorbed
through a cationic exchange procedure, while the second layer is adsorbed in the acetate salt
form. Because the deintercalation of the biopolymer is very difficult, the -NH+ −
3 Ac species
18 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

(A)

2 µm 200 nm

(B)

2 µm 200 nm

Figure 16. TEM images of ESO (epoxidized plant oils)-clay nanocomposites with clay content of (a) 5%, and
(b) 15%. Reprinted with permission from [61], H. Uyama et al., Chem. Mater. 15, 2492 (2003). © 2003, American
Chemical Society.

belonging to the chitosan second layer act as anionic exchange sites and, in this way, such
nanocomposites become suitable systems for the detection of anions. These materials have
been successfully used in the development of bulk-modified electrodes exhibiting numerous
advantages such as easy surface renewal, ruggedness, and long-term stability. The resulting
sensors were applied in the potentiometric determination of several anions, showing a higher
selectivity toward monovalent anions. This selectivity behavior could be explained by the
special arrangement of the polymer as a nanostructured bidimensional system.
The interlayer space in the nanocomposites prepared from chitosan-clay ratios of 0.25:1
and 0.5:1 can be related to the thickness of one chitosan sheet and, thus, to its intercalation
as a monolayer covering the interlayer surface of the clay, as shown in Figure 18. Above
such chitosan-clay ratios, the increase of the basal spacing can be explained as the uptake
of two chitosan layers by the clay.

7. MATERIALS PROPERTIES OF BIODEGRADABLE


POLYMER/LS NANOCOMPOSITES
PLS nanocomposites consisting of a polymer and clay (modified or not) frequently exhibit
remarkably improved mechanical and materials properties as compared to those of pris-
tine polymers containing small amounts (≤5 wt%) of layered silicate. Improvements can

CH3
OH C–O OH
NH3+
O NH
HO O HO O O
O O HO O O HO O
+
NH3 NH3+
OH OH

Figure 17. Chitosan chemical structure.


Biodegradable Polymer/Layered Silicate Nanocomposites A Review 19

Figure 18. Intercalation of chitosan into Na+ -MMT. Reprinted with permission from [61], M. Darder et al., Chem.
Mater. 15, 3774 (2003). © 2003, American Chemical Society.

include higher moduli, increased strength and heat resistance, decreased gas permeability
and flammability, and increased biodegradability of biodegradable polymers.

7.1. Mechanical Properties


7.1.1. Dynamic Mechanical Analysis (DMA)
DMA results are expressed by three main parameters: (1) the storage modulus (G
) corre-
sponding to the elastic response to the deformation; (2) the loss modulus (G

), correspond-
ing to the plastic response to the deformation, and (2) tan (i.e., the (G

/G
) ratio), useful
for determining the occurrence of molecular mobility transitions such as the glass transi-
tion temperature (Tg ). DMA analysis has been studied to track the temperature dependence
of G
, G

, and tan of pure PLA upon nanocomposite formation with five different types
of OMLS. Figure 19 shows the temperature dependence of G
, G

, and tan of pure PLA


20 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

(a) (b)
109

G′′/Pa
108 ω = 6.28 rad.s–1 ω = 6.28 rad.s–1
Strain = 0.05% Strain = 0.05%
Heating rate = 2 °C/min Heating rate = 2 °C/min
107

108
G′′/Pa

107
PLA PLA (e)
PLA/C18-MMT PLA/qC218-MMT 109
6
10

G′/Pa
10–1 108 ω = 6.28 rad.s–1
Strain = 0.05%
tan δ

Heating rate = 2 °C/min


10–2 107
108

G′′/Pa
10–3
0 40 80 120 160 0 40 80 120 160 107
Temperature/ °C Temperature / °C
106 PLA
PLA/qC13(OH)-Mica
(c) (d) 105
109
10–1
G′/Pa

tan δ
–1 –1
10 8 ω = 6.28 rad.s ω = 6.28 rad.s
Strain = 0.05% Strain = 0.05% 10–2
Heating rate = 2 °C/min Heating rate = 2 °C/min
107
10–3
108 0 40 80 120 160
G′′/Pa

Temperature /°C
107 PLA
PLA
PLA/qC18-MMT PLA/qC16-SAP
106

10–1
tan δ

10–2

10–3
0 40 80 120 160 0 40 80 120 160
Temperature/ °C Temperature / °C

Figure 19. Temperature dependence of storage modulus (G


), loss modulus (G

), and tan of pure PLA and


corresponding nanocomposites: (a) PLA/C18 -MMT4; (b) PLA/qC2 -MMT4; (c) PLA/qC18 -MMT4; (d) PLA/qC16 -
SAP4; (e) PLA/qC13 (OH)-mica 4. Reprinted with permission from [63], S. Sinha Ray et al., Macromol. Rapid
Commun. 24, 815 (2003). © 2003, Wiley-VCH Verlag GmbH & Co.

and various PLACNs. For all nanocomposites, significant enhancement of G


can be seen
in the investigated temperature range, indicating that all OMLS have strong influence on
the elastic properties of pure PLA. Below Tg , the enhancement of G
is also clear for all
PLACNs [63]. In the temperature range of −20—0 C, the increments in G
are 37% for
PLA/C18 -MMT, 52% for PLA/qC218 -MMT, 45% for PLA/qC18 -MMT, 31% for PLA/qC16 -SAP
(saponite), and 23% for PLA/qC13 (OH)-mica nanocomposites compared to that of pure
PLA [63]. In the temperature range of 80–90 C, all nanocomposites exhibit much greater
enhancement in G
(103% for PLA/C18 -MMT, 105% for PLA/qC218 -MMT, 96% for PLA/qC18 -
MMT, and 111% for PLA/qC13 (OH)-mica) than that of pure PLA, with the exception of
PLA/qC16 -SAP nanocomposite with a 45% increment. This is due to the mechanistic rein-
forcement by clay particles at high temperature. Above Tg , when materials become soft,
the reinforcement effect of clay particles becomes prominent due to the restricted move-
ment of the polymer chains and, hence, strong enhancement of modulus appeared in this
study [32]. The restriction of polymer chain movement by the clay particles is high in the
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 21

case of PLA/qC13 (OH)-mica because of the low value of clay (see Table 4 and Figure 3).
For this reason, PLA/qC13 (OH)-mica nanocomposites show high increments in G
at high
temperature ranges compared to that of other PLACNs. Furthermore, at room tempera-
ture (25 C), PLA/C18 -MMT, PLA/qC18 -MMT, PLA/qC218 -MMT, and PLA/qC16 -SAP showed
higher increments in G
of 38%, 44%, 51%, and 30%, respectively, than that of pure PLA,
while that of PLA/qC13 (OH)-mica showed only 26% higher. These increments come from
the extended intercalation, the higher degree of crystallization, and also the high aspect ratio
of dispersed clay particles in MMT-based nanocomposite systems.
On the other hand, above Tg the enhancement of G

is significant in all nanocomposites


in comparison with that of below Tg , indicating that plastic response to the deformation of
pure PLA is prominent in the presence of OMLS when materials become soft. However, the
presence of OMLS does not lead to a significant shift and broadening of the tan curves
for all PLACNs compared to that of pure PLA. This behavior has been ascribed to the
restricted segmental motions in the organic-inorganic interface neighborhood of intercalated
PLACNs.
The increment in G
directly depends upon the aspect ratio of dispersed clay particles,
which is also clearly observed in PBSCNs. The temperature dependence of G
of PBS and var-
ious PBSCNs are shown in Figure 20). The nature of enhancement of G
in PBSCNs with tem-
perature is somewhat different from well established intercalated polypropylene/LS nanocom-
posites [24] and well known exfoliated nylon 6/LS nanocomposite systems (N6CNs) [12, 64].
In the latter system, there is a maximum of 40–50% increment of G
as compared to that
of matrices at well below Tg ; above Tg , there is a strong enhancement (>200%) in G
. This
behavior is common for nanocomposites evaluated to date, and the reason is the strong rein-
forcement effect of the clay particles above Tg when materials become soft. But with PBSCNs,

PBS
PBSCN1
PBSCN2
PBSCN3
109
G′/Pa

PBSCN4

108

ω = 6.28 rad/s
8
Strain = 0.05%
10 Heating rate = 2°/min
G′′/Pa

107

106

10–1
tan δ

10–2

–40 –20 0 20 40 60 80 100


Temperature / °C

Figure 20. Temperature dependence of G


, G

and tan for neat PBS and various PBSCNs prepared with
C18 -MMT. Reprinted with permission from [50], S. Sinha Ray et al., Macromolecules 36, 2355 (2003). © 2003,
American Chemical Society.
22 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

the order of enhancement of G


is almost the same below and above Tg ; this behavior may be
due to the extremely low Tg (−29 C) of the PBS matrix. At the temperature range of −50 C
to −10 C, the increments in G
are 18% for PBSCN1, 31% for PBSCN2, 67% for PBSCN3,
and 167% for PBSCN4 compared to that of neat PBS. Furthermore, at room temperature,
PBSCN3 and PBSCN4 show higher increments in G
(82% and 248%, respectively) than that
of neat PBS, while those of PBSCN1 and PBSCN2 are 18.5% and 44% higher, respectively.
At 90 C, only PBSCN4 exhibits very strong enhancement of G
compared to that of the other
three PBSCNs.
In Figure 21, Okamoto summarizes the clay content dependence of G
of various types
of nanocomposites obtained well below Tg . The Einstein coefficient (kE ) derived by using
Halpin and Tai’s theoretical expression (modified by Nielsen) is shown in Figure 21, and
represents the aspect ratio (Lclay /dclay ) of dispersed clay particles without intercalation.
From Figure 21, it can be clearly observed that PBSCNs show very high increment in G

compared to other nanocomposites having the same clay content in the matrix. PPCNs are
well known for intercalated systems; N6CNs are already well established exfoliated nanocom-
posites; PLACNs will soon be established intercalated-and-flocculated nanocomposites; and
PBSCNs are intercalated-and-extended flocculated nanocomposites systems [50, 51]. Due to
the strong interaction between hydroxylated edge–edge groups, the clay particles are some-
times flocculated in the polymer matrix. As a result of this flocculation, the length of the clay
particles increases enormously, as does the overall aspect ratio. For the preparation of high-
molecular-weight PBS, di-isocyanate end-groups are generally used as chain extenders [65].
These isocyanate end-group chain extenders make urethane bonds with hydroxy-terminated,
low-molecular-weight PBS, and each high-molecular-weight PBS chain contains two such
bonds (see the schematic illustration in Figure 22). These urethane-type bonds lead to strong
interaction with the silicate surface by forming hydrogen bonds and, hence, strong floccula-
tion (see Fig. 23). For this reason, the aspect ratio of dispersed clay particles is much higher
in PBSCNs compared to all other nanocomposites, which results in high enhancement of
the modulus.

