Download as pdf or txt
Download as pdf or txt
You are on page 1of 71

United Arab Emirates University

Scholarworks@UAEU

Theses Electronic Theses and Dissertations

6-2020

POLY (LACTIC ACID)/NANOCLAY NANOCOMPOSITES FOR


PACKAGING APPLICATIONS
Maissa Akram Adi

Follow this and additional works at: https://scholarworks.uaeu.ac.ae/all_theses

Part of the Chemical Engineering Commons

Recommended Citation
Akram Adi, Maissa, "POLY (LACTIC ACID)/NANOCLAY NANOCOMPOSITES FOR PACKAGING
APPLICATIONS" (2020). Theses. 815.
https://scholarworks.uaeu.ac.ae/all_theses/815

This Thesis is brought to you for free and open access by the Electronic Theses and Dissertations at
Scholarworks@UAEU. It has been accepted for inclusion in Theses by an authorized administrator of
Scholarworks@UAEU. For more information, please contact mariam_aljaberi@uaeu.ac.ae.
ii

Declaration of Original Work

I, Maissa Akram Adi, the undersigned, a graduate student at the United Arab Emirates
University (UAEU), and the author of this thesis entitled “Poly (Lactic Acid)/Nanoclay
Nanocomposites for Packaging Applications”, hereby, solemnly declare that this thesis
is my own original research work that has been done and prepared by me under the
supervision of Dr. Muhammad Iqbal, in the College of Engineering at UAEU. This
work has not previously been presented or published, or formed the basis for the award
of any academic degree, diploma or a similar title at this or any other university. Any
materials borrowed from other sources (whether published or unpublished) and relied
upon or included in my thesis have been properly cited and acknowledged in
accordance with appropriate academic conventions. I further declare that there is no
potential conflict of interest with respect to the research, data collection, authorship,
presentation and/or publication of this thesis.

Student’s Signature: Maissa Date: 09/July/2020


iii

Copyright © 2020 by Maissa Akram Adi


All Rights Reserved
iv

Advisory Committee

1) Advisor: Dr. Muhammad Zafar Iqbal


Title: Assistant Professor
Department of Chemical and Petroleum Engineering
College of Engineering

2) Co-advisor: Prof. Basim Abu-jdayil


Title: Professor
Department of Chemical and Petroleum Engineering
College of Engineering
v

Approval of the Master Thesis

This Master Thesis is approved by the following Examining Committee Members:

1) Advisor (Committee Chair): Dr. Muhammad Zafar Iqbal


Title: Assistant Professor
Department of Chemical and Petroleum Engineering
College of Engineering

Signature ____________________________ Date 10/7/2020


____________

2) Member: Nayef Ghasem


Title: Professor
Department of Chemical and Petroleum Engineering
College of Engineering

Signature ____________________________ Date _____________

3) Member (External Examiner): Matthew liberatore


Title: Professor
Department of Chemical Engineering
Institution: University of Toledo, USA

Signature ____________________________ Date __July 10, 2020__


vi

This Master Thesis is accepted by:

Dean of the College of Engineering: Professor Sabah Alkass

Signature Date 15/7/2020

Dean of the College of Graduate Studies: Professor Ali Al-Marzouqi

Signature Date 15/7/2020

Copy ____ of ____


vii

Abstract

With increasing stringent environmental regulations and attemps to reduce dependence


on fossil fuels for polymer production, bio-based polymers are receiving appreciable
global attention. In this context, bio-based poly (lactic acid) (PLA) is one of the most
frequently used polymers in packaging application. However, despite of its high
versatility, this polymer is limited in packaging applications due to its low mechanical
properties. The two-dimensional nanofillers such as nanoclay and graphene are well
known for improving thermo-mechanical properties of polymers at low
concentrations. However, increasing higher filler concentration to improve barrier
properties of polymers reduces their mechanical properties which is a drawback. This
thesis investigates effects of two-dimensional nanoclays on thermal, mechanical, and
barrier properties of PLA/clay nanocomposites.

The nanocomposites of PLA with clay and chemically modified clay were synthesized
via solution blending method, and their thermal and mechanical properties were
investigated. Several characterization techniques were used to understand chemical
structure of nanoclays and thermo-physical properties of PLA/clay nanocomposites.
The x-ray diffraction indicated increased dispersion and exfoliation of modified clay
in PLA compared to unmodified clay. A significant decrease in percent crystallinity of
PLA-modified clay nanocomposites was observed. A higher Young’s modulus and
yield strength and reduced elongation at break revealed increased filler/polymer
interactions in PLA/modified clay nanocomposites. Moreover, results from gas
permeability revealed 75% reduction in Nitrogen gas permeability for PlA/Modified
nanocomposites at 5wt%. The results could be used as fundamental understanding of
clay/PLA interactions for thin film packaging applications. However, a detailed
analysis of extensional rheology is required to assess feasibility of these
nanocomposites for film blowing applications.

Keywords: PLA, Clay, Modified clay, Nanocomposites, Packaging, Dispersion.


‫‪viii‬‬

‫)‪Title and Abstract (in Arabic‬‬

‫بولي (حمض الالكتيك)‪/‬نانوكالي نانوكومبوزت لتطبيقات التعبئة والتغليف‬


‫الملخص‬

‫مع زيادة اللوائح البيئية الصارمة ومحاوالت تقليل االعتماد على األحافير إلنتاج البوليمر‪ ،‬تحظى‬
‫البوليمرات الحيوية باهتمام عالمي ملموس‪ .‬في هذا السياق‪ ،‬يعد بولي (حمض الالكتيك) الحيوي‬
‫القائم على البوليمرات أحد أكثر البوليمرات استخدا ًما في التعبئة والتغليف‪ .‬ومع ذلك‪ ،‬على الرغم‬
‫من تنوعها العالي‪ ،‬فإن هذا البوليمر محدود في تطبيقات التعبئة والتغليف بسبب خصائصه‬
‫الميكانيكية المنخفضة‪ .‬تُعرف حشوات النانو ثنائية األبعاد مثل طبقة النانو والجرافين بتحسين‬
‫الخصائص الميكانيكية الحرارية للبوليمرات بتركيزات منخفضة‪ .‬ومع ذلك‪ ،‬فإن زيادة تركيز‬
‫حشو أعلى لتحسين خصائص حاجز البوليمرات يقلل من خصائصها الميكانيكية وهو عيب‪ .‬تبحث‬
‫هذه الرسالة في تأثيرات الطبقات النانوية ثنائية األبعاد على الخصائص الحرارية والميكانيكية‬
‫والحاجزة للمركبات النانوية ‪/PLA‬الطين‪.‬‬

‫تم تصنيع المركبات النانوية من ‪ PLA‬مع الصلصال والصلصال المعدل كيميائيا ً عبر طريقة‬
‫مزج المحلول ‪ ،‬وتم فحص خصائصها الحرارية والميكانيكية‪ .‬تم استخدام العديد من تقنيات‬
‫التوصيف لفهم التركيب الكيميائي للطبقات النانوية والخواص الحرارية الفيزيائية للمركبات‬
‫النانوية ‪/PLA‬الطين‪ .‬أشار حيود األشعة السينية إلى زيادة تشتت وتقشير الطين المعدل في ‪PLA‬‬
‫مقارنة بالطين غير المعدل‪ .‬لوحظ انخفاض معنوي في النسبة المئوية لبلورة المركبات النانوية‬
‫المعدلة من ‪ .PLA‬أظهر معامل يونغ األعلى وقوة الخضوع وانخفاض االستطالة عند الكسر‬
‫زيادة تفاعالت الحشو ‪ /‬البوليمر في ‪ PLA /‬مركبات نانوية معدلة من الطين‪ .‬عالوة على ذلك‪،‬‬
‫تم عرض نتائج نفاذية الغازالتي كشفت عن انخفاض ‪ %75‬في نفاذية غاز النيتروجين عند نسبة‬
‫‪ %5‬من ‪/PLA‬الطين المعدل‪ .‬يمكن استخدام النتائج كفهم أساسي لتفاعالت الطين‪PLA/‬‬
‫لتطبيقات تغليف األغشية الرقيقة‪ .‬ومع ذلك‪ ،‬هناك حاجة إلى تحليل مفصل للريولوجيا الموسعة‬
‫لتقييم جدوى هذه المركبات النانوية لتطبيقات نفخ األفالم‪.‬‬

‫مفاهيم البحث الرئيسية‪ :‬حمض بولي الكتيك‪ ،‬طين‪ ،‬طين معدّل‪ ،‬مركبات نانوية‪ ،‬تعبئة وتغليف‪،‬‬
‫تشتت‪.‬‬
ix

Acknowledgements

First of all, I would like to thank God who enabled me to complete this work.

Then, I would like to express my deepest thanks and appreciation to my main

supervisor Dr. Muhammad Zafar Iqbal, Department of Chemical and Petroleum

Engineering (UAEU) for his guidance, support and knowledge through my two years

of research work. Moreover, My extended thanks and gratitude to my co-supervisor

Prof. Basim Abu-Jdayil for his continuous support and cooperation as usual.

Additionally, I would like to thank Dr. Hussein Awad, Eng. Essa Lwisa,

Emmanuel Galiwango for their assistance with, XRD and TGA characterizations.

Many thanks to my beloved parents, family and friends for their

encouragement and care.


x

Dedication

To my beloved parents and family


xi

Table of Contents

Title ............................................................................................................................... i
Declaration of Original Work ...................................................................................... ii
Advisory Committee ................................................................................................... iv
Approval of the Master Thesis ..................................................................................... v
Abstract ...................................................................................................................... vii
Title and Abstract (in Arabic) ................................................................................... viii
Acknowledgements ..................................................................................................... ix
Dedication .................................................................................................................... x
Table of Contents ........................................................................................................ xi
List of Tables ............................................................................................................ xiii
List of Figures ........................................................................................................... xiii
List of Abbreviations ................................................................................................ xiv
Chapter 1: Background .............................................................................................. 15
1.1 Overview ...................................................................................................... 15
1.2 Relevant Literature ...................................................................................... 17
1.2.1 Polymers in Packaging Applications .................................................. 17
1.2.2 Poly (lactic acid) (PLA) ...................................................................... 19
1.2.3 Polymer Nanocomposites ................................................................... 19
Chapter 2: PLA/Clay Nanocomposites ...................................................................... 29
2.1 Introduction ................................................................................................. 29
2.2 Experimental ................................................................................................ 34
2.2.1 Materials ............................................................................................. 34
2.2.2 Preparation of Nanocomposites .......................................................... 34
2.3 Characterization tools .................................................................................. 35
2.4 Results and Discussion ................................................................................ 36
2.4.1 Characterization of Nano-Clay ........................................................... 36
2.4.2 Characterization of PLA nanocomposites .......................................... 42
Chapter 3: Conclusions and Future Recommendation ............................................... 59
3.1 Conclusions ................................................................................................. 59
3.2 Future Recommendations ............................................................................ 60
References .................................................................................................................. 61
xii

List of Tables

Table 1: Different applications of PLA/composites................................................... 16


Table 2: Several biodegradable polymers with different applications ....................... 19
Table 3: DSC data of PLA, PLA/Clay and PLA/Modified clay
nanocomposites ............................................................................................ 47
Table 4: Decomposition temperature and overall weight loss of PLA
and PLA nanocomposites ............................................................................ 49
Table 5: Elongation at break of PLA, PLA/Clay and PLA/Modified clay
nanocomposites ............................................................................................ 54
xiii

List of Figures

Figure 1: Distribution of various polymers used for packaging industry .................. 18


Figure 2: Structure of 2:1 layered silicate clay nanoparticles .................................... 24
Figure 3: Two distinct approaches applied to the production of conventional
micro- and nano-composites ...................................................................... 26
Figure 4: PLA structure ............................................................................................. 31
Figure 5: FTIR spectra of nano clay particles ............................................................ 37
Figure 6: XRD of clay nanoparticles ......................................................................... 39
Figure 7: XRD of PLA/Modified clay nanocomposites ............................................ 39
Figure 8: XPS of clay and Modified clay nanoparticles. ........................................... 41
Figure 9: FTIR spectrum of PLA/Modified clay nanocomposites............................. 43
Figure 10: DSC curves of PLA, PLA/Clay and PLA/Modified Clay
nanocomposites at 7 wt% clay loadings .................................................... 45
Figure 11: The effect of PLA/Clay and PLA/Modified clay nanocomposites
on crystallinity ........................................................................................... 47
Figure 12: Thermal stability of PLA and PLA nanocomposites ................................ 49
Figure 13: Stress-strain curves of PLA, PLA/Clay, and PLA/Modified clay
composites ................................................................................................. 50
Figure 14: Young’s modulus (MPa) of PLA nanocomposites ................................... 52
Figure 15: Yield strength (MPa) of PLA nanocomposites......................................... 52
Figure 16: Elongation at break for PLA composites .................................................. 54
Figure 17: Nitrogen gas permeability of PLA/Clay nanocomposites ........................ 57
Figure 18: Nitrogen gas permeability of PLA/Modified clay
nanocomposites ......................................................................................... 58
xiv

List of Abbreviations

DSC Differential Scanning Colorimeter

∆Hcc Cold Crystallization Enthalpy

FTIR Fourier-Transform Infrared Spectroscopy

∆Hm Melting Enthalpy

PLA Poly Lactic Acid

Tcc Cold Crystallization Temperature

Tg Glass Transition Temperature

Tm Melting Temperature

TGA Thermo-Gravimetric Analysis

Xc Percentage of Crystallinity

XPS X-ray Photoelectron Spectroscopy

XRD X-ray Diffractometer


15

Chapter 1: Background

1.1 Overview

Bio-based poly (lactic acid) (PLA), a biodegradable polyester produced from

renewable sources, [1] contains a key-position within the showcase of biopolymers for

different applications such as biomedical, packaging, material strands [2].