7.1.2. Tensile Properties


The effect of clay content on the tensile properties of PLC/LS nanocomposites has been
studied (see Tables 6 and 7). The Young’s modulus of nanocomposites is higher compared
to neat PCL, as result of their intercalated/exfoliated structure. For instance, the Young’s
modulus is significantly increased from 216 MPa for pure PCL to more than 390 MPa for
the composite that contains 10 wt% of MMT-(OH)2 [66]. PCL is a ductile polymer able to
sustain large deformations (700% at break). Adding nanoclays only slightly decreases the
elongation at break, as shown in Tables 6 and 7. PCL remains ductile with an elongation at

10
PLACN1
PLACN2 160 70 15
PLACN3 T=20 °C
PLACN4
G′nanocomposite /G′matrix

PLACN5
PLACN6

N6CN1.6
T=0 °C N6CN3.7
PBSCN1
PBSCN2
PBSCN3
T=–50 °C PBSCN4
1 PBSCN5
PBSCN6
0.1 1 10 100
Vol % of clay

Figure 21. Plots of G


nanocomposite /G
matrix versus volume percent of clay for various nanocomposites. The Einstein
coefficient (kE ) is shown with the number in the box. The lines show the calculated results from Halpin and Tai’s
theory with various kE s.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 23

C O
O N O
N C
H
H

PBS, Mw is 50300 gm/mol.

Figure 22. Formation of urethane bondings in high-molecular-weight PBS.

break higher than 550%. However, higher clay content (10 wt%) has a detrimental effect,
as confirmed by an ultimate elongation lower than 10% for the two types of OMLS.

7.1.3. Flexural Properties


Flexural modulus and strength of pure PLA and various PLACNs measured at 25 C are
summarized in Table 8. Compared to that of pure PLA, there is a significant increase in
flexural modulus for all PLACNs except PLA/qC16 -SAP nanocomposites. The high modulus
value in PLA/qC13 (OH)-mica and the low modulus value in PLA/qC16 -SAP may be due
to the different aspect ratios of dispersed clay particles in the PLA matrix. On the other
hand, flexural strength is also remarkably increased in PLA/C18 -MMT, PLA/qC18 -MMT, and
PLA/qC218 -MMT, but there is not such a large increase in the cases of PLA/qC13 (OH)-mica
or PLA/qC16 -SAP. This behavior may be due to some brittleness appearing in materials in
the presence of qC13 (OH)-mica or qC16 -SAP, as revealed by distortion values (see Table 8).

7.1.4. Heat Distortion Temperature


Heat distortion temperature (HDT) of a polymeric material is an index of heat resistance
toward applied load. Most PLS nanocomposite studies report HDT as a function of clay
content, characterized by ASTM D-648. Kojima and his colleagues [13] first showed that
the HDT of pure nylon 6 increases up to 80 C after nanocomposite preparation with MMT.
In another investigation [67], they reported clay content dependence of HDT in nylon 6/LS
nanocomposites. There is a marked increase in HDT, from 65 C for the neat nylon 6 to
150 C for 4.7 wt% nanocomposites; beyond the weight percent of MMT, the HDT of the
nanocomposites levels off. They also conducted HDT tests of various nylon 6/LS nanocom-
posites prepared with clay having different lengths, and found that HDT also depends upon
the aspect ratio of dispersed clay particles [13].
Since the degree of crystallinity of nylon 6/LS nanocomposites is independent of the amount
and nature of clay, the HDT of nylon 6/LS nanocomposites is related to the presence of strong
interactions between matrix and silicate surfaces by forming hydrogen bonds (see Fig. 24).
Although nylon 6 in nanocomposites stabilizes in a different crystal phase (-phase) than that
found in pure nylon 6, this different crystal phase is not responsible for the higher mechanical
properties of nylon 6/clay nanocomposites because the -phase is a very soft crystal phase.

N N N N
H N
H H H H
O O O O O O O O O O OH HO O Si O O O O O O O O O
Si Si Si Si Si Si Si Si Si Si Si Si Si Si Si Si Si Si Si
OH HO

Figure 23. Formation of hydrogen bonds between PBS and clay, which leads to flocculation of the dispersed
silicate layers.
24 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

Table 6. Tensile properties of PCL/LS nanocomposites containing MMT-Alk.

Elongation
Sample OMLS (wt%) Young’s Modulus (Mpa) at Break (%) Tensile Strength (Mpa)

PCLC9 1 262 ± 13 659 ± 27 33 ± 1


PCLC10 3 282 ± 9 528 ± 58 26 ± 3
PCLC11 5 307 ± 18 598 ± 43 28 ± 1
PCLC12 10 371 ± 15 9±1 18 ± 1

Source: Reprinted with permission from [66], B. Lepoittevin et al., Polymer 42, 4017 (2001). © 2001, Elsevier
Science Ltd.

On the other hand, the increased mechanical properties of nylon 6/LS nanocomposites that
accompany increasing clay content is due to the mechanical reinforcement effect.
Okamoto et al. examined the HDT of neat PLA and various PLA/LS nanocompos-
ites (PLACNs) with different load conditions. As seen in Figure 25(a) [48], in PLACN7
(inorganic clay content = 5 wt%) there is a marked increase of HDT with an intermedi-
ate load of 0.98 MPa, from 76 C for the neat PLA to 98 C for PLACN4 (inorganic clay
content = 3 wt%). The value of HDT gradually increases with increasing organoclay content;
in PLACN7, the value increases up to 111 C.
On the other hand, imposed load dependence on HDT is clearly observed in PLACNs.
Figure 25(b) shows the typical load dependence in PLACN7. The increase of HDT of neat
PLA due to nanocomposite preparation is a very important property improvement, not only
from the industrial point of view but also pertaining to molecular control on the silicate
layers (i.e., crystallization through interaction between PLA molecules and SiO4 tetrahedral
layers in the MMT). When there is a high load (1.81 MPa), it is very difficult to achieve
high HDT enhancement without strong interaction between the polymer matrix and organo-
MMT [13]. With all of the PLACNs studied here, the values of the melting temperature
(Tm ) do not change significantly as compared to that of neat PLA. Thus, the improvement of
HDT with an intermediate load (0.98 MPa) originates in the better mechanical stability
of PLACNs due to mechanical reinforcement by the dispersed clay particles, higher levels of
crystallinity (
c ), and intercalation. This is qualitatively different from the behavior of a nylon
6/LS nanocomposite system, where the MMT layers stabilize in a different crystalline phase
(-phase) [67] than that found in neat nylon 6, with the strong hydrogen bonding between
the silicate layers and nylon 6 as a result the discrete lamellar structure on both sides of the
clay (see Fig. 24).

7.1.5. Izod-Impact Properties


Table 9 shows the Izod-impact strength values for PCL/LS nanocomposites prepared with
MMT-Na+ , MMT-Alk of MMT-(OH)2 as a function of clay content [66]. The Izod-impact
strength systematically decreases with increasing clay content. It drops from 48 J/m for
unfilled PCL to 33 and 13 J/m when 1 wt% and 10 wt%, respectively, of the MMT-(OH)2
nanoclay are incorporated.

7.1.6. Thermal Stability


The thermal stability of polymeric materials is usually studied by thermogravimetric analysis
(TGA). The weight loss due to the formation of volatile degradation products is monitored
as a function of temperature ramp. When the heating is conducted under an inert gas
Table 7. Tensile properties of PCL/LS nanocomposites containing MMT-(OH)2 .

Elongation
Sample OMLS (wt%) Young’s Modulus (Mpa) at Break (%) Tensile Strength (Mpa)

PCLC1 1 259 ± 11 705 ± 47 36 ± 2


PCLC2 3 272 ± 16 563 ± 62 25 ± 4
PCLC3 5 313 ± 23 560 ± 46 24 ± 3
PCLC4 10 399 ± 23 7±1 17 ± 0.5

Source: Reprinted with permission from [66], B. Lepoittevin et al., Polymer 42, 4017 (2001). © 2001, Elsevier
Science Ltd.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 25

Table 8. Comparison of practical materials properties of pure PLA and various nanocomposites.

Practical PLA/ PLA/ PLA/ PLA/ PLA/


Materials Properties PLA C18 -MMT qC218 -MMT qC18 -MMT qC16 -SAP qC13 (OH)-mica

Flexural modulus 4.84 5.66 5.43 5.57 4.5 6.11


(GPa at 25 C)
Flexural strength 86 132 102 134 93 94
(/MPa at 25 C)
Distortion at break (%) 1.9 3.2 3.9 3.1 1.5 1.5
HDT ( C with 0.98 MPa 76 94 91.3 93 98 93
load)
O2 gas permeability 200 172 171 177 120 71
(mL/mm) (m2 /day.MPa)
O2 gas permeability 200 168 167 178 169 68
(mL/mm) (m2 /day.MPa)a
a
Calculated on the basis of the Nielsen theoretical equation (Eq. (2)) in this chapter.
Source: Reprinted with permission from [63], S. Sinha Ray et al., Macromol. Rapid Commun. 24, 815 (2003). © 2003, Wiley-VCH
Verlag GmbH & Co.

17.2 Å

O H O

N
N N n

H
H O H
H

O O O O O O O
Si Si Si Si Si Si Si

Figure 24. Schematic illustration of the formation of hydrogen bonds in nylon 6/MMT nanocomposites.

(a) 120 (b) 160


PLA
110 140 PLACN7

120
HDT / °C

HDT / °C

100

100
90
80
80
Load = 0.98 MPa 60
70
0 2 4 6 8 0.4 0.8 1.2 1.6 2
Organoclay / wt.% Load /MPa

Figure 25. (a) OMLS (wt %) dependence of HDT of neat PLA and various PLACNs. (b) Load dependence of
HDT of neat PLA and PLACN7. Reprinted with permission from [48], S. Sinha Ray et al., Polymer 44, 857 (2003).
© 2003, Elsevier Science Ltd.

Table 9. Izod-impact properties of PCL/LS nanocomposites contain-


ing MMT-Na, MMT-Alk, and MMT-(OH)2 .