PLA meets numerous requirements as a packaging thermoplastic and is

proposed as a product gum for common packaging applications. Owing to its

biodegradability, PLA can also be utilized in manufacturing bioplastics, valuables for

creating loose-fill packaging, compost sacks, food packaging, and disposable flatware.

In the form of fibers and non-woven materials, PLA has shown potential in upholstery,

disposable garments, overhangs, feminine hygiene items, and nappies. PLA is also

used as a hydrophobic block of amphiphilic block copolymers to create vesicle films

of polymersomes [3]. For a few years, normal foods purveyors such as Newman’s

Possess Organics and Wild Oats have been quietly utilizing a few PLA items, but the

material got its greatest boost when Wal-Mart, the world’s biggest retailer, declared

that they will offer a few deliver-in PLA containers.

Being an imperative aliphatic polyester and biodegradable at the same time,

PLA has broad applications in biomedical engineering such as suture, bone fixation

material, drug delivery microsphere, and tissue engineering [4, 5]. A major application

field for microspheres is controlled drug delivery whereas a protein or peptide is

homogeneously dispersed within PLA microspheres. The most commonly utilized

materials for these systems are polylactide (PLA) and its copolymers with glycolide

(PLGA). The adjuvant impact of biodegradable poly (D and L-lactic acid) granules is
16

competent of antigen release taking after intraperitoneal injection [3]. Another

common application of PLA is in bio sutures. In this application a surgeon imposes a

little radius of curvature but due to the even smaller suture diameter, the strains are

frequently small. Similarly, small stresses and strains are normal for delicate tissue

applications [6]. Table 1 provides a brief overview of several applications of PLA

composites.

Table 1: Different applications of PLA/composites

PLA/Composite Application
PLA/graphene oxide-polyethylene glycol [7] Tissue engineering
PLA/Poly capro lactone matrix reinforced with Bone regenerative implants
phosphate-based glass fibers [8]
PLA/chitosan [9] Bone tissue scaffolds
PLA/gelatin nanofibers [10] Tissue engineering
PLA/daunorubicin-nano TiO2 nanofibers [11] Drug delivery
PLA/ketoprofen nanofibers [12] Drug delivery
PLA/cork [13] Fused deposition modelling
PLA/Ibuprofen nanofibers [14] Drug delivery

PLA/hydroxyapatite nanofibers [15] Bone tissue regeneration

The bio-based PLA is one of the most frequently used polymers in medical

packaging application. However, despite of its high versatility, these polymers are

limited in widespread packaging applications due to poor mechanical properties. The

two-dimensional fillers such as nanoclay are well known for improving thermal and

mechanical properties of polymers at low concentrations. This will help in producing

economical nanocomposites for packaging and beverage applications with low filler

concentrations, and improved thermal and mechanical properties.


17

1.2 Relevant Literature

1.2.1 Polymers in Packaging Applications

Plastics industry overwhelmingly depends on the fossil-based polymers (~107

million tons in 2015) [16, 17] owing to their affordability and superior performance

compared to naturally occurring and bio-based-polymers. The packaging industry

shares a major portion (approximately 38%) of the globally produced plastics. Most

frequently used polymers in packaging application are polyethylene (PE),

polypropylene (PP), polystyrene (PS), polyvinyl chloride (PVC), and polyethylene

terephthalate (PET) [18-20]. Polyolefins such as high density PE (HDPE), low density

PE (LDPE) and linear low density PE (LLDPE) dominate the packaging industry

(Figure 1) [17]. The density of PE plays an important role in increasing its flexibility

for various applications. For example, high strength containers (milk and detergent

bottles) are produced from HDPE; flexible films and squeezable containers are

manufactured from LDPE; and LLDPE is used to produce grocery bags and cling wrap

films owing to its excellent flexibility and better mechanical properties [17, 21].

However, despite of their high versatility, a limiting property of polymeric materials

in packaging is their inherent permeability to gases and vapors, including oxygen,

carbon dioxide, and organic vapors [19, 20]. Nevertheless, several copolymers such as

poly (ethylene-co-vinyl alcohol) (EVA) have excellent oxygen barrier properties

required for food packaging applications. However, copolymers are typically

coextruded as an internal layer or blended with other polymers.


18

Others,
PUR, 0.20%
0.40%
LDPE/LL PET, 23%
DPE,
30%
PVC, 2%

PS, 5%

HDPE, PP, 19%


21%

Figure 1: Distribution of various polymers used for packaging industry

Recently, bio-based polyolefins derived from biomass have exhibited

properties identical to those of petroleum-based polyolefins [17, 22, 23]; aiming at

replacing conventional fossil-based polyolefins. However, bio-polyolefins are 30%

higher in cost compared to fossil-based polyolefins [17, 24], limiting their use in

economical packaging applications. In general, the problems associated with

biodegradable polymers are threefold: performance, processing, and cost. Although

these factors are somewhat interrelated, the problems due to “performance and

processing” are common to all biodegradable polymers despite of their origin [25]. In

particular, brittleness, low heat distortion temperature, high gas and vapor

permeability, poor resistance to prolonged processing operations have limited their

applications. The application of nanotechnology to these polymers may open new

possibilities for improving not only the properties but also the cost-performance

efficiency [26].
19

1.2.2 Poly (lactic acid) (PLA)

PLA is a highly versatile biodegradable aliphatic polyester derived from 100%

renewable resources such as corn and sugar beets. PLA offers great promises in a wide

range of commodity applications. Despite of its excellent balance of properties, the

commercial viability has historically been limited by high production costs (greater

than $2/lb). Until now PLA has enjoyed little success in replacing petroleum-based

plastics in commodity applications, with initial uses limited to biomedical applications

such as the sutures [27]. Among the biodegradable class of polymers, PLA is the most

anticipated polymer for large scale packaging applications. Table 2 shows some

applications of biodegradable polymers [28].

Table 2: Several biodegradable polymers with different applications

Product Composition Applications

Polynat Rye flower (80%) Disposable items, flower containers


Ecofoam Starch Wrapping plastics
Biopol PHB/PHV Razors, bottles
Eco-PLA PLA Sanitary products, sport clothes,
conditioning and packaging
Ecoflex Co-polyester Agricultural films
BAK 1095 Polyester amide Disposable items, flower containers

1.2.3 Polymer Nanocomposites

Polymer nanocomposites are the class of materials in which nanoscale

particulates such as layered clays or spherical inorganic minerals are dispersed within

polymer matrices. Compared to pure polymers, polymer nanocomposites exhibit


20

markedly improved functional properties depending on the type and concentration of

nanoparticles [29-31]. Significant scientific and technological interest has been

focused on polymer nanocomposites (PNCs) over the last few decades [32, 33]. These

interests are further boosted with discoveries such as carbon nanotubes, graphene, and

metal-carbon complexes such as MXenes.

In general, most of the early and some of the recent efforts focused on

understanding the underlying synthetic and physical chemistry [34] in conjunction

with the issues of mechanical, thermal and physical properties [35] of polymer

nanocomposites. Only recently more consideration has been taken to understand and

exploit the physics of polymers in nanocomposites. This was persuaded by a

developing acknowledgment to move beyond defining polymers with nanofillers, and

towards genuinely designed, planned, and utilitarian nanocomposites. In this way,

broader commercial utilization, the current fragmented understanding of the elemental

angles of the structure, property and processing relationship of these materials must be

addressed.

Tremendous interests in PNCs were proclaimed by the advancement of nylon-

6 based clay nanocomposites developed by the Toyota research group where an

uncommon set of mechanical, thermal, and physical properties were accomplished by

exfoliating clay layers with a significant division of the polymer chains ionically

connected by utilizing clay bound initiator during ring opening polymerization [36].

The early consideration was on the exfoliation of clay layers and its effects on property

improvement; as of late, it has ended up apparently that the nature of polymer-clay

interactions and high density of surface grafting changes crystalline polymorphs of


21

nylon-6. The altered polymorphs can be related to the changes within the mechanical

and thermal properties of those nanocomposites [37].

Exceptional issues that might significantly affect the properties and handling

of PNCs include changes in glass transition temperatures (Tg), as well as the crystal

polymorphs and propensities [38]. Favorable interactions between polymer and

nanoparticle usually result in increased Tg of the polymer. It was also found that,

favorable interaction between polymer and filler results in a higher zero shear

viscosity, an earlier shear‐thinning transition, and a stronger shear‐thinning behavior

[39]. Also, the favorable polymer filler interaction restricts the mobility of the polymer

chains, which results in the increase of the storage modulus and complex viscosity

[40].

1.2.4 Clay Nano-Composites

Nanomaterials are the materials with nanoscopic dimensions which not only have

potential technological applications in areas such as chemistry, physics, optics,

materials science, biomedical sciences, mechanical devices [41], drug delivery [42],

bioencapsulation but also are of fundamental interest in that the properties of a material

can transition between bulk and molecular scales [41]. Generally, different nanofillers

exist in numerous sizes and shapes, and can be basically classified into three

categories depending on the dimensionality: Zero, one-, two-, and three-dimensional

fillers. Spherical nanoparticles are considered zero-dimensional (0D), fibrous

nanoparticles such as nanotubes or carbon nanofiber are one-dimensional (1D), plate-

like nanoparticles are considered as 2D fillers, and nanoparticles such as graphite are

considered 3D due to breadth of their structures. The plate-like nanofillers (2D) are

layered materials with a thickness on the order of 1 nm, but with an aspect ratio
22

following their two remaining dimensions of at least 25. The most well-known 2D

nanofillers graphene and nanoclays. Graphene and clay have a thickness below 100

nm and characterized by an aspect ratio of ~100. The most common 3D nanofillers are

graphite and metal oxides [43].

Among various nanoparticles, nanoclay, carbon nanotubes, and metal/ ceramic

nanoparticles are the most common nanofillers being studied to fabricate a variety of

high-performance polymer nanocomposites [44-46]. The most important nanoparticles

which are known to improve barrier performance of polymers against gases are the 2D

nanomaterials such as nanoclay and graphene [47]. The plate-like morphology of the

2D nanofillers provides a tortuous path to the penetrating gas molecules, increasing

their gas and vapor resistance. At the meantime, both of these nanofillers have also

been reported to improve mechanical properties of the polymers. Compared to

different nanoparticles, nanoclay is especially attractive as a reinforcement because it

is environmentally friendly, readily available in large quantities at relatively low cost

[48], and its intercalation chemistry is well studied and understood. The clay-based

nanofillers are commonly alluded to aluminosilicate (also called layered silicates)

based nanofillers.

Layered silicates such as montmorillonite, hectorite and saponite are widely

used for preparing polymer composites to improve functional properties of polymer.

The most pertinent factor in clay/polymer nanocomposites is controlling the effective

dispersion of the clay by keeping up the dispersion of clay layers at nanometric scale

instead of micrometric scale [49, 50]. This is attributed to high surface area of the clay

platelets scattered inside polymer, offering huge surface for polymer adsorption and

inter-gallery intercalation of clay [51].


23

Layered silicate clay nanocomposites speak to a modern lesson of crossover

materials, making them interesting in academic and industrial domains due to their

excellent thermal, obstruction, thermomechanical and fire-resistance properties at low

nanofiller content [52]. These enhancements depend on the scattering level of layered

silicates (either intercalated or exfoliated structures) inside the polymeric lattice

(known as dispersion) [53]. Within the domain of polymeric nanocomposites, 2:1

layered silicates (also called 2:1 phyllosilicates) such as montmorillonite and saponite

have been broadly explored as nanofillers [54]. In terms of structure, silicate layers

appear as a superposition of two tetrahedrally facilitated silicon atoms on the other

hand isolated by one octahedrally facilitated magnesium hydroxide moieties. Whether

isomorphic substitution occur between trivalent Al3+ particles and divalent Mg2+

particles inside the silicate layer, structure of montmorillonite is obtained (with a layer

thickness near to 1 nm) [55, 56] (Figure 2).

The electronic insufficiency display at the silicate layer (negative charges) is

counteracted by Na+, K+, Li+ and Ca2+ particles residing inside silicate layers. The

nonsufficient electrostatic forces empower joining of polar particles inside these

displays as well as critical swelling capacity for the layered silicates in aqueous media.

Compared to distinctive sorts of layered silicate such as vermiculites, illites, kaolinite,

and smectite layered silicates exhibited superior characteristics, specifically

montmorillonite. Agreeing to their most elevated swelling capacity in aqueous media,

montmorillonites are heavily studied and utilized in manufacturing polymer

nanocomposites [57].

It is difficult to achieve the complete exfoliation of nanosheets, but rather a

mixture between intercalated and exfoliated structures is obtained. These complex


24

morphologies are so-called cluttered morphology or deliberate exfoliated morphology.