Izod-impact (J/m)
OMLS (wt%) MMT-Na MMT-Alk MMT-(OH)2

1 33 ± 5 28 ± 6 33 ± 3
3 22 ± 2 22 ± 2 18 ± 3
5 19 ± 1 15 ± 1 13 ± 1
10 15 ± 1 16 ± 3 13 ± 1
Source: Reprinted with permission from [66], B. Lepoittevin et al., Polymer
42, 4017 (2001). © 2001, Elsevier Science Ltd.
26 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

flow, a nonoxidative degradation occurs, while the use of air or oxygen allows following the
oxidative degradation of the samples. Generally, the incorporation of clay in the polymer
matrix enhances thermal stability by acting as a superior insulator and mass transport barrier
to the volatile products generated during decomposition.
The thermal stability of the PCL/LS nanocomposites has also been studied by TGA.
Generally, the degradation of PCL fits a two-step mechanism [35]: First, a statistical rup-
ture of the polyester chains by pyrolysis of ester groups with the release of CO2 , H2 O,
and hexanoic acid; and second, -caprolactone (cyclic monomer) is formed as a result of
an unzipping depolymerization process. The thermograms of nanocomposites prepared with
organoclay and pure PCL recovered after clay extraction are presented Figure 26. Both
intercalated and exfoliated nanocomposites show higher thermal stability when compared
to pure PCL or the corresponding microcomposites. The nanocomposites reached a high
of 25 C in decomposition temperature at 50% weight loss. The shift of the degradation
temperature may be ascribed to: (1) a decrease in oxygen; (2) a decrease in the permeabil-
ity/diffusivity of volatile degradation products due to the homogeneous incorporation of clay
sheets; (3) a barrier of high aspect ratio fillers; and (4) char formation. The thermal stability
of nanocomposites systematically increases with increasing clay; however, above a loading of
5 wt%, the thermal stability is no longer improved.
A completely different behavior is observed in synthetic biodegradable aliphatic polyester
(BAP)/clay nanocomposite systems. Here, the thermal degradation temperature and thermal
degradation rate are systematically increased with an increasing amount of organoclay, up
to 15 wt% [68]. Like PS/LS nanocomposites, a small amount of clay also increased the
residual weight of BAP/OMMT because of the restricted thermal motion of the polymer in
the silicate layers. The residual weight of various materials at 450 C increased in the order
BAP < BAP03 < BAP06 < BAP09 < BAP15 (here, the number indicates the weight percent
of clay). This type of improved thermal properties is also observed in other systems like
SAN [69], the intercalated nanocomposites prepared by emulsion polymerization.
Many researchers believe the role of clay in nanocomposite structure might be the main
reason for the difference in TGA results of these systems compared to the so far reported
systems. The clay acts as a heat barrier that could enhance the overall thermal stability of
the system, as well as assisting in the formation of char after thermal decomposition. Thus,
in the beginning stage of thermal decomposition, the clay could shift the decomposition
temperature higher. However, after that, this heat barrier effect would result in a reverse
thermal stability. In other words, the stacked silicate layers could hold accumulated heat

100

75
weight (%)

PCL
50

1 wt%

3 wt%
25
5 wt%

10 wt%
0
250 300 350 400 450 500
temperature °C

Figure 26. Temperature dependence of weight loss under an air flow for neat PCL and PCL nanocomposites
containing 1, 3, 5, and 10 wt% (relative to inorganics) of MMT-Alk (heating rate: 20 K/min). Reprinted with
permission from [35], B. Lepoittevin et al., Polymer 43, 1111 (2002). © 2002, Elsevier Science Ltd.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 27

that could be used as a heat source to accelerate the decomposition process, in conjunction
with the heat flow supplied by an outside heat source.

7.1.7. Gas Barrier Properties


Nanoclays are believed to increase the barrier properties by creating a maze or “tortuous
path” (see Fig. 27) that retards the progress of gas molecules through the matrix resin.
The direct benefit of the formation of this type of path is clearly observed in polyimide/clay
nanocomposites, which show dramatically improved barrier properties with simultaneous
decrease in the thermal expansion coefficient [70, 71]. The polyimide/LS nanocomposites
revealed a several-fold reduction in the permeability of small gases (e.g., O2 , H2 O, He,
CO2 , and the organic vapor ethylacetate) with the presence of a small fraction of OMLS.
For example, at 2 wt% clay loading, the permeability coefficient of water vapor was decreased
ten-fold for synthetic mica relative to pristine polyimide. By comparing nanocomposites
made with layered silicates of various aspect ratios not only was the permeability noted to
decrease with increasing aspect ratio.
Okamoto and Yamada [72] measured the O2 gas permeability for exfoliated PLA/synthetic
mica nanocomposites. Figure 28 shows a plot of the relative permeability coefficient value
as a function of the weight percent of clay, PPLACN /PPLA , where PPLACN and PPLA are the
nanocomposite and pure PLA permeability coefficients, respectively. The curve fitting was
achieved by using the Nielsen theoretical expression [77], allowing the prediction of gas
permeability as a function of the length and width of the filler particles as well as their
volume fraction within the PLA matrix.
In the Nielsen model [73], where platelets of length ( Lclay ) and width ( Dclay ) of the
clay (which are dispersed parallel in the polymer matrix), the tortuosity factor () can be
expressed as:
 
Lclay
 =1+ clay (1)
2Dclay

where clay is the volume fraction of dispersed clay particles. Therefore, the relative perme-
ability coefficient (PPLS nano /PNeat ) is given by:
PPCNnano 1
=  −1 = (2)
PNeat 1 + !Lclay /2Dclay "clay #

where PPCNnano and PNeat are the permeability coefficients of PCN and neat polymer,
respectively.

7.1.8. Biodegradability
Another most interesting and exciting aspect of nanocomposite technology is the signif-
icant improvement in the biodegradability of biodegradable polymers after nanocompos-
ite preparation with OMLS. Aliphatic polyesters are among the most promising materials
for the production of environmentally friendly biodegradable plastics. Biodegradation of
aliphatic polyester is well known, in that some bacteria degrade them by producing enzymes
which attack the polymer. Tetto et al. [74] first reported on the biodegradability of PCL-
based nanocomposites, where PCL/LS nanocomposites showed improved biodegradability
compared to pure PCL. According to these researchers, the improved biodegradability of

“Tortuous path” in layered silicate


Conventional composites nanocomposites

Figure 27. Formation of “tortuous paths” in polymer/clay nanocomposites.


28 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

O2 TR / ml.mm.m–2.day–1.MPa–1
200 Theoretical curve based on
Lclay /Dclay = 275
Experimental value
150

100

50

0
0 2 4 6 8 10
OMSFM / wt %

Figure 28. Oxygen gas permeability of neat PLA and various PLACNs as a function of organoclay content measured
at 20 C and 90% relative humidity. The filled circles represent the experimental data. Theoretical fits are based on
the Nielsen tortuosity model.

PCL after nanocomposite formation may be due to the catalytic role of the OMLS in the
biodegradation mechanism. But it is still unclear how clay increases the biodegradation rate
of PCL.
In 2002, Lee et al. [55] reported on the biodegradation of aliphatic polyester-based
nanocomposites under compost. Figure 29(a) and (b), respectively, show the clay content
dependence of biodegradation of APES-based nanocomposites prepared with two differ-
ent types of clays. These researchers assumed that the biodegradation was retarded due to
improvement of the barrier properties of the aliphatic APSE after nanocomposite prepara-
tion with clay. However, they provided no data about permeability.
Very recently, Yamada and Okamoto et al. [42, 48] first reported on the biodegradability of
neat PLA and corresponding nanocomposites prepared with trimethyl octadecylammonium-
modified MMT (C3 C18 -MMT) with details mechanism. The compost used was prepared from
food waste and tests were carried out at 58 C ± 2 C. Figure 30(a) shows an actual picture
of the samples of neat PLA and PLACN4 (C3 C18 -MMT = 4 wt%) recovered from compost
with time. The decreased molecular weight (Mw ) and residual weight percentage (Rw ) of the
initial test samples with time are shown in Figure 30(b). The biodegradability of neat PLA
is significantly enhanced after PCN preparation. Within one month, the decrease in Mw and
the extent of weight loss are almost the same for both PLA and PLACN4. However, after
one month a sharp change occured in weight loss of PLACN4, and within two months it was
completely degraded in compost. The degradation of PLA in compost is a complex process
involving four main phenomena: (1) water absorption; (2) ester cleavage and formation of
oligomer fragments; (3) solubilization of oligomer fragments; and (4) diffusion of soluble
oligomers by bacteria [75]. Therefore, the factor that increases the hydrolysis tendency of
PLA ultimately controls the degradation of PLA.

100 100
APES (a) APES/10A (97/3 wt %) (b)
APES30B (97/3 wt %) APES/10A (95/5 wt %)
APES30B (95/5 wt %) APES/10A (90/10 wt %)
Biodegradability (wt%)

80 80
Biodegradability (wt%)

APES/10A (90/20 wt %)
APES30B (90/10 wt %)
APES/10A (70/30 wt %)
APES30B (80/20 wt %)
APES
APES30B (70/30 wt %)
60 60

40 40

20 20

0 0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
Time (Day) Time (Day)

Figure 29. Biodegradability of APES nanocomposites with: (a) Closite 30B and (b) Closite 10A. Reprinted with
permission from [55], S. R. Lee et al., Polymer 43, 2495 (2002). © 2002, Elsevier Science Ltd.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 29

(a) after 32 days after 50 days after 60 days

PLA

PLACN4

(b)
PLACN4
PLA
150

MW x 10–3 /(gm/mol)
100
Rw /%

80 100

60

40 50

20
PLACN4
0
0 PLA

0 10 20 30 40 50 60 70
Time /days

Figure 30. (a) Actual picture of biodegradability of neat PLA and PLACN4 recovered from compost with time.
The initial shape of the crystallized samples was 3 × 10 × 0%1 cm3 . (b) Time dependence of residual weight, Rw and
of matrix, Mw of PLA and PLACN4 under compost at 58 ± 2 C. Reprinted with permission from [48], S. Sinha
Ray et al., Polymer 44, 857 (2003). © 2003, Elsevier Science Ltd.