Hence, an organic modification of layered silicates is required with, e.g. cationic trade

with natural cations [53, 58] to increase effective polymer-clay interactions.

Figure 2: Structure of 2:1 layered silicate clay nanoparticles

The chemical modification helps improving dispersion of clay inside polymers.

The modification is accomplished by substituting cations (Na+, K+, and Li+) on the

surface of clay layers with cationic surfactants counting primary, secondary, tertiary,

and quaternary alkylammonium or alkylphosphonium cations. These surfactants alter

the nature of the clay surface from hydrophilic to hydrophobic as well as expanding

interlayer dividing of clay and in a similar manner, they advance more fondness and

space for the polymers [59, 60]. In this respect, the inorganic cations found in the

interior galleries (Na+, Ca2+, etc.) are for the most part, traded by natural ammonium,

or sulfonium and/or phosphonium cations bearing at slightest one long alkyl chain,

beside other substituted groups. From the literature, ammonium alkyls appear to be an
25

excellent organic modifying operators. These natural cations play a vital part to

diminish the surface vitality between layers as well as to strengthen the fondness

toward polymeric framework. Besides, use of natural cations with long alkyl chains

increments, the interlayer separation is prompted in silicate layers. This allows a better

diffusion of polymeric chains within silicate galleries [58].

Clay/polymer nanocomposites are a commonplace case of nanotechnology.

This class of nanocomposites employ smectite type clays, such as hectorite,

montmorillonite, and manufactured mica, as fillers to improve polymers properties.

Smectite type clays have a layered structure. Each layer is constructed from

tetrahedrally facilitated Si molecules melded into an edge shared octahedral plane of

either Al(OH)3 or Mg(OH)2. According to the nature of the holding between these

atoms, layers ought to display amazing mechanical properties parallel to the layer

course [61]. Hundreds or thousands of these layers are stacked together with week van

der Waals strengths to create a clay molecule. With such a setup, it is conceivable to

tailor clays into various different structures in a polymer [61]. The major intrigued in

utilizing clays for polymer enhancement was to break down clay particle aggregates

into individual particles to create micro-sized filler reinforced polymers (Figure 3). It

can be envisioned that the excellent mechanical properties of each layer in clay

particles cannot work successfully in such a framework. The week interlayer bonding

may act as harm initiation sites in applications. It is common to utilize high clay

stacking to achieve satisfactory enhancement of the modulus, while strength and

toughness of the polymer are decreased [62].

The guideline utilized in clay/polymer nanocomposites is to separate not only

clay aggregates but moreover individual silicate layers in a polymer, as outlined


26

schematically in Figure 2. By achieving this, the great mechanical properties of the

individual clay layers can work viably, whereas the number of fortifying components

increase dramatically since each clay molecule contains hundreds or thousands of

layers. As a result, several properties can be essentially moved forward with a low

level of filler loading, usually less than 5 wt% [62].

Figure 3: Two distinct approaches applied to the production of conventional micro-


and nano-composites

The primary commercial application of these materials was clay/nylon-6

nanocomposites as timing belt covers for Toyota cars, in collaboration with Ube in

1991 after which, Unitika presented nylon-6 nanocomposites for engine covers on

Mitsubishi’s GDI engines [63]. In 2001, General Motors and Basell declared the

application of clay/polyolefin nanocomposites as a step assistant component for GMC

Safari and Chevrolet Astro vans. This was taken after by the application of these

nanocomposites in the entryways of Chevrolet Impalas [64]. Recently, Noble

polymers, united states, have created clay/polypropylene nanocomposites for auxiliary

situate backs within the Honda Acura, whereas Ube is developing clay/nylon-12
27

nanocomposites for automotive fuel lines and fuel framework components.

Additionally, clay/polymer nanocomposites have been utilized to make strides

obstruction resistance in refreshment applications. Alcoa CSI has connected multilayer

clay/polymer nanocomposites as obstruction liner materials for enclosure applications

[65]. Honeywell has made commercial clay/nylon-6 nanocomposite items for drink

bundling applications [66]. More recently, Mitsubishi Gas Chemical and Nanocor

have developed Nylon-MXD6 nanocomposites for multilayered polyethylene

terephthalate (PET) bottle applications [67].

Mechanical properties of polymer clay nanocomposites are profoundly related

to their microstructure which in turn is straightforwardly related to extent of exfoliation

and dispersion of clay platelets within the polymer. The dispersion of clay platelets

within melt polymer depends on thermal diffusion of polymer chains within the

exhibitions and mechanical shearing activity [68] . Most likely, good exfoliation of

clay platelets in a polymer matrix leads to upgraded Young’s modulus, storage

modulus, and tensile strength, but significantly diminished tensile ductility and impact

strength compared to neat polymer [69, 70]. For example, Compared to Nylon 6,

clay/nylon 6 nanocomposites with ~4 wt% clay loading achieved 47% and 67%

enhancement in tensile yield strength and Young’s modulus respectively [62]. The

tensile modulus as well as the tensile strength of compatibilized polyethylene organo

MMT nanocomposites at 3wt% MMT contents, were enhanced by 23% and 14%

respectively in comparison with neat polyethylene [48]. Also, supercritical CO2

processed nanoclay (modified MMT) dispersed in PET increased tensile strength

(10.7%) and the Young's modulus (21.4%) compared to neat PET at only 1 wt% clay

concentration. Further improvement was observed when MMT loading increased to 3

wt% [71]. The tensile strength of PLA nanocomposites with Na+ modified MMT
28

decreased by 25% compared with the neat PLA. The reduced tensile strength indicates

poor dispersion of Na+ modified MMT in PLA. Meanwhile, the tensile strength of

PLA/MMT nanocomposites films was still comparable to that of commonly used

plastic films such as HDPE (22-31 MPA), PP (31-38 MPa, and PS (45-83 MPa) [72].

1.3 Thesis Objectives

The most frequently used polymers in packaging application are petroleum-

based polyethylene terephthalate (PET). The objective of this thesis is to investigate

the physical and the mechanical properties of bio-based poly (lactic acid) (PLA)/clay

nanocomposites using 2D nanoclay fillers (MMT clay and modified clay). This thesis

further helps in manufacturing economical nanocomposites for packaging and

beverage applications with low filler concentrations and improved mechanical

properties. Various characterization techniques such as morphological, thermal and

mechanical were used to examine the effect of fillers on resultant properties.

This thesis consists of three chapters as follows: Chapter 1 contains literature

review of polymers in packaging applications, PLA, and polymer nanocomposites.

Chapter 2 covers literature review about PLA/clay nanocomposites. It also discusses

the experimental technique used to prepare PLA/Clay nanocomposites and the

characterization of nanoclay particles (clay, modified clay) as well as the

characterization of PLA nanocomposites. Lastly, conclusions and future

recommendations are included in Chapter 3.


29

Chapter 2: PLA/Clay Nanocomposites

2.1 Introduction

Approximately 99% of the plastics used nowadays are petroleum-based, and

the packaging industry alone expands over 38% of these plastics. Despite of the fact

that polyolefins and their subsidiaries offer great set of properties for packaging

applications, they are inferred from non-renewable resources. Petroleum costs are

unsteady and unpredictable worldwide, and the petroleum supply will inevitably get

to be exhausted [73]. Similarly, most of these polymers have saturated (carbon-carbon)

bonds in their backbones resisting biodegradation. Consequently, white contamination

(emerging from the proliferation of plastic packaging materials) and landfilling issues

are putting an expanding strain on the society and environment. Biodegradable

polymers can provide a key turning point as renewable plastics within the course

toward green packaging [17].

Biodegradable polymers incorporate common polymers (synthesized from

living species such as plants, creatures and microorganisms, and extricated as

polymers) and polymers synthesized from bio-based monomers industry [21, 74].

Natural polymers inferred from biomass (barring microorganisms) are plenteous and

renewable in nature [75]. Natural polymers are mostly biodegradable (capable of being

converted/decomposed into simple/ basic atoms by microorganisms) and hence, they

reduce the strain on landfills. These polymers exist within the frame of polysaccharides

(e.g. cellulose, starches and their derivatives, seaweed extricates such as carrageenan

and alginates, chitosan and pectin), proteins (whey protein, gelatin, corn zein, soy

protein, collagen and wheat gluten) and lipids. Cellulose is the most copious

biopolymer but it is non-thermoplastic and in this way is troublesome to machine


30

handle [76]. Interestingly, the viscose handle depends on hazardous materials, such as

carbon disulfide. In expansion, cellophanes are hydrophilic and lose their mechanical

and barrier properties upon presentation to water. Subsequently, cellophanes are

utilized for packaging applications as they have been generally supplanted by PP

movies. Other cellulose-based polymers such as cellulose nitrates and cellulose acetic

acid derivations were popular earlier to the rise of polyolefins within the 1950s [21].

These polymers were utilized for extraordinary applications especially within the

packaging industry. Currently, starch and its blends are utilized in compostable bags

and as free-fills in packaging disseminations as a substitution for polystyrene due to

their non-adhesive characteristics.

In general, the destitute water resistance and challenges experienced within the

machine processability of the key normal polymers restricted their product scale

applications [77]. Critical advances have been made within the field of bio-based

biodegradable nanocomposites for food packaging applications. Poly (lactic acid)

(PLA)- a well-known bio-based polymer- has pulled in gigantic consideration.

PLA is commercially synthesized through the ring opening polymerization

(ROP) of lactides (Figure 4) [78]. A two-step strategy for creating high molecular

weight PLA (with normal molecular weight (MW) ranging from 100 to 500 kg mol−1)

is used. In the first step, a low MW PLA is created by polymerizing lactides in liquid

state, followed by solid state polymerization to achieve high MW PLA. Naturally,

lactides transcendently exist as L- and D- isomers. Consequently, the PLA synthesized

via ROP consists of transcendently poly(L-lactic corrosive) (PLLA) with a few mol%

of the D-lactide isomer [79].


31

Figure 4: PLA structure. [reported from [78]]

The thermal properties of PLA such as Tg and Tm rely on the D-contents in

PLA. For example, Tg and Tm of PLA with 94% L-lactide contents are 66 and 141°C,

respectively. Similarly, expanding the percent composition of the D-isomer decreases

crystallinity of PLA. High D-lactide content (more than 10%) renders PLA into

completely an amorphous polymer. In addition, relative humidity is another factor that

influences Tg of PLA [80]. Due to its semi-crystalline nature, PLA has reasonably good

oxygen barrier properties. For instance, PLA has 10 times better oxygen barrier

characteristics than petroleum-based PP. However, PLA may have poor barrier for

water vapor. PLA exhibits 1/10th barrier compared to PP, independent of the relative

humidity [81].

Moreover, mechanical properties of PLA also depend on its D- & L-contents

[21]. The 100% L-lactide PLA is brittle. The presence of 2–10 mol% D-lactide isomers

renders PLA for packaging applications. PLA with 94% L-lactide composition is a

stiff fabric having a Young’s modulus of 2.5 GPa, which is 2 times higher than PP (1.3

GPa) and PHA (1.5 GPa). The tensile strength of PLA (74–82 MPa) is more than the
32

bio-based poly hydroxyalkanoates (PHAs) (15–40 MPa). In conclusion, PLA with

94% L-lactide has reasonable mechanical properties for packaging applications [82].

PLA does not viably block UV light [83] For example, up to 85% of UV-B and

95% of UV-A is transmitted by PLA films, which limits the utilize of transparent PLA

for applications where UV exposure can harm a packaged product. Hence, owing to

low oxygen permeability and higher UV transmission, PLA containers are not suitable

for juices unless extra measures are taken to block UV light and enhance their barrier

properties, and consequently hold the taste and appearance while extending item’s

shelf-life. Nevertheless, PLA is transparent within the visible region, which may be an

alluring highlight for the production of clear packages with effortlessly seeable items

[83].

Due to the exhaustion of petroleum assets, PLA is presently showing increasing

interests as an important bio-sourced polymer elective in long terms for several

applications. Nevertheless, PLA endures from a few deficiencies such as low thermal

resistance, heat distortion temperature and rate of crystallization. On the other hand, a

few other particular properties are required by distinctive applications such as fire

retardancy, antistatic to conductive electrical characteristics, anti-UV, antibacterial or

barrier properties, etc.

There are numerous nanofillers (one dimensional such as nanotubes and

nanofibers , two dimensional like graphene and clay and three dimensional like

graphite), that have been considered, with palatable accomplishments, within the

design of PLA nanocomposites [58]. The incorporation of clay nanoparticles has

shown to improve performance in thermal stability, stiffness, and processability of

polymers. The organically modified clays have been broadly considered. The coming
33

about PLA/clay nanocomposites are interesting in the packaging industry since of their

prevalent mechanical, thermal, gas-barrier, and optical properties indeed at low clay

content. [84].

The use of nanoclays to reinforce PLA has been reported in literature. For

example, Ray et al. [85-90] demonstrated intercalation of nanoclays in PLA and

resulted properties. The Young’s modulus increased by 25% with increasing clay

contents whereas a noteworthy decrease in oxygen penetrability was observed.

Fukushima et al. [91] synthesized nanocomposites by with 5 and 7 wt% of clays using

a commercial organic modified clay (Cloisite 30B), an organically modified

magnesium sodium fluoro-hectorite, and unmodified sepiolite. They observed good

dispersion of clay in PLA with significant enhancements in thermal and thermo-

mechanical properties of all nanocomposites. Specifically, 30B clay is a naturally

altered montmorillonite (OMMT) where the natural modifier has two hydroxyl groups

The improved performance of PLA nanocomposites was attributed to the interactions

between hydroxyl groups in clay and the carboxylic groups in PLA [92-94]. A number

of reports have shown that the 30B clay intercalates or exfoliates well in PLA [92, 95,

96]. In addition, PLA/ 30 B clay frameworks were studied for degradation behavior

[97], controlled drug release [98], and water vapor transmission characteristic [72].

Generally, elongation at break decreases with increasing filler contents in

PNCs. However, Jollands and Gupta [99] reported reported increased elongation at

break from 2.3 to 3.2% for the nanocomposite compared to neat PLA. Gamez Perez

et al. [100] observed elongation at break rising from 4.0% for neat PLA to 5.7% and

11% for nanocomposites at 0.5 and 2.5 wt% of OMMT, respectively. Wang et al. [101]

reported ductile PLA/ modified MMT nanocomposites where elongation at break


34

increased from 5.4 to 7.9% whereas tensile modulus also progressed from 1.65 to 1.93

GPa at 1 wt% clay loading compared to neat PLA. More interestingly, elongation at

break improved by 26.5 times as reported by Li et al. [102] for naturally modified

rectorite (OREC)/ PLA melt processed nanocomposites.

This thesis focuses on understanding how clay and modified clay nanoparticles

influence the physical, thermal, crystallization and mechanical behavior of PLA/Clay

and PLA/Modified clay nanocomposites.

2.2 Experimental

2.2.1 Materials

Nanoclay hydrophilic bentonite with a bulk density 600-1100 g/cm3 (682659,

Sigma Aldrich), Nanoclay-surface modified contains 35-45 wt% dimethyl dialkyl

(C14-C18) amine with a bulk density 200-500 g/cm3 (682624, Sigma Aldrich), Semi

crystalline poly (lactic acid) pellets with an L-lactide:D-lactide ratio from 24:1 to 32:1

and molecular weight of 2.41X105 g/ mol (PLA 4032D, Zhejiang Zhongfu Industrial

Limited, Zhejiang, China) And tetrahydrofuran (anhydrous, ≥99.9%, Sigma Aldrich)

were used as received.

2.2.2 Preparation of Nanocomposites

The PLA/Nanocomposites were prepared with different clay and modified clay

concentrations (0, 1, 3, 5, 7 wt%) using solvent blending technique. Approximately,

2.5 g of PLA and appropriate amounts of clay nanoparticles were sonicated in THF

solvent at room temperature for 270 minutes after which the temperature of the mixure

was measured to be found as 60°C. The mixture was refluxed at 60°C under constant
35

magnetic stirring at 210 RPM for 2 hours. The hot mixture was dropped into deionized

water resulting in coagulation of the nanocomposite suspension followed by drying

under fumehood for 24 hours.

The dried samples were hot pressed using a Carver hydraulic press at 175°C

for 15 minutes to produce thin sheets (~9 cm diameter, ~0.1 mm thick) followed by

cooling for ~3 minutes. Finally, samples for tensile test were cut using a manual cutter

(Pioneer Die-tecs) using ASTM D-638-V die into standard dumbbell-shaped

specimens. A micrometer (0.001 mm resolution) was used to measure the thickness of

each specimen.

2.3 Characterization tools

FT/IR-4700 Spectrometer with PRO ONE Single reflection J142A-ATR

accessory (Jasco, Japan) was used to collect the Fourier transform infrared spectrum

of clay, modified clay, and PLA nanocomposites over the spectral range of 400-5000

cm-1. A thin potassium bromide (KBr) pellet was used as background.

PANalytical-XPERT-3 (Philips, Netherland) was used to obtain x-ray

diffraction (XRD) spectrum using 45 kV and 40 mA working voltage and current,

respectively, and Cu-Kα radiation (λ = 1.542 Å). The XRD spectra were collected

within the range 10°≤2θ≤60° for clay, modified clay, and PLA nanocomposites at

2°/min scanning speed at room temperature.

X-ray photoelectron spectroscopy (XPS), also known as electron spectroscopy

for chemical analysis (ESCA), was outsourced at CoreLabs at King Abdullah

University of Science and Technology (KAUST), KSA.


36

A differential scanning calorimeter (TA Instruments, 25DSC) was used to

determine thermal properties such as glass transition temperature (Tg), cold

crystallization temperature (Tcc), melting temperature (Tm), enthalpy of heating (ΔHm),

enthalpy of cold crystallization (ΔHcc), and percentage crystallinity of PLA/clay

nanocomposites. The experiments were performed under constant inert nitrogen flow

(50 mL/min). The samples were heated from room temperature to 200°C at 10°C/min

to remove thermal history followed by cooling to 20°C at 10°C/min to record Tc, and

subsequently heated to 200°C at the same heating rate to record Tm. Thermal

Gravimetric Analysis (TGA) was performed by using a thermogravimetric analyzer

50Q (TA Instruments, USA) in order to evaluate thermal stability of clay nanoparticles

and nanocomposites. Approximately, 20 mg sample was heated up to 800°C at 20°C

/min under inert nitrogen flow (50 mL/min).

Mechanical characteristics were determined by a Universal Tensile Testing

Instrument (Shimadzu, AGS-X Series, 20 kN) following ASTM D638 at an extension

rate of 10 mm/min at room temperature. For each sample, five specimens were tested

and average were calculated and reported for tensile properties.

2.4 Results and Discussion

2.4.1 Characterization of Nano-Clay

The FTIR spectra of nano-clay and modified nano-clay are shown in Figure 5.

Major peaks corresponding to Si and water were same in both spectra. The absorption

peak at ~978 cm-1 was assigned to Si-OH group [103] At ~1642 cm-1, amide carbonyl

group was observed . At ~3613 cm-1 stretching -OH was shown [104], while the band
37

at 1032 cm−1 might be attributed to the siloxane (Si-O-Si) group stretching. The band

at 913 cm−1 was due to Al-OH [105].

For modified nano-clay Some bands appearing in more intense at 2950 cm-1

and 2880 cm-1 those were attributed to the C–H asymmetric and symmetric stretching

of methyl group, which confirmed a successful reaction between dimethyl alkyl amine

group and clay [104, 106]. Moreover, the band around 1400 cm−1 is attributed to the

presence of N–H on the structure due to amine group modification. It can be noticed

that the –OH stretching which is at ~3613 cm-1 is more intense in modified clay than

clay which could be attributed to the reaction between the –OH and the water

molecules which appears at ~3700 cm-1 in modified clay [106].

Figure 5: FTIR spectra of nano clay particles


38

The X-ray diffraction (XRD) is an effective method to investigate exfoliation of

nanoclay particles. In addition, information on dispersion of clay particles in

nanocomposite films can also be obtained.

In Figure 6, the nanoclay and modified nanoclay particles show one

characteristic peak at 2θ ~19.7°, indicating that the d-spacing calculated using the

Bragg’s law (   2d sin  ,where λ is the wavelength of x-rays, θ is diffraction angle,

and “d” is the interlayer spacing) between the silicate layer was about 4.5 nm,

corresponding to the (001) crystalline plane of clay nanoparticles [107]. Meanwhile,

Figure 7 shows that, in the case of PLA/modified nanoclay, a considerable shift in d-

spacing was observed, from 4.5 nm for modified nanoclay to 5.3 nm for PLA/modified

nanoclay composite. This in turn, indicates that modified nanoclay particles were

successfully intercalated in PLA resulting in a better exfoliation [108]. It is worth

stating that, the PLA shows one intense characteristic peak at 2θ ~16.6° with a

corresponding d-spacing of 5.4 nm which is almost the same as that of PLA

nanocomposites. It can be noticed from Figure 7 that the most intense peak appears at

2θ ~ 16.6° corresponding to the reflections from (110) and (200) planes of

orthorhombic α crystalline phase [109, 110]. The less intense peak appeared at

2θ ~19.1° corresponding to lattice plane (203) for PLA [92, 111].


39

Figure 6: XRD of clay nanoparticles

Figure 7: XRD of PLA/Modified clay nanocomposites

The XPS analysis revealed the presence of O1s core level spectra for both clay

and modified clay nanoparticles. As shown in Figure 8a, for clay nanoparticles, O 1s is

generally observed as a strong peak around 531 eV. Several studies done on clays
40

[112-114] and micas [115] of different clay minerals showed that O1s is observed

around 532 eV. Peak fitting revealed the presence of oxygen around 531.23 eV and

hydroxyl groups appeared ~532.234 eV [116]. Almost the same peaks appeared of O1s

core level spectra for modified clay nanoparticles as shown in Figure 8b. The slightly

higher energy in O 1s spectra for modified clay may be attributed to the amine group

attached onto the clay by reaction between the surface hydroxyls groups and the alkyl

groups of amine.

The high resolution of C1s region of modified nanoclay sample showed the

bands at 284.5, 286.1 and 287 eV, corresponding to C-C bond in the long chain, C-N

or C-O and COOR bonds, respectively [117] as illustrated in Figure 8c. The most

relevant atomic percentage variations on functionalization of bentonite with amine are

the increases in carbon and nitrogen surface contents and the decrease in oxygen

content, since they may confirm that amine was attached onto the clay by reaction

between the surface hydroxyls groups and the alkyl groups of amine. This reaction is

catalyzed by the interlayer clay water, leading to a decrease in water content (also

confirmed by IR), and consequently decrease in the O% [118].

The N1s peak was decomposed into two components located at 402.3 and 404

eV as shown in Figure 8d. The peak at 402.3 eV maybe attributed to non-protonated

nitrogen, while the peak at 404 eV can be due to the presence of protonated nitrogen

[118, 119], which in turn assigned to different binding strengths (strong electrostatic

and weak van der Waals links, respectively) for nitrogen from the amine group towards

the clay surface [117, 120]. These results suggest a change in the atoms environment

as a consequence of the grafting reaction [118].


41

Overall, it can be concluded from this section that the clay nanoparticles have

a variety of groups, which enable them to successfully intercalated in the PLA resulting

in exfoliation enhancement (illustrated by IR and XRD). In addition, interesting

composition surface characteristics were shown by the XPS analysis. It was observed

that the binding energy of O 1s of clay is slightly higher than that of modified clay due

to the modification of clay with amine group which in turn is expected to play a role

in enhancing the thermal as well as the mechanical properties of the PLA/Modified

clay nanoparticles more than that of PLA/clay nanoparticles.

a) b)

Nanoclay, O 1s Modified nanoclay, O 1s


531.234, O
Intensity (a.u)
Intensity (a.u)

532.657, OH

532.234, OH
531.657, O

510 520 530 540 550 510 520 530 540 550
Binding energy (eV) Binding energy (eV)

Figure 8: XPS of clay and Modified clay nanoparticles. XPS of (a) O s1 core level
spectra of clay nanoparticles. XPS of (b) O s1 core level spectra of Modified clay
nanoparticles. XPS of (c) C s1 core level spectra of Modified clay nanoparticles. XPS
of (d) N s1 core level spectra of Modified clay nanoparticles
42

c) d)
Modified Nanoclay, C 1s Modified Nanoclay, N 1s 403.957,
Protonated
286.957, C-C

Intensity (a.u)
Nitrogen
Intensity (a.u)

402.257,
284.457, 286.057, C-N Nonprotonated
COOR Nitrogen

270 280 290 300 310 380 390 400 410 420
Binding energy (eV) Binding energy (eV)

Figure 8: XPS of clay and modified clay nanoparticles. XPS of (a) O s1 core level
spectra of clay nanoparticles. XPS of (b) O s1 core level spectra of Modified clay
nanoparticles. XPS of (c) C s1 core level spectra of Modified clay nanoparticles. XPS
of (d) N s1 core level spectra of Modified clay nanoparticles (Continued)

2.4.2 Characterization of PLA nanocomposites

Figure 9 illustrates the FTIR spectra of PLA nanocomposite. The characteristic

amide N–H stretching vibration (primary amine) appeared at wave number (~3500 cm-

1) due to hydrogen bonding. The absorption peaks at ~3010 and ~2942 cm-1 were

assigned to aromatic and aliphatic C–H stretching vibrations [104]. The amide

carbonyl also appeared at lower wave number (~1751 cm-1) due to hydrogen bonding

and it also refers to esters because of the strong C=O stretching [104]. It can be noticed

that, absorption peak at 1450 cm-1 was given to alkane (methyl group) C-H bending.

The sulfonamide appeared at ~1359 cm-1 due to the strong S=O stretching. A medium

appearance was noticed for C–N stretching amine at the wave length of ~1183 cm-1.

At ~1081 cm-1 wave number, a strong C–O stretching aliphatic ether is appearing

[104].
43

From IR analysis, it can be concluded that the main difference between the clay

and the modified nanocomposites is the presence of characteristic amide N–H

stretching vibration (primary amine) at wave number (~3500 cm-1) for PLA/Modified

clay, while this group is not existing in the PLA/Clay nanocomposites. Most likely,

this can be as an attribution to the enhancement in the thermal and mechanical

properties of PLA/Modified clay compared to PLA/Clay nanocomposites.