These researchers concluded that the presence of terminal hydroxylated edge groups on
the silicate layers may be one of the factors responsible for this behavior. In the case of
PLACN4, the stacked (approximately four layers) and intercalated silicate layers are homo-
geneously dispersed in the PLA matrix (from TEM imaging) and these hydroxy groups
start heterogeneous hydrolysis of the PLA matrix after absorbing water from the compost.
Because this process takes some time to start, the weight loss and degree of hydrolysis of
PLA and PLACN4 are almost the same within up to one month [see Fig. 30(b)]. However,
after one month there is a sharp weight loss in the PLACN4 compared to the PLA. That
means that one month is a critical time at which to start heterogeneous hydrolysis; due to
this type of hydrolysis, the matrix breaks into very small fragments and disappears with com-
posting. This assumption was confirmed by conducting the same type of experiment with
PLACN prepared with dimethyl dioctdecyl ammonium salt-modified synthetic mica, which
has no terminal hydroxylated edge group. The degradation tendency was almost the same
as that of neat PLA [76].
Yamada and Okamoto et al. [41, 76] also conducted respirometric tests to study degrada-
tion of the PLA matrix in a compost environment at 58 C ± 2 C. For this test, the compost
was made from bean curd refuse, food waste, and cattle feces. Rather than weight loss, which
reflects the structural changes in the test sample, CO2 evolution provides an indicator of
the ultimate biodegradability of PLA in PLACN4. The PLA in PLACN4 was prepared with
(N (cocoalkyl)N , N -[bis(2-hydroxyethyl)]-N -methylammonium-modified synthetic mica); in
other words, the samples were mineralized. Figure 31 shows the time dependence of the
30 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

100 200

Degree of Biodegradation/%
(a) (b) Neat PLA
80 PLACN4

Mw×10–3 /g.mol –1
150

60
100
40

50
20 Neat PLA
PLACN4
0 0
0 10 20 30 40 50 0 2 4 6 8 10 12 14
Time/Days Time/Days

Figure 31. Degree of biodegradation (i.e., CO2 evolution), and (b) time-dependent change of matrix Mw of neat
PLA and PLACN4 (MEE clay = 4 wt%) under compost at 58 ± 2 C. Reprinted with permission from [41],
S. Sinha Ray et al., Macromol. Rapid Commun. 23, 943 (2002). © 2002, Wiley-VCH Verlag GmbH & Co.

degree of biodegradation of neat PLA and PLACN4, indicating that the biodegradability
of PLA in PLACN4 is significantly enhanced. The presence of OMLS may thus cause a
different mode of attack on the PLA component, which might be due to the presence of
hydroxy groups. Details of the mechanism of biodegradability are presented in the relevant
literature [41, 76].
K. Okamoto and M. Okamoto also investigated the biodegradability of neat PBS before
and after nanocomposite preparation with three different types of OMLS. They used alky-
lammonium or alkylphosphonium salts for the modification of pristine layered silicates, and
these surfactants are toxic for microorganisms.
Figure 32(a) shows actual pictures of samples of neat PBS and various nanocompos-
ites recovered from compost after 35 days. This figure clearly shows that many cracks
appeared in the nanocomposite samples compared to that of neat PBS. This confirms the
improved degradability of nanocomposites in compost. This kind of fracture is advantageous
for biodegradation because it creates much more surface area for further attack by microor-
ganisms. It should be noted that the extent of fragmentation is directly related to the nature
of the OMLS used for nanocomposite preparation. These researchers also conducted gel
permeation chromatography (GPC) measurement of the samples recovered from compost,
and they found that the extent of molecular weight loss was almost the same for all samples
(see Table 10). This result indicates that the extent of hydrolysis of PBS in a pure state or
in OMLS-filled systems is the same as in compost.

(a) (b)

PBS PBS/C18-MMT PBS PBS/C18-MMT

PBS/qC18-MMT PBS/qC16-SAP PBS/qC18-MMT PBS/qC16-SAP

Figure 32. Biodegradability of neat PBS and various nanocomposite sheets (a) under compost, and (b) under soil
fields. Reprinted with permission from [51], K. Okamoto et al., J. Polym. Sci. Part B: Polym. Phys. 41, 3160 (2003).
© 2003, John Wiley & Sons, Inc.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 31

Table 10. GPC results of various samples recovered from compost after 35 days.

Samples Mw × 10−3 (g/mol) Mn × 10−3 (g/mol) Mwo × 10−3 (g/mol) Mw /Mwo

PBS 16 3.8 101 0%16


PBS/C18 -MMT 17 6.6 104 0%16
PBS/qC18 -MMT 17 4.4 112 0%15
PBS/qC16 -SAP 8.7 1.2 91 0%096
Mwo is the molecular weight before composting.

Except for the PBS/qC16 -SAP system, the degree of degradation was not different for
other samples. This observation indicates that MMT or alkylammonium cations and other
properties have no effect on the biodegradability of PBS. The accelerated degradation of the
PBS matrix in the presence of qC16 -SAP may be due to the presence of the alkylphosphonium
surfactant. This type of behavior was also observed in the case of PLA/OMLS nanocomposite
systems.
Yamada and Okamoto et al. also observed nature of degradation of PBS and various
nanocomposites under soil fields. These experiments were conducted for one, two, and six
months. After one and two months, there was no change in the nature of the sample surfaces,
but after six months black or red spots appeared on the surface of nanocomposite samples.
Figure 32(b) shows the results of degradation of neat PBS and various nanocomposite sheets
recovered from soil fields after six months. These researchers concluded that these spots on
the sample surfaces were due to fungus attack, because when they put these samples into
a slurry they observed clear fungus growth. These results also indicate that nanocomposites
exhibited the same or higher levels of biodegradability compared to PBS matrices.

8. CRYSTALLIZATION OF BIODEGRADABLE
POLYMER/LS NANOCOMPOSITES
Crystallization of PLS nanocomposites might be a good method for controlling the structure
and various properties of nanocomposites.
Okamoto and Nam reported in detail on the crystallization behavior of PLA/C18 -MMT
with 4 wt% of C18 -MMT as a representative system [77]. To understand the crystallization
kinetics of pure PLA before and after nanocomposite preparation at low Tc (≤120 C), we
used time-resolved LS photometry, which is a powerful tool for estimating the overall crystal-
lization rate and its kinetics in supercooled crystalline polymer liquid [44]. Details regarding
LS experiments can be found elsewhere [78]. For the kinetics of crystallization, integrated
scattering intensity can be employed; the invariant Q is defined as:

Q= I!q"q 2 dq (3)
0

where q [scattering vector = !4*/+LS " sin!,LS /2"] and I!q" is the intensity of the scattered
light at q [78]. In the Hv mode, the invariant Q can be described by the mean-square optical
anisotropy 2 :
Q ∝ 2  ∝ s !-r − -t "2 (4)

where s is the volume fraction of the spherulites, and -r and -t are the radial and tangential
polarizabilities of the spherulites, respectively. We constructed a plot of reduced invariant
Q /Q versus time (t) with Q being Q at an infinitely long time of crystallization (up to
full solidification of the melt).
Figure 33 shows the time variation of the invariant Q /Q taken for pure PLA and
PLA/C18 -MMT at 110 C. The overall crystallization rate was determined from the slope of
Q /Q (d(Q /Q )/dt) in the crystallization region, as indicated by the solid line in Figure 33,
and we plotted in Figure 34. It is clear that the overall crystallization rate increases in
PLA/C18 -MMT, in comparison to pure PLA, as well as the rate increases in PLA/C18 -MMT
for a particular Tc . The same trend is also observed over the wide range of Tc s studied here.
32 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

1.0

0.8

0.6

Qδ / Q∞δ
0.4
Tc=110 °C
0.2 PLACN4
neat PLA

0.0
0 100 200 300 400
Time / sec.

Figure 33. Time variation of reduced invariant Q /Q during isothermal crystallization at quiescent state at
Tc = 110 C. The solid line represents the slope (overall crystallization rate). Reprinted with permission from [63],
S. Sinha Ray et al., Macromol. Rapid Commun. 24, 815 (2003). © 2003, Wiley-VCH Verlag GmbH & Co.

It should be noted that the equilibrium melting temperatures (Tm0 ) of the PLA/C18 -MMT and
pure PLA are the same. The equilibrium melting temperatures were measured by isothermal
crystallization at various temperatures by constructing a Hoffman-Weeks [48] plot, as shown
in Figure 35. Both PLA/C18 -MMT and pure PLA show the same value of Tm0 of 179.5 C;
this would nullify the effect of supercooling .T ! ≡ Tm0 − Tc " on the overall crystallization
rate, linear growth rate, G, etc. The overall crystallization rate with Tc is observed typical
rate curve as usual for semicrystalline polymer. However, the crystallization rate of PLACN4
is enhanced for every measured temperature, especially at higher Tc s. From the onset time
(t0 ), we can estimate the induction time of the crystallization until start of crystallization.
The observed value of t0 at 110 C was 74 s for pure PLA and 56 s for PLA/C18 -MMT. At all
Tc s measured here, the t0 value for PLA/C18 -MMT was lower than that of pure PLA. This
reduction of t0 in PLA/C18 -MMT is attributed to the presence of clay as the nucleating agent.
A typical example of the time variation of the diameter of a spherulite (D) for pure
PLA and PLA/C18 -MMT at higher Tc s is shown in Figure 36(a), and the linear growth rate
[G = 1/2!dD/dt"] of the spherulites is summarized in Figure 36(b). For both pure PLA
and PLA/C18 -MMT, G decreases with increasing Tc in the temperature range of 120–140 C.
However, for PLA/C18 -MMT, G shows a slightly higher value compared to that of pure
PLA. This observation indicates that the dispersed clay particles do not have much effect on
the crystallization and no big acceleration of G in the crystallization of the PLA/C18 -MMT.
This behavior suggests that the diffusion rate of bulk PLA molecules is not enhanced with

10–1

PLACN4
neat PLA

10–2
d(Qδ / Q∞δ)
dt

10–3

10–4
70 80 90 100 110 120 130 140
Crystallization Temperature/ °C

Figure 34. Tc dependence of the overall crystallization rate of pure PLA and PLACN4. Reprinted with permission
from [63], S. Sinha Ray et al., Macromol. Rapid Commun. 24, 815 (2003). © 2003, Wiley-VCH Verlag GmbH & Co.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 33

190
neat PLA
185 PLACN4

180

Tm / °C
175

170

165

160
100 120 140 160 180
Tc / °C

Figure 35. Tm versus Tc (Hoffman-Weeks) plots of pure PLA and PLACN4. Reprinted with permission from [63],
S. Sinha Ray et al., Macromol. Rapid Commun. 24, 815 (2003). © 2003, Wiley-VCH Verlag GmbH & Co.