Figure 9: FTIR spectrum of PLA/Modified clay nanocomposites

In order to study the influence of nanoclay and modified nano clay on

crystallinity (Xc), melting (Tm), cold crystallization (Tcc) and glass transition (Tg)

temperatures of PLA/Clay and PLA/modified nanocomposites, a differential scanning

calorimeter (DSC) was used. For the purpose of removing the thermal history, samples

were heated rapidly above their melting temperatures. The second heating cycle was

used to record Tm and Tg as discussed in the following section. Figure 9 illustrates Tm

and Tg values for pure PLA, PLA/clay and PLA/modified clay nanocomposites at 7
44

wt%. It can be noticed that, both clay and modified clay enhanced the glass transition

of PLA. However, this enhancement is slightly more noticeable for modified clay

where Tg rised up from ~52°C for PLA to ~54°C for PLA/modified clay composite,

attributed to increased polymer-filler interactions.

The nanocomposites PLA/Clay exhibited Tcc peak values (~105°C) which is

almost similar to that of PLA. However, PLA/Modified composites clay showed an

increase in the Tcc (~112°C) compared to that of PLA. The enhancement in the Tcc of

the PLA/Moified clay nanocomposites indicate nucleation behavior of modified clay

nanofiller. Compared to PLA Tm value (~165°C), PLA/Modified clay nanocomposites

Tm slightly increased (~167°C), while, a slightly lower Tm value was recorded for

PLA/Clay nanocomposites (~161°C).

In conclusion, the glass transition temperature as well as cold crystallization

was enhanced by modified nanoclay incorporation with PLA, whereas the melting

behavior did not change significantly. It is worth stating that, the produced

PLA/Modified nanocomposites show promising thermal characteristics compared to

several PLA composites. For instance, the maximum observed Tm (169°C) of

PLA/Clay at 5wt% and the maximum observed ΔHm (42 J/g) of PLA/Modified

nanocomposites at 5wt% are higher than reported values of PLA/cellulose crystal

composites for which the Tm and the ΔHm are 148°C and 22 J/g, respectively. While

they have higher Xc (24%) than that of PLA/ clay and modified clay nanocomposites

at 5wt% filler [121]. Also, it was found that, the PLA/graphene oxide nanosheets

composites have lower values of Tm (161°C), ΔHm (36 J/g), Tcc (109°C) and Xc (1.9%)

compared to the produced PLA/clay and modified clay at 5wt% [122]. Moreover, the

maximum Tg value of PLA/clay and modified clay nanocomposites (57°C) is higher


45

in comparison with PLA/microcrystalline cellulose (56°C) at 1wt% filler. For the same

PLA/microcrystalline cellulose, the Tm and ΔHm values of 149°C and 24 J/g,

respectively, are also lower than that of the resulted PLA/clay and modified clay

nanocomposites at 5wt% [123].

Figure 10: DSC curves of PLA, PLA/Clay and PLA/Modified Clay nanocomposites
at 7 wt% clay loadings

Table 3 summarizes the DSC data of PLA, PLA/Clay and PLA/Modified clay

nanocomposites. The highest values of the melting temperature (Tm), glass transition

temperature (Tg) and cold crystallization temperature (Tcc) were recorded for

PLA/Modified clay nanocomposites. On the other hand, it can be shown that the

enthalpy of cold crystallization (ΔHcc) increased for both PLA/Clay and

PLA/Modified clay nanocomposites compared to that of PLA with increasing clay

contents, whereas enthalpy of heating (ΔHm) and percentage crystallinity (Xc) of

PLA/Clay nanocomposites and PLA/Modified nanocomposites decreased under the


46

same conditions. However, modified clay reduced crystallinity of PLA more than what

was observed in PLA/clay nanocomposites. Thus, a higher elongation at break and

lower tensile modulus are expected for PLA/modified clay nanocomposites.

The percentage crystallinity (Xc) of PLA, PLA/Clay and PLA/Modified

nanocomposites was calculated using the following equation [124]:

H m  Hcc
X c (%)  100%
H 100% 1   
[1]

Where ΔHm is the experimental enthalpy of melting obtained from the DSC

melting endotherms, ΔH° is the enthalpy of melting of the 100% crystalline PLA

(93.7 J/g) [125] , ΔHcc is the enthalpy of cold crystallization, and θ is the weight

fraction of the filler in the matrix. The Xc decreased with increasing filler content in

both nanocomposites compared to PLA (Xc% = 42.8) (Figure 10). The PLA/Modified

clay nanocomposites exhibited more decrease om crystallinity even at higher clay

contents which might be beneficial for manufacturing transparent PLA/Modified clay

nanocomposite thin films.


47

Table 3: DSC data of PLA, PLA/Clay and PLA/Modified clay nanocomposites

Samples Filler Tg (°C) ΔHcc Tm (°C) ΔHm Xc%


Wt% (J/g) (J/g)
PLA 0 52±1.2 10.4 165±1.03 50.5 42.8

1 57.6 30.3 169.7 42.6 13.3


3 57±1.21 26.9 170±1.0 37.1 11.2
PLA/Clay
5 54±1.31 24.5 166±1.4 42.1 19.8
7 53.5 22.6 160.6 41.9 20

1 55.7 30.5 168.6 38.4 10.6


PLA/ 3 56±1.4 31 169±1.05 39.5 9.3
Modified
clay 5 57±1.62 28.5 169±1.45 36.6 9.1

7 54.3 21.6 166.7 36.8 4.5

Figure 11: The effect of PLA/Clay and PLA/Modified clay nanocomposites on


crystallinity
48

Thermogravimetric analysis (TGA) was used to investigate thermal stability

of PLA and PLA nanocomposites. Figure 11 shows the weight loss curves as well as

the corresponding derivative weight loss of pure PLA and PLA nanocomposites. The

thermal stability of PLA/Clay nanocomposite was lower than that of PLA whereas

PLA/Modified clay exhibited higher thermal stability compared to PLA. The thermal

decomposition temperature (Td) of PLA/Clay nanocomposite is lower than that of

PLA. On the other hand, Td of PLA showed an enhancement by ~7°C towards higher

temperature, from ~380°C to ~387°C for PLA/Modified clay nanocomposite.

The TGA results are presented in Table 3 quantitatively. For PLA/Clay

nanocomposite, 5% weight loss occurred at ~311°C while PLA/Modified clay

nanocomposites lost the same weight percent at ~344°C. This means that the thermal

degradation of PLA/Modified clay nanocomposite occurs at higher temperature

compared to PLA which lost 5% of its weight at 336°C. Moreover, it could be stated

from Table 4 that the overall weight loss (%) of PLA nanocomposites enhanced from

99% overall weight loss for PLA to 93.5% and 94.5% weight loss for PLA/Clay and

PLA/Modified nanocomposites respectively. The results indicate improved

interactions between PLA and modified clay that consequently, might have increased

thermal stability of the nanocomposites.


49

Figure 12: Thermal stability of PLA and PLA nanocomposites

Table 4: Decomposition temperature and overall weight loss of PLA and PLA
nanocomposites

Samples T(°C) at T(°C) at Td Overall


5% weight 50% weight (°C) weight
loss loss change (%)
PLA 336 372 380 99
PLA/ clay 311 364 374 93.5
PLA/ Modified 344 381 387 94.5
clay

The influence of PLA/Clay and PLA/Modified clay nanocomposites on

mechanical properties was investigated using a universal testing machine. Tensile

stress-strain curves and the fracture points (εf, σf) (represented by arrows) of pure PLA,

PLA/clay, and PLA/Modified clay matrices are shown in Figure 13. Compare with

PLA/clay and PLA/Modified clay, PLA is observed to experience more ductility prior

to failure that took place at 20 MPa and 10.24% elongation. This necking/strain
50

hardening could occur mainly due to the sudden deformation and then the stretching

as stress increases [126] and it can be due to dislocation movements and dislocation

generation within the crystal structure of the material [127]. On the other hand,

embrittlement behavior was observed for PLA nanocomposites system due to the

incorporation of clay and modified clay to the PLA matrix which eliminates the

orientation apparent strain hardening experienced by the pure polymer [128]. The

break stress of PLA/clay, and PLA/Modified clay is higher compared to pure PLA for

the same strain, indicating that the addition of clay and modified clay improved

effectively the modulus and strength of PLA matrix. It is worth stating that, the fracture

stress value of the prepared PLA/Modified clay composite (~30 MPa) is higher than

that of PLA/ microcrystalline cellulose (MCC) and silver (Ag) films (~ 21 MPa) [123]

While it is lower than that of PLA/cellulose nanofiber films [129] and PLA/graphene

nanoplatelet [128] with ~ 58 MPa and ~59 MPa, respectively.

Figure 13: Stress-strain curves of PLA, PLA/Clay, and PLA/Modified clay


composites
51

The effects of clay and modified clay on the Young’s modulus and yield

strength of the PLA/nanocomposites are shown in Figures 14 and 15 respectively. The

PLA/clay and PLA/modified clay exhibited Young’s moduli of ~697 MPa and ~764

MPa at 7 wt%, respectively, which were higher than neat PLA (~502 MPa). In general,

the PLA/modified clay nanocomposites showed lower modulus value than that of

PLA/Clay nanocomposites which might be attributed to low crystallinity observed in

DSC analysis for PLA/modified clay nanocomposites. However, PLA/clay

nanocomposites have lower of tensile yield strength than PLA/Modified clay on each

% increase. The modified clay had a noticeable beneficial effect on the tensile yield

strength of the nanocomposites which increased with increasing modified clay content

(1 and 3 wt%). A dramatic increase in the yield strength from ~12 MPa for PLA up to

16 MPa for PLA/Modified clay composite by 3 wt% modified clay content, indicating

33% enhancement in the yield strength. This improvement can be due to the well

dispersed of modified clay particles in the polymer matrix in comparison with

nanoclay particles. The reported values of PLA/graphene nanoplatelets [128],

PLA/cellulose nanocrystals [121] and PLA/fiber [130] are 1290 MPa, 1442MPa and

8300 MPa, respectively. Several studies reported the yield strength value of different

types of PLA composites. For instance, the yield strength of PLA/nanofiber [131]

PLA/cork [13] and PLA/starch [132] are 50.7 MPa, 38.3 MPa and 28 MPa

respectively, which illustrate that clay is similar to cork and starch in improvement but

orders of magnitude less than graphene and cellulose nanomaterials.


52

900

Young's Modulus (Mpa)


PLA-Clay
800
PLA-MClay
700
600
500
400
300
200
100
0
0 1 3 5 7
Filler wt%

Figure 14: Young’s modulus (MPa) of PLA nanocomposites

18
Yield Strength (Mpa)

16 PLA-Clay
PLA-MClay
14
12
10
8
6
4
2
0
0 1 3 5 7
Filler wt%

Figure 15: Yield strength (MPa) of PLA nanocomposites


53

Usually, a well-developed polymer/clay nanocomposite results in highly

increased mechanical strength compared to the pure polymer matrix since uniform

dispersion of the nano sized clay particles produces an ultra-high interfacial area and

ionic bonds between the nanoclay and the host polymer [133-135]. In the present

study, the addition of modified nanoclay enhanced both the tensile yield strength

compared to nanoclay. This indicates that the modified clay particles are well

dispersed in the polymer matrix in comparison with nanoclay particles [136] due to

the reaction occurring between the hydroxyl group and amine group of modified clay

and the carboxyl groups of PLA. Also, the increased degree of modified clay

dispersion found through narrow-angle X-ray scattering measurements promotes the

formation of a larger interfacial area and subsequent interfacial interaction between

modified clay and PLA. Therefore, the stress transfer from the polymeric matrix to the

inorganic phase increases, leading to enhance the tensile strength and modulus of the

PLA composites [72].

Table 5 and Figure 16 show the elongation at break values of PLA/Clay and

PLA/Modified clay nanocomposites. Generally, the elongation values decreased with

increasing clay content in both PLA/Clay and PLA/Modified clay nanocomposites.

The elongation reduction of PLA/Modified clay composites is lower than that of

PLA/Modified clay composites. These results propose that PLA/Modified clay

nanocomposites exhibited higher toughness and ductility compared to PLA/Clay

nanocomposites, which showed a brittle behavior. A similar behavior was observed by

Guohua et al (2006) for PVA/methylated-cornstarch composites [137]. Though the

elongation at break decreased after compounding with nanoclays, those of the

composite films are still comparable and competitive to those of widely used plastic

films such as PVA/methylated-cornstarch, and starch films, of which the elongation at


54

break values are in the range of 1.53-1.64 and 0.94-2.47, respectively [137, 138].

Moreover, the reported elongation at break values of PLA/Clay and PLA/Modified

clay at 3 wt% are 5.88 and 8.36 respectively, which were found to be greater than that

of cellulose nanocrystals (5.5%) for the same % filler content [121].

Table 5: Elongation at break of PLA, PLA/Clay and PLA/Modified clay


nanocomposites

Samples Filler (wt%) Elongation at break (%)

PLA 0 10
PLA/Clay 1 5
3 6
5 5
7 5.2
PLA/Modified 1 11
clay 3 8
5 6
7 3

Figure 16: Elongation at break for PLA composites


55

Among the nanoclays tested, modified nanoclay was more effective in

maintaining tensile yield strength property than the nanoclay, indicating that the

modified nanoclay is more compatible and has better homogeneous dispersion with

PLA, as compared to the nanoclay. However, high modified nanoclay content could

led the particles of the modified nanoclay to agglomerate within the matrix, causing a

poor dispersion and thus less enhancement in mechanical properties. For mechanical

properties enhancement such as modulus and tensile yield strength, the optimum

modified clay content is 7 wt% and 3 wt%, respectively.

For the nitrogen gas permeability test, the samples were cut into circular shape

with diameter of 0.037 m and 0.0001 m thickness. Then the sample was placed

between the two plates which are made of stainless steel. The Nitrogen gas

permeability of each sample was measured under a pressure ranged between 0.5 and 4

bar.

The apparent permeability (ka), in m2, was calculated at constant pressure using

the Hagen-Poiseuille expression for a compressible fluid. Since gases are

compressible, the inlet pressure P (applied test pressure (absolute)) at which the flow

rate (Q) is measured must be taken into account in addition to the outlet pressure Patm

(atmospheric pressure). The following equation was used to calculate the apparent

permeability (ka) [139]:

(2.𝑄.𝑃𝑎𝑡𝑚 .𝐿.𝜇)
ka= [2]
(𝑃𝑖2 −𝑃𝑎𝑡𝑚
2 )𝐴
56

Where,

A= cross-sectional Area (m2)

L= sample thickness crossed by the flow (m)

 = gas dynamic viscosity (Ns/m2). For nitrogen,  = 1.76×10^-5 Ns/m2 at 20°C [140]

Pi= inlet pressure (Pa)

Q= measured gas flowrate (m3/s)

Generally, it can be noticed that as the pressure increases the permeability of

neat PLA as well as of the nanocomposites films decrease as shown in Figure 17 and

18. From Figure 17, it can be noticed that, for PLA/Clay nanocomposites, there is

almost no noticeable enhancement in the permeability except for 7 wt% of clay content

with 60% reduction in nitrogen permeability at 400000 Pa. This improvement in

barrier property is because of the tortuous path being created by the clay particles

which retards the diffusion of the gas molecules through the matrix region. As clays

are crystalline materials, they are expected to increase the barrier properties by creating

a maze or “tortuous path” that restricts the progress of the gas molecules to pass

through the polymer matrices [141]. In comparison, almost all the PLA/Modified

nanocomposite films showed better nitrogen barrier properties than neat PLA films as

shown in Figure 18. Amongst all the PLA/Modified nanocomposites, the

nanocomposite having 5 wt% of modified clay content exhibited the highest reduction

in nitrogen permeability, approximately 76% which is higher than that (26%) of

PLA/Clay the same filler content. The enhanced nitrogen reduction of PLA/Modified

clay nanocomposites compared to that of PLA/Clay nanocomposites is attributed to

the improved dispersion and exfoliation of PLA/Modified clay in comparison to that


57

of PLA/Clay nanocomposites [142] as was supported by the XRD analysis. This is

due to the dispersion of amine of the modified clay in the PLA matrix which resulted

in the effective hydrogen bonding interaction among amine group and carbonyl back

bone of PLA which in turn improved the tortuosity passage for nitrogen gas [109].

PLA PLA-Clay 1 wt% PLA-Clay 3 wt% PLA-Clay 5 wt% PLA-Clay 7 wt%


Gas permeability (m2)

3.5E-21
3E-21
2.5E-21
2E-21
1.5E-21
1E-21
5E-22
0

Pressure (Pa)

Figure 17: Nitrogen gas permeability of PLA/Clay nanocomposites


58

PLA PLA-Modified Clay 1 wt%


PLA-Modified Clay 3 wt% PLA-Modified Clay 5 wt%
PLA-Modified Clay 7 wt%
3.5E-21

Gas Permeability (m2) 3E-21

2.5E-21

2E-21

1.5E-21

1E-21

5E-22

Pressure (Pa)

Figure 18: Nitrogen gas permeability of PLA/Modified clay nanocomposites


59

Chapter 3: Conclusions and Future Recommendation

3.1 Conclusions

In conclusion, in this study, clay and modified clay nanoparticles were

incorporated with PLA to produce PLA/clay and PLA/modified clay nanocomposites

using solvent blending technique. PLA chains intercalate galleries of clay

nanoparticles. The thermal, mechanical, and morphological properties of the produced

nanocomposites were characterized. The effect of modified clay on the exfoliation and

dispersion improvement of the PLA/modified clay nanocomposites was confirmed by

the appearance of wider peak than that of the PLA/clay nanocomposites in the XRD

for which the peak also shifted to lower 2θ. Generally, the highest values of the melting

temperature (Tm), glass transition temperature (Tg) and cold crystallization

temperature (Tcc) were recorded for PLA/Modified clay nanocomposites. On the other

hand, it can be shown that the enthalpy of cold crystallization (ΔHcc) increased for both

PLA/Clay and PLA/Modified clay nanocomposites compared to that of PLA, whereas,

the enthalpy of heating (ΔHm) and percentage crystallinity (Xc) of PLA/Clay

nanocomposites and PLA/Modified nanocomposites decreased in comparison with

PLA. Due to the chemical structure and low crystallinity of modified nanoclay, its

addition reduced crystallinity of PLA more than the addition of nanoclay which is

beneficial for manufacturing transparent nanocomposite thin films. Therefore, higher

elongation at break and lower tensile modulus were shown by PLA/Modified clay

nanocomposites, thus, higher toughness and ductility. In addition, analysis of

mechanical data indicates that the optimum modified clay concentration in the

PLA/Modified clay nanocomposites might be 3 wt%. Higher than 3 wt% modified

clay might cause destitute dispersion of modified clay nanoparticles due to


60

agglomeration coming about in diminished mechanical properties. Moreover, the

results of nitrogen gas permeability revealed that, the PLA/Modified clay

nanocomposites exhibited the best nitrogen gas permeability reduction at 5 wt%,

which is higher than that achieved by PLA/Clay nanocomposites at the same filler

load.

3.2 Future Recommendations

Further work may include extensional rheology to assess feasibility of these

nanocomposites for film blowing applications. Moreover, melt blending technique

could be used for the preparation of the same type of PLA nanocomposites prepared

in this study in order to investigate the differences. This would be accommodating in

investigating the impact of method utilized to prepare such samples on the properties

of the delivered PLA nanocomposites.


61

References

[1] E. Castro-Aguirre, F. Iniguez-Franco, H. Samsudin, X. Fang, and R. Auras, "Poly (lactic


acid)—Mass production, processing, industrial applications, and end of life," Advanced
Drug Delivery Reviews, vol. 107, pp. 333-366, 2016.
[2] M. Murariu and P. Dubois, "PLA composites: From production to properties,"
Advanced Drug Delivery Reviews, vol. 107, pp. 17-46, 2016.
[3] K. M. Nampoothiri, N. R. Nair, and R. P. John, "An overview of the recent
developments in polylactide (PLA) research," Bioresource Technology, vol. 101, no.
22, pp. 8493-8501, 2010.
[4] Y. Zhao, Z. Wang, J. Wang, H. Mai, B. Yan, and F. Yang, "Direct synthesis of poly (D,
L‐lactic acid) by melt polycondensation and its application in drug delivery," Journal
of Applied Polymer Science, vol. 91, no. 4, pp. 2143-2150, 2004.
[5] R. Mehta, V. Kumar, H. Bhunia, and S. N. Upadhyay, "Synthesis of poly (lactic acid):
a review," Journal of Macromolecular Science, Part C: Polymer Reviews, vol. 45, no.
4, pp. 325-349, 2005.
[6] J. S. Bergstroem and D. Hayman, "An overview of mechanical properties and material
modeling of polylactide (PLA) for medical applications," Annals of Biomedical
Engineering, vol. 44, pp. 330-340, 2016.
[7] C. Zhang, L. Wang, T. Zhai, X. Wang, Y. Dan, and L.-S. Turng, "The surface grafting
of graphene oxide with poly (ethylene glycol) as a reinforcement for poly (lactic acid)
nanocomposite scaffolds for potential tissue engineering applications," Journal of the
Mechanical Behavior of Biomedical Materials, vol. 53, pp. 403-413, 2016.
[8] R. P Pawar, S. U Tekale, S. U Shisodia, J. T Totre, and A. J Domb, "Biomedical
applications of poly (lactic acid)," Recent Patents on Regenerative Medicine, vol. 4, pp.
40-51, 2014.
[9] L. Li, S. Ding, and C. Zhou, "Preparation and degradation of PLA/chitosan composite
materials," Journal of Applied Polymer Science, vol. 91, pp. 274-277, 2004.
[10] E. Hoveizi, M. Nabiuni, K. Parivar, S. Rajabi‐Zeleti, and S. Tavakol, "Functionalisation
and surface modification of electrospun polylactic acid scaffold for tissue engineering,"
Cell Biology International, vol. 38, pp. 41-49, 2014.
[11] C. Chen et al., "Poly (lactic acid)(PLA) based nanocomposites—a novel way of drug-
releasing," Biomedical Materials, vol. 2, pp. 1-4, 2007.
[12] J.-Y. Park and I.-H. Lee, "Controlled release of ketoprofen from electrospun porous
polylactic acid (PLA) nanofibers," Journal of Polymer Research, vol. 18, no. 6, pp.
1287-1291, 2011.
[13] F. Daver, K. P. M. Lee, M. Brandt, and R. Shanks, "Cork–PLA composite filaments for
fused deposition modelling," Composites Science and Technology, vol. 168, pp. 230-
237, 2018.
62

[14] A. P. S. Immich, M. L. Arias, N. Carreras, R. L. Boemo, and J. A. Tornero, "Drug


delivery systems using sandwich configurations of electrospun poly (lactic acid)
nanofiber membranes and ibuprofen," Materials Science and Engineering: C, vol. 33,
pp. 4002-4008, 2013.
[15] S. Takenaka, M. Ishida, M. Serizawa, E. Tanabe, and K. Otsuka, "Formation of carbon
nanofibers and carbon nanotubes through methane decomposition over supported cobalt
catalysts," The Journal of Physical Chemistry B, vol. 108, no. 31, pp. 11464-11472,
2004.
[16] R. Geyer, J. R. Jambeck, and K. L. Law, "Production, use, and fate of all plastics ever
made," Science Advances, vol. 3, pp. 1-6, 2017.
[17] M. Rabnawaz, I. Wyman, R. Auras, and S. Cheng, "A roadmap towards green
packaging: the current status and future outlook for polyesters in the packaging
industry," Green Chemistry, vol. 19, pp. 4737-4753, 2017.
[18] K. Marsh and B. Bugusu, "Food packaging—roles, materials, and environmental
issues," Journal of Food Science, vol. 72, pp. 39-55, 2007.
[19] A. Arora and G. W. Padua, "Nanocomposites in food packaging," Journal of Food
science, vol. 75, no. 1, pp. 43-49, 2010.
[20] Y. Cui, S. I. Kundalwal, and S. Kumar, "Gas barrier performance of graphene/polymer
nanocomposites," Carbon, vol. 98, pp. 313-333, 2016.
[21] S. E. M. Selke and J. D. Culter, Plastics Packaging: Properties, Processing,
Applications, and Regulations. Carl Hanser Verlag GmbH Co KG, 2016.
[22] R. Mülhaupt, "Green polymer chemistry and bio‐based plastics: dreams and reality,"
Macromolecular Chemistry and Physics, vol. 214, pp. 159-174, 2013.
[23] P. F. H. Harmsen, M. M. Hackmann, and H. L. Bos, "Green building blocks for bio‐
based plastics," Biofuels, Bioproducts and Biorefining, vol. 8, no. 3, pp. 306-324, 2014.
[24] L. Shen, J. Haufe, and M. K. Patel, "Product overview and market projection of
emerging bio-based plastics PRO-BIP 2009," Report for European Polysaccharide
Network of Excellence (EPNOE) and European Bioplastics, vol. 243, pp. 1-245, 2009.
[25] G. Scott, "Green’polymers," Polymer Degradation and Stability, vol. 68, pp. 1-7, 2000.
[26] A. Sorrentino, G. Gorrasi, and V. Vittoria, "Potential perspectives of bio-
nanocomposites for food packaging applications," Trends in Food Science &
Technology, vol. 18, pp. 84-95, 2007.
[27] R. E. Drumright, P. R. Gruber, and D. E. Henton, "Polylactic acid technology,"
Advanced Materials, vol. 12, pp. 1841-1846, 2000.
[28] I. Vroman and L. Tighzert, "Biodegradable polymers," Materials, vol. 2, pp. 307-344,
2009.
[29] M. Ros-Chumillas, Y. Belissario, A. Iguaz, and A. López, "Quality and shelf life of
orange juice aseptically packaged in PET bottles," Journal of Food Engineering, vol.
79, no. 1, pp. 234-242, 2007.
63