the addition of clay at every measured temperature; thus, the overall crystallization rate is
affected only by nucleation of the clay particles [79, 80].
Figure 37 demonstrates that the number of heterogeneous nuclei (N ) can be estimated
from a rough approximation [using Eq. (4)]. The calculated value of N at 130 C was
9%3 × 10−7 /m−3 for pure PLA and 55%7 × 10−7 /m−3 for PLA/C18 -MMT. The time variation
of the volume fraction of the spherulites increases in proportion to NG3 ( the overall
crystallization rate). This fact suggests that the overall crystallization rate of PLA/C18 -MMT
at high temperature (Tc = 130 C) is about a one-half order of magnitude higher than that
of matrix PLA without clay. The difference in N between pure PLA and PLA/C18 -MMT
at Tc = 130 C is higher than at low Tc . This suggests that PLA/C18 -MMT exhibits hetero-
geneous nucleation kinetics, which depend on more originating from the well dispersed
clay particles in the matrix at high temperature. It should be noted that the spherulites of
PLA/C18 -MMT have a lower ordering than those of pure PLA, due to the presence of dis-
persed clay particles in the spherulites [77]. Hence, if an aggregation of clay particles (which
are not nucleated during crystallization) exists inside the spherulite, then the occurrence of
the regular orientation of lamella stacks inside the spherulite may be disrupted (see Fig. 38).
Figure 38 shows WAXD profiles of neat PLA and PLACN4 after crystallization at 110 C
for 1.5 hours. The neat PLA exhibits a very strong reflection at 20 = 17%1 degrees due to
diffraction from the (200) and/or (110) planes, and another reflection peak at 20 = 19%5
degrees occurring from the (203) plane. On the other hand, in PLACN4 these peaks are
shifted toward the lower diffraction angle accompanied by another small peak at 20 = 15%3
degrees. After calculation, it was confirmed that this reflection is due to the (010) diffrac-
tion plane. These profiles indicate neat PLA crystals are typical orthorhombic crystals [81];

(a) (b) 0.12


Spherulite Diameter /µm

140 neat PLA neat PLA


120 PLACN4 PLACN4
100 Tc =130 °C 0.11
G / µms–1

80
60
0.10
40
20
0 0.09
0 500 1000 1500 2000 110 120 130 140 150
Crystallization Time / sec Crystallization Temperature /°C

Figure 36. (a) Spherulite diameter as a function of crystallization time at Tc = 130 C; and (b) linear growth rate of
pure PLA and PLACN4 as a function of Tc . Reprinted with permission from [63], S. Sinha Ray et al., Macromol.
Rapid Commun. 24, 815 (2003). © 2003, Wiley-VCH Verlag GmbH & Co.
34 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

1e–1
neat PLA
1e–2
PLACN4

1e–3

1e–4

N /µm–3
1e–5

1e–6

1e–7

1e–8

1e–9
80 90 100 110 120 130 140 150
Crystallization Temperature /°C

Figure 37. Nucleation density (N ) of pure PLA and PLACN4 as a function of Tc . Reprinted with permission
from [63], S. Sinha Ray et al., Macromol. Rapid Commun. 24, 815 (2003). © 2003, Wiley-VCH Verlag GmbH & Co.

however, the PLACN4 sample crystallized in a defect-ridden crystalline form. This unstable
growth of crystallites of PLA in the presence of MMT particles may be due to the interca-
lation of PLA chains into the silicate galleries.

9. MELT RHEOLOGY OF BIODEGRADABLE


POLYMER/LS NANOCOMPOSITES
The measurement of rheological properties of any polymeric material in a molten state is cru-
cial to gain fundamental understanding of the processability of that material. In polymer/LS
nanocomposites, measurements of melt rheological properties are not only important to
understand the nature of processability of these materials, but also to determine the
strength of polymer-layered silicate interactions and the structure-property relationship in
the nanocomposites, because melt rheological behaviors are strongly influenced by their
nanoscale structure and interfacial characteristics.

9.1. Dynamic Oscillatory Shear Measurement


Generally, the rheology of polymer melts strongly depends on the temperature at which mea-
surement is carried out. It is well known that for the thermorheological simplicity, isotherms
of G
!"1 G

!" and complex viscosity !2∗ !"" can be superimposed by horizontal shifts

10000
(110) *
8000 or
(200)
Intensity /A.U.

6000
(203)
* (105)
4000 (010) * * PLACN4

neat PLA
2000
C18MMT

0
10 15 20 25 30
2Θ / degrees

Figure 38. Typical WAXD patterns of pure PLA and PLACN4 samples crystallized at 110 C for 1.5 hours.
Reprinted with permission from [260], J. Y. Nam et al., Macromolecules 36, 7126 (2003). © 2003, American Chemical
Society.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 35

along the frequency axis:

bT G
!aT 1 Tref " = bT G
!1 T "
bT G

!aT 1 Tref " = bT G

!1 T "
2∗ !aT 1 Tref " = 2∗ !1 T "

where aT and bT are the frequency and vertical shift factors, respectively, and Tref is the
reference temperature. All isotherms measured for pure PLA and for various PLA/C18 -MMT
can be superimposed along the frequency axis.
In polymer samples, it is expected that at the temperatures and frequencies at which
the rheological measurements were carried out, the polymer chains should be fully relaxed
and exhibit characteristic homopolymer-like terminal flow behavior (i.e., the curves can be
expressed by a power law of G
!" ∝ 2 and G

!" ∝ ".
The linear dynamic viscoelastic master curves for the neat PLA and various PLACNs are
shown in Figure 39 [48]. These curves were generated by applying the time-temperature
superposition principle and shifting to a common temperature (Tref ), using both the fre-
quency shift factor (aT ) and the modulus shift factor (bT ). The moduli of the nanocompos-
ites increased with increasing clay loading at all frequencies (). At high frequencies, the
qualitative behavior of G
!" and G

!" was essentially same and unaffected by frequen-


cies. However, at low frequencies G
!" and G

!" increased monotonically with increasing


clay content. In the low frequency region, the curves can be expressed by the power law
of G
!" ∝ 2 and G

!" ∝  for neat PLA, suggesting that this is similar to those of


the narrow Mw distribution homopolymer melts. On the other hand, for aT < 5 rad.s−1 ,
the viscoelastic response [particularly G
!"] for all of the nanocomposites displayed sig-
nificantly diminished frequency dependence as compared to the matrices. In fact, for all
PLACNs, G
!" became nearly independent at low aT  and exceeded G

!", characteristic
of materials exhibiting a pseudosolid-like behavior [82]. The terminal zone slope values of
both neat PLA and PLACNs were estimated at the lower aT  region ( < 10 rad.s−1 ), and
are presented in Table 11. The lower slope values and the higher absolute values of the
dynamic moduli indicate the formation of “spatially linked” structures in the PLACNs in the
molten state [83]. Because of this structure or highly geometric constraints, the individual
stacked silicate layers are incapable of freely rotating. Hence, by imposing small aT , the
relaxations of the structure are almost completely prevented. This type of prevented relax-
ation due to the highly geometric constraints of the stacked and intercalated silicate layers
leads to the pseudosolid-like behavior observed in PLACNs. This behavior probably corre-
sponds to the shear-thinning tendency, which dramatically appears in the viscosity curves
(aT  < 5 rad.s−1 ) (2∗  versus aT " [84]. Such features are highly dependent on the shear
rate in the dynamic measurement due to the formation of shear-induced alignment of the
dispersed clay particles [85].
The temperature dependence frequency shift factors (aT , Williams-Landel-Ferry type [86])
used to generate the master curves shown in Figure 39 are shown in Figure 40. The depen-
dence of the frequency shift factors on the silicate loading suggests that the temperature-
dependent relaxation process observed in the viscoelastic measurements are somehow
affected by the presence of silicate layers [82]. In case of nylon 6/LS nanocomposites, where
the hydrogen bonding of the already formed hydrogen-bonded molecule to the silicate sur-
face, the system exhibits a high level of flow activation energy [estimated from slope in
Figure 40(a)], near one order higher in magnitude compared to that of neat nylon 6 [87].
The shift factor (bT ) shows significant deviation from a simple density effect, but it would
be expected that the values would not vary far from unity [86]. One possible explanation is
internal structure development occurring in PLACNs during measurement (shear process).
The alignment of the silicate layers probably supports the PCN melts to withstand the shear
force, thus leading to the increase in the absolute values of G
!" and G

!".
Figure 41 represents the clay content-dependent (weight percent) flow activation energy
(Ea ) of pure PLA and various PLA/C18 -MMTs obtained from Arrhenius fits of master
curves. It is clearly observed that Ea values significantly increased in PLA/C18 -MMTs con-
taining 3 wt% of C18 -MMT and then correlates fairly well with increasing C18 -MMT content.
36 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

105
0.5
104

103 1

bTG′/Pa
2 2
10

101 PLA
0
PLACN4
10 PLACN5
PLACN7
10–1
Tref = 175 °C
4
10
1

bT G′′/Pa
103 2

102

101
|η*|/Pa.s

104

103
10–2 10–1 100 101 102
aTω /rad.s–1

Figure 39. Reduced frequency dependence of storage modulus, loss modulus, and complex viscosity of neat PLA
and various PLACNs. Reprinted with permission from [63], S. Sinha Ray et al., Macromol. Rapid Commun. 24, 815
(2003). © 2003, Wiley-VCH Verlag GmbH & Co.

Table 11. Terminal slopes of G


and G

versus aT  for PLA


and various PLACNs.

System G
G

PLA 1%3 0%9


PLACN4 0%2 0%5
PLACN5 0%18 0%4
PLACN7 0%17 0%32

(a) (b)
101
0 Tref = 175 °C Tref =175 °C PLA
10 PLACN4
PLACN5
PLACN7
bT

100
aT

10–1

PLA
PLACN4
PLACN5
PLACN7
10–2 10–1
2.08 2.12 2.16 2.2 2.24 2.08 2.12 2.16 2.2 2.24
1/T ×103/K 1/T × 103/K

Figure 40. (a) Frequency shift factors (aT ) and (b) modulus shift factors (bT ) as a function of temperature.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 37

240

220

Ea/KJ.mol–1
200

180

Tref = 175 °C
160

–1 0 1 2 3 4 5 6
MMT/ wt.%

Figure 41. Flow activation energy of pure PLA and various PLA/C18 -MMT nanocomposites as a function of MMT
content. Reprinted with permission from [63], S. Sinha Ray et al., Macromol. Rapid Commun. 24, 815 (2003).
© 2003, Wiley-VCH Verlag GmbH & Co.

This result indicates that in the presence of MMT, it is very difficult for the materials to flow.
This behavior is also ascribed to the formation of spatially linked structures in PLA/C18 -
MMTs in molten states.

9.2. Steady Shear Flow


The steady shear rheological behaviors of neat PBS and various PBSCNs are shown in
Figure 42. The steady viscosity of PBSCNs is enhanced considerably with time at all shear
rates, and at a fixed shear rate the steady viscosity increases monotonically with increasing
silicate loading [50]. On the other hand, all intercalated PBSCNs exhibit strong rheopexy
behavior, and this becomes prominent at low shear rates, while neat PBS exhibits a time-
independent viscosity at all shear rates. With increasing shear rates, the shear viscosity attains
a plateau after a certain time, and the time required to attain this plateau decreases with
increasing shear rates. This type of behavior may be due to the planer alignment of the clay
particles toward the flow direction under shear. When the shear rate is very slow (0.001 s−1 ),
clay particles take a longer time to attain complete planer alignment along the flow direction,
and this measurement time (1,000 s) is too short to attain such alignment. Therefore, strong
rheopexy behavior results. On the other hand, under high shear rates (0.005 s−1 or 0.01 s−1 ),
this measurement time is long enough to attain such alignment, and hence, nanocomposites
show time-independent shear viscosity after certain periods of time.
Figure 43 shows shear rate dependence of viscosity for neat PBS and corresponding
nanocomposites measured at 120 C. Neat PBS exhibits almost Newtonian behavior at all
shear rates, whereas nanocomposites exhibit non-Newtonian behavior. At very low shear
rates, shear viscosity of nanocomposites initially exhibits some shear-thickening behavior that
corresponds to the rheopexy behavior we observed at very low shear rates (see Fig. 42).
After that initial period, all nanocomposites show very strong shear-thinning behavior at all
shear rates; this behavior is analogous to the results obtained with dynamic oscillatory shear
measurements [48]. Additionally, at very high shear rates the viscosities of nanocomposites
are comparable to that of neat PBS. These observations suggest that the silicate layers are
strongly oriented toward the flow direction at high shear rates, and shear-thinning behavior
at high shear rates is dominated by that of neat polymers.
PLS nanocomposites always exhibit significant deviation from the empirical Cox-Merz
relation [88], while all neat polymers obey that relation. (The Cox-Merz relation
requires that for ˙ = , the viscoelastic data should obey the relationship 2!" ˙ = 2∗ !").
They speculated there are two possible reasons for nanocomposite deviation from the Cox-
Merz relation: First, this rule is only applicable for homogenous systems like homopolymer
melts, but nanocomposites are heterogeneous systems. For this reason, this relation works
well in the case of neat polymers [50]. Second, the structure formation is different when
nanocomposites are subjected to dynamic oscillatory shear and steady shear measurements.
38 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

104
PBS Temperature =120 °C

103 Shear rate =0.001s–1


Shear rate =0.005s–1
Shear rate = 0.01s–1
102
PBSCN2
105

104

103
PBSCN3
105
η/Pa.s
104

103
PBSCN4
105

104

103
PBSCN6

104

103
100 101 102 103
Time/ s

Figure 42. Time variation of shear viscosity for PBSCN. Reprinted with permission from [50], S. Sinha Ray et al.,
Macromolecules 36, 2355 (2003). © 2003, American Chemical Society.

9.3. Elongational Flow and Strain-Induced Hardening


Okamoto and his colleagues [89] first conducted elongation tests of PP/LS nanocomposites
(PPCN4) in the molten state at constant Hencky strain rates (˙0 ), using elongation flow
optorheometry [90]. They also attempted to control the alignment of the dispersed silicate
layers with nanometer dimensions of intercalated PPCNs under uniaxial elongational flow.
Figure 44(a) shows double-logarithmic plots of transient elongational viscosity (2E ) against
time (t) observed for PLA/C18 -MMT containing 4 wt% of OMLS at 170 C with differ-
ent Hencky strain rates (˙0 ) ranging from 0.01–1 s−1 . We observed a very strong tendency

106
Temp. = 120 °C PBS
5 PBSCN2
10
PBSCN3
PBSCN4
η/Pa.s

104 PBSCN6

103

102
10–4 10–3 10–2 10–1 100 101
.
γ/s–1

Figure 43. Shear viscosity as a function of shear rates for the shear rate sweep test. Reprinted with permission
from [50], S. Sinha Ray et al., Macromolecules 36, 2355 (2003). © 2003, American Chemical Society.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 39

(a) (b)
108 101
–1 PLACN3
Hencky strain rate/s

Elongational viscosity / Pa.s


PLACN3
1 Temperature = 170 °C
107 Temperature = 170 °C
0.5

Hencky strain
0.1
106 0.05 100
0.01

105

104 10–1 –3
0.1 1 10 100 1000 10 10–2 10–1 100 101
Time / s Hencky strain rate /s–1

Figure 44. (a) Time variation of elongational viscosity for PLA/C18 -MMT(4) melt at 170 C; (b) Strain rate depen-
dence of uprising Hencky strain. Reprinted with permission from [63], S. Sinha Ray et al., Macromol. Rapid
Commun. 24, 815 (2003). © 2003, Wiley-VCH Verlag GmbH & Co.

toward strain-induced hardening in the PLA/C18 -MMT melt. In the early stage, 2E gradually
increases with t but almost independent of ˙0 , which we generally call the linear region of
the viscosity curve. After a certain time (t2E ), which we call the uprising time [marked with
upward arrows in Figure 44(a)], strongly dependent on ˙0 , we saw rapid upward deviation
of 2E from the curves of the linear region. On the other hand, we tried but were unable to
accurately measure the elongational viscosity of pure PLA. We concluded that very low shear
viscosity of pure PLA is the main reason for this, because the minimum viscosity range of
our instrument was greater than 104 Pa.s. However, we confirmed that neither strain-induced
hardening in elongation nor rheopexy in shear flow took place in pure PLA having the same
molecular weight and polydispersity as PLA/C18 -MMT.
As in PP/LS systems, the extended Trouton rule [320 !; ˙ t) 2E !˙0 ; t)] also does not hold
for PLA/C18 -MMT melts, as opposed to the melt of pure polymers. These results indicate
that in PLA/C18 -MMT, the flow-induced internal structural changes also occur in elongation
flow [89], but the changes are quite different from shear flow. The strong rheopexy observed
in shear measurements for PLA/C18 -MMT at very slow shear rates reflects the fact that the
shear-induced structural change involved a process with an extremely long relaxation time.
Regarding elongation-induced structure development, Figure 44(b) shows Hencky strain
rate dependence of the uprising Hencky strain (2E " = ˙0 × t2E taken for PLA/C18 -MMT at
170 C. The 2E increases systematically with the ˙0 . The lower the value of ˙0 , the smaller
the value of 2E . This tendency probably corresponds to the rheopexy of PLA/C18 -MMT
under slow shear flow.

10. PROCESSING OPERATIONS


In the preceding sections, dynamic measurements indicated the formation of spatially linked
structures in PLA/C18 -MMT melts. Shear measurements revealed very strong rheopexy
behavior in PLA/C18 -MMT under very slow shear fields. Under uniaxial elongation flow,
PLA/C18 -MMT exhibited very high viscosity and a tendency toward strong strain-induced
hardening, which probably originated in the perpendicular alignment of the silicate lay-
ers to the stretching direction. This strain-induced hardening behavior is an indispensable
characteristic for foam processing, due to its capacity to withstand the stretching force expe-
rienced during the latter stages of bubble growth. To evaluate the performance potential of
biodegradable PLA/C18 -MMT in foam applications, Okamoto et al. conducted foam process-
ing of one representative nanocomposite, PLA/C18 -MMT, using a newly developed pressure
cell technique. The goal was to devlop an advanced biodegradable foam with excellent mate-
rials properties [91].
Figure 45 shows typical scanning electron microscope (SEM) images of the freeze frac-
ture surfaces of foamed neat PLA and two different foamed nanocomposites. PLA and
PLA/MMT(ODA)5 were foamed at 140 C and PLA/MMT(SBE)5 was foamed at 165 C.
40 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

Neat PLA 100 µm 000000 PLA/ODA5 10 µm 000000 PLA/SBE5 5 µm 000000

Figure 45. SEM images of freeze fracture surfaces of neat PLA and two different nanocomposite foams. Reprinted
with permission from [91], Y. Fujimoto et al., Macromol. Rapid Commun. 24, 457 (2003). © 2003, Wiley-VCH
Verlag GmbH & Co.

All of the foams exhibited desirable closed-cell structures. Homogeneous cells were formed
in both nanocomposites, while the neat PLA foam showed nonuniform cell structures having
large cell size (∼230 /m). Also, the nanocomposite foams showed smaller cell size (d) and
greater cell density (Nc ) compared to neat PLA foam, suggesting that the dispersed silicate
particles act as nucleating sites for cell formation [92, 93].
For the nanocomposite foams, we calculated the distribution function of cell size from
SEM images, and these are presented in Figure 46. The nanocomposite foams conformed
well to Gaussian distributions. In the case of PLA/MMT(SBE)5 [see Fig. 46(b)], we can
see that the width of the distribution peaks, which indicates the cell size dispersity, became
narrow, accompanied by finer dispersion of silicate particles. From the SEM images, we
quantitatively calculated various morphological parameters of two different nanocomposite
foams, which are summarized in Table 12.
From Table 12, we can see that the PLA/MMT(SBE)5 (nanocellular) foam has a smaller
d value ( 360 nm) and a huge Nc ! 1%2 × 1014 cell · cm−3 " compared to that of
PLA/MMT(ODA)5 (microcellular) foam (d 2%59 /m and Nc 3%56 × 1011 cell · cm−3 ).
These results indicate that the nature of the dispersion has vital role in controlling the size of
cells during foaming. On the other hand, the very high value of Nc in the PLA/MMT(SBE)5
foam indicates that the final 5f is controlled by the competitive process in the cell nucle-
ation (its growth and coalescence). Cell nucleation in nanocomposite systems took place
in the boundary between the matrix polymer and the dispersed silicate particles. For this
reason, cell growth and coalescence are strongly affected by the characteristic parame-
ter (see Table 13) and the storage and loss modulus ( viscosity component) of the
materials during processing. This may create nanocellular foams without the loss of mechan-
ical properties in polymeric nanocomposites. Okamoto et al. described a novel foaming
approach for biodegradable polylactide/layered silicate nanocomposites that results in con-
trolled structure of nanocomposite foams (from microcellular to nanocellular).