[30] K. Sonneveld, "What drives (food) packaging innovation?," Packaging Technology and
Science: An International Journal, vol. 13, pp. 29-35, 2000.
[31] M. Ozdemir, C. U. Yurteri, and H. Sadikoglu, "Physical polymer surface modification
methods and applications in food packaging polymers," Critical Reviews in Food
Science and Nutrition, vol. 39, pp. 457-477, 1999.
[32] M. Moniruzzaman and K. I. Winey, "Polymer nanocomposites containing carbon
nanotubes," Macromolecules, vol. 39, pp. 5194-5205, 2006.
[33] K. I. Winey and R. A. Vaia, "Polymer nanocomposites," MRS Bulletin, vol. 32, pp. 314-
322, 2007.
[34] J. Pyun and K. Matyjaszewski, "Synthesis of nanocomposite organic/inorganic hybrid
materials using controlled/“living” radical polymerization," Chemistry of Materials,
vol. 13, pp. 3436-3448, 2001.
[35] C. E. Powell and G. W. Beall, "Solid State Mater," Science, vol. 10, pp. 73-80, 2006.
[36] A. Usuki, N. Hasegawa, M. Kato, and S. Kobayashi, "Polymer-clay nanocomposites,"
in Inorganic Polymeric Nanocomposites and Membranes: Springer, 2005, pp. 135-195.
[37] S.-Y. Park, Y.-H. Cho, and R. A. Vaia, "Three-dimensional structure of the zone-drawn
film of the nylon-6/layered silicate nanocomposites," Macromolecules, vol. 38, pp.
1729-1735, 2005.
[38] P. Rittigstein and J. M. Torkelson, "Polymer–nanoparticle interfacial interactions in
polymer nanocomposites: confinement effects on glass transition temperature and
suppression of physical aging," Journal of Polymer Science Part B: Polymer Physics,
vol. 44, no. 20, pp. 2935-2943, 2006.
[39] R. Guo, J. Azaiez, and C. Bellehumeur, "Rheology of fiber filled polymer melts: Role
of fiber‐fiber interactions and polymer‐fiber coupling," Polymer Engineering &
Science, vol. 45, pp. 385-399, 2005.
[40] H. X. Huang and J. J. Zhang, "Effects of filler–filler and polymer–filler interactions on
rheological and mechanical properties of HDPE–wood composites," Journal of Applied
Polymer Science, vol. 111, pp. 2806-2812, 2009.
[41] C. R. Martin, "Nanomaterials: a membrane-based synthetic approach," Science, vol.
266, pp. 1961-1966, 1994.
[42] R. Gref, Y. Minamitake, M. T. Peracchia, V. Trubetskoy, V. Torchilin, and R. Langer,
"Biodegradable long-circulating polymeric nanospheres," Science, vol. 263, pp. 1600-
1603, 1994.
[43] N. Herron and D. L. Thorn, "Nanoparticles: uses and relationships to molecular cluster
compounds," Advanced Materials, vol. 10, pp. 1173-1184, 1998.
[44] P. Das, S. Schipmann, J.-M. Malho, B. Zhu, U. Klemradt, and A. Walther, "Facile
access to large-scale, self-assembled, nacre-inspired, high-performance materials with
tunable nanoscale periodicities," ACS Applied Materials & Interfaces, vol. 5, no. 9, pp.
3738-3747, 2013.
[45] X. Sun, H. Sun, H. Li, and H. Peng, "Developing Polymer Composite Materials: Carbon
Nanotubes or Graphene?," Advanced Materials, vol. 25, pp. 5153-5176, 2013.
64

[46] K. Hu, D. D. Kulkarni, I. Choi, and V. V. Tsukruk, "Graphene-polymer nanocomposites


for structural and functional applications," Progress in Polymer Science, vol. 39, pp.
1934-1972, 2014.
[47] R. Coles, D. McDowell, and M. J. Kirwan, Food Packaging Technology. CRC Press,
2003.
[48] S. C. Tjong, "Structural and mechanical properties of polymer nanocomposites,"
Materials Science and Engineering: R: Reports, vol. 53, pp. 73-197, 2006.
[49] Y. Fukushima and S. Inagaki, "Synthesis of an intercalated compound of
montmorillonite and 6-polyamide," in Inclusion Phenomena in Inorganic, Organic, and
Organometallic Hosts: Springer, 1987, pp. 365-374.
[50] F. G. Giese and C. J. van Oss, "Colloid and Surface Properties of Clays and Related
Minerals. 2002," ed: New York: Marcel Dekker, Inc.
[51] C. S. Chern, "Emulsion polymerization mechanisms and kinetics," Progress in Polymer
Science, vol. 31, pp. 443-486, 2006.
[52] S. S. Ray and M. Okamoto, "Polymer/layered silicate nanocomposites: a review from
preparation to processing," Progress in Polymer Science, vol. 28, pp. 1539-1641, 2003.
[53] B. Chen et al., "A critical appraisal of polymer–clay nanocomposites," Chemical Society
Reviews, vol. 37, pp. 568-594, 2008.
[54] E. P. Giannelis, R. Krishnamoorti, and E. Manias, "Polymer-silicate nanocomposites:
model systems for confined polymers and polymer brushes," in Polymers in Confined
Environments: Springer, 1999, pp. 107-147.
[55] G. Lagaly, "From clay mineral crystals to colloidal clay mineral dispersions," Surfactant
Science Series, pp. 427-427, 1993.
[56] S. Caillère, S. Hénin and M. Rautureau, "Mineralogie des argiles. ii: classification et
nomenclature," Masson, Paris, 1982.
[57] E. Eslinger and D. Peaver, "Clay minerals for petroleum geologists and engineers,
SEPM Short course n 22, Soc," Economic Paleontologists and Mineralogists, Tulsa,
USA, 1988.
[58] J.-M. Raquez, Y. Habibi, M. Murariu, and P. Dubois, "Polylactide (PLA)-based
nanocomposites," Progress in Polymer Science, vol. 38, pp. 1504-1542, 2013.
[59] A. Blumstein, "Polymerization of adsorbed monolayers. II. Thermal degradation of the
inserted polymer," Journal of Polymer Science Part A: General Papers, vol. 3, no. 7,
pp. 2665-2672, 1965.
[60] R. Krishnamoorti, R. A. Vaia, and E. P. Giannelis, "Structure and dynamics of polymer-
layered silicate nanocomposites," Chemistry of Materials, vol. 8, pp. 1728-1734, 1996.
[61] Z. Wang, H. Wang, and M. E. Cates, "Effective elastic properties of solid clays,"
Geophysics, vol. 66, pp. 428-440, 2001.
[62] F. Gao, "Clay/polymer composites: the story," Materials Today, vol. 7, pp. 50-55, 2004.
[63] C. Edser, "Auto applications drive commercialization of nanocomposites," Plastics
Additives and Compounding, vol. 4, pp. 30-33, 2002.
65

[64] H. Cox, T. Dearlove, W. Rodges, M. Verbrugge, and C. S. Wang, "Nanocomposite


systems for automotive applications,", Springer, 2004, pp 329-356.
[65] A. Y. Goldman and C. J. Copsey, "Multilayer barrier liner material with nanocomposites
for packaging applications," 4th World Congress in Nanocomposites, EMC, San
Francisco. 2004. pp. 1-3.
[66] R. Conway, "Using nanotechnology to provide active and passive barriers in beverage
packaging application," Future of Nanomaterials Conference, Pira, Birmingham. 2004.
pp. 29-30.
[67] J. Radford, "Enhanced barrier properties in containers and films," Annals of the Future
of Nanomaterials Conference, Birminghan. 2004.
[68] E. Manias, A. Touny, L. Wu, K. Strawhecker, B. Lu, and T. C. Chung,
"Polypropylene/montmorillonite nanocomposites. Review of the synthetic routes and
materials properties," Chemistry of Materials, vol. 13, no. 10, pp. 3516-3523, 2001.
[69] S. C. Tjong and Y. Z. Meng, "Preparation and characterization of melt‐compounded
polyethylene/vermiculite nanocomposites," Journal of Polymer Science Part B:
Polymer Physics, vol. 41, pp. 1476-1484, 2003.
[70] J.-H. Lee, D. Jung, C.-E. Hong, K. Y. Rhee, and S. G. Advani, "Properties of
polyethylene-layered silicate nanocomposites prepared by melt intercalation with a PP-
g-MA compatibilizer," Composites Science and Technology, vol. 65, pp. 1996-2002,
2005.
[71] F. Yang, C. Mubarak, R. Keiegel, and R. M. Kannan, "Supercritical carbon dioxide
(scCO2) dispersion of poly (ethylene terephthalate)/clay nanocomposites: Structural,
mechanical, thermal, and barrier properties," Journal of Applied Polymer Science, vol.
134, pp. 1-11, 2017.
[72] J.-W. Rhim, S.-I. Hong, and C.-S. Ha, "Tensile, water vapor barrier and antimicrobial
properties of PLA/nanoclay composite films," LWT-Food Science and Technology, vol.
42, pp. 612-617, 2009.
[73] L. Shen, J. Haufe, and M. K. Patel, "Product overview and market projection of
emerging bio-based plastics," Report by Utrecht University commissioned by European
Polysaccharide network of excellence and European bioplastics, June 29, 2009.
[74] C. N. Cutter, "Opportunities for bio-based packaging technologies to improve the
quality and safety of fresh and further processed muscle foods," Meat Science, vol. 74,
pp. 131-142, 2006.
[75] Y. Lu and S. Ozcan, "Green nanomaterials: On track for a sustainable future," Nano
Today, vol. 10, pp. 417-420, 2015.
[76] D. Klemm, B. Heublein, H. P. Fink, and A. Bohn, "Cellulose: fascinating biopolymer
and sustainable raw material," Angewandte Chemie International Edition, vol. 44, no.
22, pp. 3358-3393, 2005.
[77] Q. Zhu, X. Zhou, J. Ma, and X. Liu, "Preparation and characterization of novel
regenerated cellulose films via sol–gel technology," Industrial & Engineering
Chemistry Research, vol. 52, pp. 17900-17906, 2013.
66

[78] O. Dechy-Cabaret, B. Martin-Vaca, and D. Bourissou, "Controlled ring-opening


polymerization of lactide and glycolide," Chemical Reviews, vol. 104, pp. 6147-6176,
2004.
[79] H. Ohara, S. Sawa, and T. Kawamoto, "Method for producing polylactic acid," U.S.
Patent 5 508 378, April, 16, 1996.
[80] R. Auras, B. Harte, and S. Selke, "Effect of water on the oxygen barrier properties of
poly (ethylene terephthalate) and polylactide films," Journal of Applied Polymer
Science, vol. 92, pp. 1790-1803, 2004.
[81] R. A. Auras, B. Harte, S. Selke, and R. Hernandez, "Mechanical, physical, and barrier
properties of poly (lactide) films," Journal of Plastic Film & Sheeting, vol. 19, no. 2,
pp. 123-135, 2003.
[82] E. A. Flexman, "Toughened poly (lactic acid) compositions," U.S. Patent 7 354 973 B,
April, 8, 2008.
[83] D. Mondal, M. Griffith, and S. S. Venkatraman, "Polycaprolactone-based biomaterials
for tissue engineering and drug delivery: Current scenario and challenges,"
International Journal of Polymeric Materials and Polymeric Biomaterials, vol. 65, pp.
255-265, 2016.
[84] S.-M. Lai, S.-H. Wu, G.-G. Lin, and T.-M. Don, "Unusual mechanical properties of
melt-blended poly (lactic acid)(PLA)/clay nanocomposites," European Polymer
Journal, vol. 52, pp. 193-206, 2014.
[85] S. Sinha Ray, K. Okamoto, K. Yamada, and M. Okamoto, "Novel porous ceramic
material via burning of polylactide/layered silicate nanocomposite," Nano Letters, vol.
2, pp. 423-425, 2002.
[86] S. Sinha Ray, K. Yamada, M. Okamoto, and K. Ueda, "Polylactide-layered silicate
nanocomposite: a novel biodegradable material," Nano Letters, vol. 2, pp. 1093-1096,
2002.
[87] S. Sinha Ray, P. Maiti, M. Okamoto, K. Yamada, and K. Ueda, "New
polylactide/layered silicate nanocomposites. 1. Preparation, characterization, and
properties," Macromolecules, vol. 35, pp. 3104-3110, 2002.
[88] S. S. Ray, K. Yamada, A. Ogami, M. Okamoto, and K. Ueda, "New polylactide/layered
silicate nanocomposite: nanoscale control over multiple properties," Macromolecular
Rapid Communications, vol. 23, pp. 943-947, 2002.
[89] S. Sinha Ray, K. Yamada, M. Okamoto, A. Ogami, and K. Ueda, "New
polylactide/layered silicate nanocomposites. 3. High-performance biodegradable
materials," Chemistry of Materials, vol. 15, pp. 1456-1465, 2003.
[90] S. S. Ray and M. Okamoto, "Biodegradable polylactide and its nanocomposites:
opening a new dimension for plastics and composites," Macromolecular Rapid
Communications, vol. 24, pp. 815-840, 2003.
[91] K. Fukushima, D. Tabuani, and G. Camino, "Poly (lactic acid)/clay nanocomposites:
effect of nature and content of clay on morphology, thermal and thermo-mechanical
properties," Materials Science and Engineering: C, vol. 32, pp. 1790-1795, 2012.
67