35 30
30 PLA/ODA5 25 PLA/SBE5

25
20
Fraction / %

Fraction / %

20
15
15
10
10
5 5

0 0
0 1 2 3 4 5 6 7 8 9 10 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Cell size / µm Cell size / µm

Figure 46. Cell size distribution of two different nanocomposite foams. Average values of d in /m and variances
6d2 in /m2 in the Gaussian fit through the data are 2.59 and 0.551, respectively, for PLA/ODA5 foam, and 0.36 and
0.011, respectively, for PLA/SBE5 foam. Reprinted from with permission [91], Y. Fujimoto et al., Macromol. Rapid
Commun. 24, 457 (2003). © 2003, Wiley-VCH Verlag GmbH & Co.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 41

Table 12. Morphological parameters of two different nano-


composite foams.

Parameters PLA/ODA5 PLA/SBE5

5f (g.cm−3 ) 0%46 0%57


d(/m) 2%59 0%36
Nc × 10−11 (cell.cm−3 ) 3%56 1172
(/m) 0%66 0%26
d(clay ) 10%1 4%5
d(Lclay ) 5%78 1%8
(Lclay ) 1%47 1%3

11. CREATING POROUS CERAMIC MATERIALS VIA


PLA/LS NANOCOMPOSITES
Very recently, a new route for the preparation of porous ceramic materials from thermo-
setting epoxy/LS nanocomposites was first demonstrated by Brown et al. [94]. This route
offers attractive potential for diversification and applications of PLS nanocomposites (CNs).
Okamoto and coworkers reported on their development of a novel porous ceramic material
via burning of the PLA/LS system (PLACN) [40]. In the PLACN containing 3.0 wt.% inor-
ganic clay. Figure 47 shows a SEM image of the fractured surface of the porous ceramic
material prepared from simple burning of PLACN in a furnace up to 950 C. After complete
burning, as seen in the figure, the PLACN became a white mass with a porous structure.
The bright lines in the SEM image correspond to the edges of the stacked silicate layers.
In the porous ceramic material, the silicate layers form a “house of cards” structure con-
sisting of large plates having lengths of ∼1,000 nm and thicknesses of ∼30–60 nm. This
implies that the further stacked platelet structure is formed during burning. The material
exhibits an open cell-type structure having 100–1,000-nm diameter voids, a BET surface
area of 31 m2 /g−1 , and low density of porous material (0.187 g/mL−1 estimated by the buoy-
ancy method). The BET surface area value of MMT is 780 m2 /g and that of the porous
ceramic material is 31m2 /g, suggesting approximately 25 MMT plates stacked together. When
MMT is heated above 700 C (but below 960 C), first all of the OH groups are eliminated
from the structure and, thus, MMT is decomposed into a nonhydrated aluminosilicate. This
transformation radically disturbs the crystalline network of the MMT, and the resulting
diffraction pattern is indeed often typical of an amorphous (noncrystalline) phase. A rough
estimate of the compression modulus (K) is on the order of ∼1.2 MPa, which is five orders
of magnitude lower than the bulk modulus of MMT (∼102 GPa) [27]. In the stress-strain
curve, the linear deformation behavior is well described in the early stage of the defor-
mation (i.e., the deformation of the material closely resembles that of ordinary polymeric

Table 13. Comparison of some characteristic parameters of PLA/ODA5


and PLA/SBE5.

Parameters PLA/ODA5 PLA/SBE5


−3 −1
Mw × 10 (g.mol ) 161 163
Mw /Mn 1.58 1.61
Tg ( C) 59.2 59.7
Tm ( C) 169.8 169.3
d001 (nm) 3.03 2.85
Lclay (nm) 448 ± 200 200 ± 25
clay (nm) 255 ± 137 80 ± 20
D(nma ) 38 12.37
Lclay (D) 12 18
D(d001 ) 12.5 4.3
G
PLACN /G
bPLA 1.65 1.43
a
Calculated on the basis of Scherrer equation. b G
PLACN and G
PLA are the stor-
age modulus of PLACN and PLA, respectively, at 25 C. Mw and Mw /Mn for neat
PLA are 177,000 g/mol and 1.58, respectively.
42 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

Figure 47. SEM image of porous ceramic material after being coated with a platinum layer (∼10 nm thickness).
Reprinted with permission from [40], S. Sinha Ray et al., Nano Letts. 2, 423 (2002). © 2002, American Chemi-
cal Society.

foams) [95]. This open cell-type porous ceramic material with its “house-of-cards” structure
is expected to provide strain recovery (up to 8% strain) and an excellent energy dissipation
mechanism after unloading in the elastic region, probably having each plate bend like a leaf
spring (see Fig. 48). This new porous ceramic material possesses elastic features and is very
lightweight. This new route for the preparation of porous ceramic materials via burning of
nanocomposites can be expected to pave the way for much a broader range of applications
for PLS nanocomposites.

200

Deformation rate = 5%/min


Temp = 25 °C
150
Stress / KPa.

100

50

A
0
0 5 10 15 20
Strain / %

60 Deformation rate = 5%/min


Temp = 25 °C

45
1st run
2nd run
Stress / KPa.

3rd run
30

15

B
0
0 2 4 6 8
Strain / %

Figure 48. Stress-strain curve (a) and the strain recovery behavior (b) of porous ceramic material under compres-
sion test. The author conducted the compression test using porous ceramic material of 2 × 2 × 1%5 mm3 size.
Reprinted with permission from [40], S. Sinha Ray et al., Nano Letts. 2, 423 (2002). © 2002, American
Chemical Society.
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 43

Figure 49. Schematic illustration of the main concept of PLA/OMLS nancomposites.

12. CURRENT STATUS AND FUTURE PROSPECTS OF


BIODEGRADABLE NANOCOMPOSITES
Development of PLS nanocomposites is one of the latest evolutionary steps in polymer tech-
nology. Nanocomposites offer attractive potential for diversification and new applications of
conventional polymeric materials.
Since the possibility of direct melt intercalation was first demonstrated by Giannelis and
his colleagues [25], melt intercalation has become a mainstream method of preparing inter-
calated polymer nanocomposites without in situ intercalative polymerization. It is a quite
effective technology within the PLS nanocomposite industry.
Biodegradable polymer-based nanocomposites have a great deal of future promise for
potential applications as high-performance biodegradable materials. These are entirely new
types of materials based on plant and other natural materials (OMLS). When disposed of in
compost, these are safely decomposed into CO2 , water, and humus through the activity of
microorganisms. The CO2 and water could become corn or sugar cane again through plant
photosynthesis (see Fig. 49). Undoubtedly, these unique properties originating in controlled
nanostructures pave the way to a much broader range of applications (already commer-
cially available through Unitika Ltd., Japan), and open a new dimension for plastics and
composites.

REFERENCES
1. R. Chandra, R. Rustgi, Progress in Poly. Sci. 23, 7, 1273 (1998).
2. M. Okamoto, Polymer/Clay Nanocomposites, “Encyclopedia of Nanoscience and Nanotechnology”
(H. S. Nalwa, Ed.), Vol. 8, p. 1. American Scientific Publishers, Stevenson Ranch, CA, 2004.
3. M. Alexander, P. Dubois, Mater. Sci. Eng. Res. 28, 1 (2000).
4. E. P. Giannelis, Organomet. Chem. 12, 675 (1998).
5. K. Yano, A. Usuki, A.Okada, T. Kurauchi, and O. Kamigaito, J. Polym. Sci. A: Polym. Chem. 31, 2493 (1993).
6. J. W. Gilman, Appl. Clay Sci. 15, 31 (1999).
7. S. Sinha Ray, K. Yamada, M. Okamoto, and K. Ueda, Nano Letts. 2, 1093 (2002).
8. R. A. Vaia, K. D. Jandt, E. J Kramer, and E. P. Giannelis, Macromolecules 28, 8080 (1995).
9. J.-S. Chen, M. D. Poliks, C. K. Ober, Y. Zhang, U. Wiesner, and E. P. Giannelis, Polymer 43, 4895 (2002).
10. A. Blumstein, J. Polym. Sci. Part A: Polym. Chem. 3, 2653 (1965).
11. B. K. G. Theng, “Formation and Properties of Clay-Polymer Complexes.” Elsevier, Amsterdam, 1979.
12. A. Usuki, Y. Kojima, A. Okada, Y. Fukushima, T. Kurauchi, and O. Kamigaito, J. Mater. Res. 8, 1174 (1993).
13. A. Usuki, Y. Kojima, A. Okada, Y. Fukushima, T. Kurauchi, and O. Kamigaito, J. Mater. Res. 8, 1185 (1993).
14. National Lead Co., U.S. Patent No. 2531396 (1950).
15. Union Oil Co., U.S. Patent No. 3084117 (1963).
44 Biodegradable Polymer/Layered Silicate Nanocomposites A Review