[92] V. Krikorian and D. J. Pochan, "Poly (L-lactic acid)/layered silicate nanocomposite:


fabrication, characterization, and properties," Chemistry of Materials, vol. 15, pp. 4317-
4324, 2003.
[93] N. Najafi, M. C. Heuzey, and P. J. Carreau, "Polylactide (PLA)-clay nanocomposites
prepared by melt compounding in the presence of a chain extender," Composites Science
and Technology, vol. 72, pp. 608-615, 2012.
[94] P. Krishnamachari, J. Zhang, J. Lou, J. Yan, and L. Uitenham, "Biodegradable poly
(lactic acid)/clay nanocomposites by melt intercalation: a study of morphological,
thermal, and mechanical properties," International Journal of Polymer Analysis and
Characterization, vol. 14, no. 4, pp. 336-350, 2009.
[95] M. A. Paul et al., "(Plasticized) polylactide/(organo‐) clay nanocomposites by in situ
intercalative polymerization," Macromolecular Chemistry and Physics, vol. 206, pp.
484-498, 2005.
[96] S. Y. Lee and M. A. Hanna, "Tapioca starch‐poly (lactic acid)‐Cloisite 30B
nanocomposite foams," Polymer Composites, vol. 30, pp. 665-672, 2009.
[97] N. Najafi, M. C. Heuzey, P. J. Carreau, and P. M. Wood-Adams, "Control of thermal
degradation of polylactide (PLA)-clay nanocomposites using chain extenders," Polymer
Degradation and Stability, vol. 97, pp. 554-565, 2012.
[98] R. Nanda, A. Sasmal, and P. L. Nayak, "Preparation and characterization of chitosan–
polylactide composites blended with Cloisite 30B for control release of the anticancer
drug paclitaxel," Carbohydrate Polymers, vol. 83, pp. 988-994, 2011.
[99] M. Jollands and R. K. Gupta, "Effect of mixing conditions on mechanical properties of
polylactide/montmorillonite clay nanocomposites," Journal of Applied Polymer
Science, vol. 118, pp. 1489-1493, 2010.
[100] J. Gamez‐Perez et al., "Influence of crystallinity on the fracture toughness of poly (lactic
acid)/montmorillonite nanocomposites prepared by twin‐screw extrusion," Journal of
Applied Polymer Science, vol. 120, pp. 896-905, 2011.
[101] B. Wang, T. Wan, and W. Zeng, "Rheological and thermal properties of
polylactide/organic montmorillonite nanocomposites," Journal of Applied Polymer
Science, vol. 125, pp. 364-371, 2012.
[102] B. Li, F.-X. Dong, X.-L. Wang, J. Yang, D.-Y. Wang, and Y.-Z. Wang, "Organically
modified rectorite toughened poly (lactic acid): Nanostructures, crystallization and
mechanical properties," European Polymer Journal, vol. 45, pp. 2996-3003, 2009.
[103] J. Madejová, "FTIR techniques in clay mineral studies," Vibrational Spectroscopy, vol.
31, pp. 1-10, 2003.
[104] A. Kausar, S. Haider, and B. Muhammad, "Nanocomposite based on
polystyrene/polyamide blend and bentonite: morphology, thermal, and nonflammability
properties," Nanomaterials and Nanotechnology, vol. 7, pp. 1-14, 2017.
68

[105] P. Kumararaja, K. M. Manjaiah, S. C. Datta, and B. Sarkar, "Remediation of metal


contaminated soil by aluminium pillared bentonite: synthesis, characterisation,
equilibrium study and plant growth experiment," Applied Clay Science, vol. 137, pp.
115-122, 2017.
[106] O. Zabihi, H. Khayyam, B. L. Fox, and M. Naebe, "Enhanced thermal stability and
lifetime of epoxy nanocomposites using covalently functionalized clay: experimental
and modelling," New Journal of Chemistry, vol. 39, pp. 2269-2278, 2015.
[107] B. Yalcin, Z. Ergungor, Y. Konishi, M. Cakmak, and C. Batur, "Molecular origins of
toughening mechanism in uniaxially stretched nylon 6 films with clay nanoparticles,"
Polymer, vol. 49, pp. 1635-1650, 2008.
[108] Suryani, H. Agusnar, B. Wirjosentono, T. Rihayat, and Nurhanifa, "Thermal
degradation of Aceh’s bentonite reinforced poly lactic acid (PLA) based on renewable
resources for packaging application," AIP Conference Proceedings. vol. 2049. no. 1.
AIP Publishing LLC, 2018.
[109] S. Loganathan, J. Jacob, R. B. Valapa, and S. Thomas, "Influence of linear and branched
amine functionalization in mesoporous silica on the thermal, mechanical and barrier
properties of sustainable poly (lactic acid) biocomposite films," Polymer, vol. 148, pp.
149-157, 2018.
[110] S. M. Bhasney, R. Patwa, A. Kumar, and V. Katiyar, "Plasticizing effect of coconut oil
on morphological, mechanical, thermal, rheological, barrier, and optical properties of
poly (lactic acid): A promising candidate for food packaging," Journal of Applied
Polymer Science, vol. 134, no. 41, p. 45390, 2017. https://doi.org/10.1002/app.45390
[111] K. Zhang, A. K. Mohanty, and M. Misra, "Fully biodegradable and biorenewable
ternary blends from polylactide, poly (3-hydroxybutyrate-co-hydroxyvalerate) and poly
(butylene succinate) with balanced properties," ACS Applied Materials & Interfaces,
vol. 4, pp. 3091-3101, 2012.
[112] C. Mosser, A. Mosser, M. Romeo, S. Petit, and A. Decarreau, "Natural and synthetic
copper phyllosilicates studied by XPS," Clays and Clay Minerals, vol. 40, pp. 593-599,
1992.
[113] T. L. Barr, S. Seal, H. He, and J. Klinowski, "X-ray photoelectron spectroscopic studies
of kaolinite and montmorillonite," Vacuum, vol. 46, pp. 1391-1395, 1995.
[114] W. D. Johns and S. Gier, "X-ray photoelectron spectroscopic study of layer charge
magnitude in micas and illite-smectite clays," Clay Minerals, vol. 36, pp. 355-367,
2001.
[115] K. G. Bhattacharyya, "XPS study of mica surfaces," Journal of Electron Spectroscopy
and Related Phenomena, vol. 63, pp. 289-306, 1993.
[116] J. T. Kloprogge and B. J. Wood, "Baseline studies of the clay minerals society source
clays by x-ray photoelectron spectroscopy," Clay Science, vol. 22, pp. 85-94, 2018.
[117] M. Gamba, P. Kovář, M. Pospíšil, and R. M. T. Sánchez, "Insight into thiabendazole
interaction with montmorillonite and organically modified montmorillonites," Applied
Clay Science, vol. 137, pp. 59-68, 2017.
69

[118] C. Pereira et al., "Copper acetylacetonate anchored onto amine-functionalised clays,"


Journal of Colloid and Interface Science, vol. 316, pp. 570-579, 2007.
[119] A. Xue, S. Zhou, Y. Zhao, X. Lu, and P. Han, "Effective NH2-grafting on attapulgite
surfaces for adsorption of reactive dyes," Journal of Hazardous Materials, vol. 194, pp.
7-14, 2011.
[120] P. M. Naranjo, E. L. Sham, E. R. Castellón, R. M. Torres Sánchez, and E. M. Farfán
Torres, "Identification and quantification of the interaction mechanisms between the
cationic surfactant HDTMA-Br and montmorillonite," Clays and Clay Minerals, vol.
61, pp. 98-106, 2013.
[121] S. H. Sung, Y. Chang, and J. Han, "Development of polylactic acid nanocomposite films
reinforced with cellulose nanocrystals derived from coffee silverskin," Carbohydrate
Polymers, vol. 169, pp. 495-503, 2017.
[122] H.-D. Huang et al., "Improved barrier properties of poly (lactic acid) with randomly
dispersed graphene oxide nanosheets," Journal of Membrane Science, vol. 464, pp. 110-
118, 2014.
[123] E. Fortunati, I. Armentano, A. Iannoni, and J. M. Kenny, "Development and thermal
behaviour of ternary PLA matrix composites," Polymer Degradation and Stability, vol.
95, pp. 2200-2206, 2010.
[124] J. Henricks, M. Boyum, and W. Zheng, "Crystallization kinetics and structure evolution
of a polylactic acid during melt and cold crystallization," Journal of Thermal Analysis
and Calorimetry, vol. 120, pp. 1765-1774, 2015.
[125] S. Jia, D. Yu, Y. Zhu, Z. Wang, L. Chen, and L. Fu, "Morphology, crystallization and
thermal behaviors of PLA-based composites: wonderful effects of hybrid GO/PEG via
dynamic impregnating," Polymers, vol. 9, no. 9, pp. 528-546, 2017.
[126] M. Nofar et al., "Mechanical and bead foaming behavior of PLA-PBAT and PLA-PBSA
blends with different morphologies," European Polymer Journal, vol. 90, pp. 231-244,
2017.
[127] E. P. DeGarmo, J. T. Black, R. A. Kohser, and B. E. Klamecki, Materials and Process
in Manufacturing. Prentice Hall Upper Saddle River, 1997.
[128] Y. Gao, O. T. Picot, E. Bilotti, and T. Peijs, "Influence of filler size on the properties of
poly (lactic acid)(PLA)/graphene nanoplatelet (GNP) nanocomposites," European
Polymer Journal, vol. 86, pp. 117-131, 2017.
[129] M. Jonoobi, J. Harun, A. P. Mathew, and K. Oksman, "Mechanical properties of
cellulose nanofiber (CNF) reinforced polylactic acid (PLA) prepared by twin screw
extrusion," Composites Science and Technology, vol. 70, pp. 1742-1747, 2010.
[130] K. Oksman, M. Skrifvars, and J. F. Selin, "Natural fibres as reinforcement in polylactic
acid (PLA) composites," Composites Science and Technology, vol. 63, pp. 1317-1324,
2003.
70

[131] M. Kowalczyk, E. Piorkowska, P. Kulpinski, and M. Pracella, "Mechanical and thermal


properties of PLA composites with cellulose nanofibers and standard size fibers,"
Composites Part A: Applied Science and Manufacturing, vol. 42, no. 10, pp. 1509-1514,
2011.
[132] A. Nuona, X. Li, X. Zhu, Y. Xiao, and J. Che, "Starch/polylactide sustainable
composites: Interface tailoring with graphene oxide," Composites Part A: Applied
Science and Manufacturing, vol. 69, pp. 247-254, 2015.
[133] M. Alexandre and P. Dubois, "Polymer-layered silicate nanocomposites: preparation,
properties and uses of a new class of materials," Materials Science and Engineering: R:
Reports, vol. 28, pp. 1-63, 2000.
[134] J. K. Pandey, A. P. Kumar, M. Misra, A. K. Mohanty, L. T. Drzal, and R. Palsingh,
"Recent advances in biodegradable nanocomposites," Journal of Nanoscience and
Nanotechnology, vol. 5, pp. 497-526, 2005.
[135] S. S. Ray and M. Bousmina, "Biodegradable polymers and their layered silicate
nanocomposites: in greening the 21st century materials world," Progress in Materials
Science, vol. 50, pp. 962-1079, 2005.
[136] N. Ogata, G. Jimenez, H. Kawai, and T. Ogihara, "Structure and thermal/mechanical
properties of poly (l‐lactide)‐clay blend," Journal of Polymer Science Part B: Polymer
Physics, vol. 35, pp. 389-396, 1997.
[137] Z. Guohua, L. Ya, F. Cuilan, Z. Min, Z. Caiqiong, and C. Zongdao, "Water resistance,
mechanical properties and biodegradability of methylated-cornstarch/poly (vinyl
alcohol) blend film," Polymer Degradation and stability, vol. 91, pp. 703-711, 2006.
[138] G. Scarascia-Mugnozza, E. Schettini, G. Vox, M. Malinconico, B. Immirzi, and S.
Pagliara, "Mechanical properties decay and morphological behaviour of biodegradable
films for agricultural mulching in real scale experiment," Polymer Degradation and
Stability, vol. 91, pp. 2801-2808, 2006.
[139] A. Talah and F. Kharchi, "A modified test procedure to measure gas permeability of
hollow cylinder concrete specimens," International Journal of Engineering and
Technology, vol. 5, no.1, pp. 91-95, 2013.
[140] V. Picandet, A. Khelidj, and H. Bellegou, "Crack effects on gas and water permeability
of concretes," Cement and Concrete Research, vol. 39, pp. 537-547, 2009.
[141] A. Bhatia, R. K. Gupta, S. N. Bhattacharya, and H. J. Choi, "Analysis of gas
permeability characteristics of poly (lactic acid)/poly (butylene succinate)
nanocomposites," Journal of Nanomaterials, vol. 2012, pp. 1-11, 2012.
[142] G. Choudalakis and A. D. Gotsis, "Permeability of polymer/clay nanocomposites: a
review," European Polymer Journal, vol. 45, no. 4, pp. 967-984, 2009.

Digitally signed by
Shrieen
DN: cn=Shrieen,
o=United Arab
Emirates University,
ou=UAEU Library
Digitizatio,
email=shrieen@uae
u.ac.ae, c=AE
Date: 2021.08.23
12:36:02 +04'00'

You might also like