16. Unitika Ltd., Japanese Kokai Patent Application No. 109998 (1976).
17. Y. Fukushima and S. Inagaki, J. Incl. Phen. 5, 473 (1987).
18. P. Aranda and E. Ruiz-Hitzky, Chem. Mater. 4, 1395 (1992).
19. R. Krishnamoorti, R. A. Vaia, and E. P. Giannelis, Chem. Mater. 8, 1728 (1996).
20. R. A. Vaia, R. K. Teukolsky, and E. P. Giannelis, Chem. Mater. 6, 1017 (1994).
21. D. J. Greenland, J. Colloid. Sci. 18, 647 (1963).
22. J. Billingham and C. Breen Yarwood, J. Vibr. Spectrosc. 14, 19 (1997).
23. C. O. Oriakhi, I. V. Farr, and M. M. Lerner, Clay and Clay Minerals 45, 194 (1997).
24. P. H. Nam, P. Maiti, M. Okamoto, T. Kotaka, N. Hasegawa, and A. Usuki, Polymer 42, 9633 (2001).
25. R. A. Vaia, H. Ishii, and E. P. Giannelis, Chem. Mater. 5, 1694 (1993).
26. E. P. Giannelis, R. Krishnamoorti, and E. Manias, Adv. Polym. Sci. 138, 107 (1999).
27. M. Okamoto, S. Morita, Y. H. Kim, T. Kotaka, and H. Tateyama, Polymer 42, 1201 (2001).
28. R. S. Sinha, P. Maiti, M. Okamoto, K. Yamada, and K. Ueda, Macromolecules 35, 3104 (2002).
29. D. K. Yang and D. B. Zax, J. Chem. Phys. 110, 5325 (1991).
30. D. L. VanderHart, A. Asano, and J. W. Gilman, Chem. Mater. 13, 3796 (2001).
31. P. B. Messersmith and E. P. Giannelis, Chem. Mater. 5, 1064 (1993).
32. P. Messersmith and E. P. Giannelis, J. Polym. Sci., Polym. Chem. 33, 1047 (1995).
33. N. Pantoustier, B. Lepoittevin, M. Alexandre, D. Kubies, C. Calberg, R. Jerome, and P. Dubois, Poly. Eng. Sci.
42, 1928 (2002).
34. B. Lepoittevin, N. Pantoustier, M. Devalckenaere, M. Alexandre, D. Kubies, C. Calberg, R. Jerome, and
P. Dubois, Macromolecules 35, 8385 (2002).
35. B. Lepoittevin, M. Devalckenaere, N. Pantoustier, M. Alexandre, D. Kubies, C. Calberg, R. Jerome, and
P. Dubois, Polymer 43, 1111 (2002).
36. K. E. Strawhecker and E. Manias, Chem. Mater. 12, 2943 (2000).
37. H. Matsuyama and J. F. Young, Chem. Mater. 11, 16 (1999).
38. S. Sinha Ray, P. Maiti, M. Okamoto, K. Yamada, and K. Ueda, Macromolecules 35, 3104 (2002).
39. N. Ogata, G. Jimenez, H. Kawai, and T. Ogihara, J. Polym. Sci. Part B: Polym. Phys. 35, 389 (1997).
40. S. Sinha Ray, K. Okamoto, K. Yamada, and M. Okamoto, Nano. Lett. 2, 423 (2002).
41. S. Sinha Ray, K. Yamada, A. Ogami, M. Okamoto, and K. Ueda, Macromol. Rapid Commun. 23, 943 (2002).
42. S. Sinha Ray, M. Okamoto, K. Yamada, and K. Ueda, Nano. Lett. 2, 1093 (2002).
43. P. Maiti, K. Yamada, M. Okamoto, K. Ueda, and K. Okamoto, Chem. Mater. 14, 4654 (2002).
44. M. Pluta, A. Caleski, M. Alexandre, M.-A. Paul, and P. Dubois, J. Appl. Polym. Sci. 52, 1497 (2002).
45. S. Sinha Ray, K. Yamada, M. Okamoto, A. Ogami, and K. Ueda, Chem. Mater. 15, 1456 (2003).
46. M.-A. Paul, M. Alexandre, P. Degee, C. Henrist, A. Rulmont, and P. Dubois, Macromol. Rapid Commun. 24,
561 (2003).
47. J.-H. Chang, Y. Uk-An, and G. S. Sur, J. Polym. Sci. Part B: Polym. Phys. 41, 94 (2003).
48. S. Sinha Ray, K. Yamada, M. Okamoto, and K. Ueda, Polymer 44, 857 (2003).
49. S. Sinha Ray, K. Okamoto, and M. Okamoto, J. Nanosci. Nanotech. 2, 171 (2002).
50. S. Sinha Ray, K. Okamoto, and M. Okamoto, Macromolecules 36, 2355 (2003).
51. K. Okamoto, S. Sinha Ray, and M. Okamoto, J. Polym. Sci. Part B: Polym. Phys. 41, 3160 (2003).
52. X. Kornmann, L. A. Berglund, J. Sterete, and E. P. Giannelis, Polym. Eng. Sci. 38, 1351 (1998).
53. H. J. Choi, J. H. Kim, and J. Kim, Macromol. Symp. 119, 149 (1997).
54. P. Maiti, C. A. Batt, and E. P. Giannelis, Polm. Mater. Sci. Eng. 88, 58 (2003).
55. S. R. Lee, H. M. Park, H. L. Lim, T. Kang, X. Li, W. J. Cho, and C. S. Ha, Polymer 43, 2495 (2002).
56. S. H. Park, H. J. Choi, S. T. Lim, T. K. Shin, and M. S. Jhon, Polymer 42, 5737 (2001).
57. R. K. Bharadwaj, A. R. Mehrabi, C. Hamilton, C. Trujillo, M. F. Murga, A. Chavira, and A. K. Thompson,
Polymer 43, 3699 (2002).
58. S. T. Lim, Y. H. Hyun, and H. J. Choi, Chem. Mater. 14, 1839 (2002).
59. H. M. Park, X. Li, C. Z. Jin, C. Y. Park, W. J. Cho, and C. S. Ha, Macromol. Mater. Eng. 8, 553 (2002).
60. H. M. Park, X. Li, C. Z. Jin, C. Y. Park, W. J. Cho, and C. S. Ha, J. Mater. Sci. 38, 909 (2003).
61. H. Uyama, M. Kuwabara, T. Tsujimoto, M. Nakono, A. Usuki, and S. Kobayashi, Chem. Mater. 15, 2492 (2003).
62. M. Darder, M. Colilla, and E. Ruiz-Hitzky, Chem. Mater. 15, 3774 (2003).
63. S. Sinha Ray and M. Okamoto, Macromol. Rapid Commun. 24, 815 (2003).
64. A. Usuki, Y. Kojima, M. Kawasumi, A. Okada, Y. Fukushima,T. Kurauchi, and O. Kamigaito, J. Mater. Res. 8,
1179 (1993).
65. T. Yasuda and E. Takiyama, U.S. Patent No. 5391644 (1995).
66. B. Lepoittevin, M. Devalckenaere, N. Pantoustier, M. Alexandre, D. Kubies, C. Calberg, R. Jerome, and
P. Dubois, Polymer 43, 4017 (2002).
67. Y. Kojima, A. Usuki, M. Kawasumi, A. Okada, T. Kurauchi, and O. Kamigaito, J. Polym. Sci. Part A: Polym.
Chem. 31, 983 (1993).
68. S. T. Lim, Y. H. Hyun, and H. J. Choi, Chem. Mater. 14, 1839 (2002).
69. J. T. Yoon, W. H. Jo, M. S. Lee, and M. B. Ko, Polymer 42, 329 (2001).
70. K. Yano, A. Usuki, A. Okada, T. Kurauchi, and O. Kamigaito, J. Polym. Sci. Part A: Polym. Chem. 31, 2493
(1993).
71. K. Yano, A. Usuki, and A. Okada, J. Polym. Sci. Part A: Polym. Chem. 35, 2289 (1997).
72. S. Sinha Ray, K. Yamada, M. Okamoto, A. Ogami, and K. Ueda, Polymer 44, 6631 (2003).
73. L. Nielsen, J. Macromol. Sci. Chem. A1(5), 929 (1967).
Biodegradable Polymer/Layered Silicate Nanocomposites A Review 45

74. J. A. Tetto, D. M. Steeves, E. A. Welsh, and B. E. Powell, ANTEC ’99 1628 (1999).
75. J. W. Liu, Q. Zhao, and C. X. Wan, Space Medicine and Medical Eng. 14, 308 (2001).
76. S. Sinha Ray, K. Yamada, M. Okamoto, and K. Ueda, Macromol. Mater. Eng. 288, 936 (2003).
77. J. Y. Nam, S. Sinha Ray, and M. Okamoto, Macromolecules 36, 7126 (2003).
78. I. T. Okamoto, Polymer 36, 2736 (1995).
79. P. Maiti, P. H. Nam, M. Okamoto, A. Usuki, and N. Hasegawa, Macromolecules 35, 2042 (2003).
80. P. Maiti, P. H. Nam, M. Okamoto, A. Usuki, and N. Hasegawa, Polym. Eng. Sci. 42, 1864 (2002).
81. W. Hoogsteen, A. R. Postema, A. J. Pennings, G. TenBrinke, and P. Zugenmaier, Macromolecules 23, 634 (1990).
82. R. Krishnamoorti and E. P. Giannelis, Macromolecules 30, 4097 (1997).
83. J. Ren, A. S. Silva, and R. Krishnamoorti, Macromolecules 33, 3739 (2000).
84. A. Akelah, in “Polymers and Other Advanced Materials: Emerging Technologies and Business Opportunities”
(P. N. Prasad, J. E. Mark, and F. J. Ting, Eds.), p. 625. Plenum, New York, 1995.
85. M. Okamoto, H. Taguchi, H. Sato, T. Kotaka, and H. Tatayama, Langmuir 16, 4055 (2000).
86. M. L. Williams, R. F. Landel, and J. D. Ferry, J. Amer. Chem. Soc. 77, 3701 (1955).
87. P. H. Nam, Master’s Thesis, Toyota Technological Institute (2001).
88. W. P. Cox and E. H. Merz, J. Polym. Sci. 28, 619 (1958).
89. M. Okamoto, P. H. Nam, P. Maiti, T. Kotaka, N. Hasegawa, and A. Usuki, Nano. Lett. 1, 295 (2001).
90. T. Kotaka, A. Kojima, and M. Okamoto, Rheol. Acta 36, 646 (1997).
91. Y. Fujimoto, S. Sinha Ray, K. Yamada, A. Ogami, M. Okamoto, and K. Ueda, Macromol. Rapid Commun. 24,
457 (2003).
92. M. Okamoto, P. H. Nam, P. Maiti, T. Kotaka, T. Nakayama, M. Takada, M. Ohshima, A. Usuki, N. Hasegawa,
and H. Okamoto, Nano. Lett. 1, 503 (2001).
93. P. H. Nam, P. Maiti, T. Kotaka, M. Okamoto, T. Nakayama, M. Takada, M. Ohshima, A. Usuki, N. Hasegawa,
and H. Okamoto, Polym. Eng. Sci. 42, 1907 (2002).
94. J. M. Brown, D. B. Curliss, and R. A. Vaia, “Proc. of PMSE Spring Meeting,” San Francisco, CA, 2000, p. 278.
95. L. J. Gibson and M. F. Ashby, Eds., “Cellular Solids,” Pergamon Press, New York, 1988, p. 8.

Вам также может понравиться