Florida State University Libraries

Electronic Theses, Treatises and Dissertations The Graduate School

2010 Morphology and Properties of / Carbon Nanotube Nanocomposite Prepared by Supercritical Carbon Dioxide Nemat Hossieny

Follow this and additional works at the FSU Digital Library. For more information, please contact [email protected] THE FLORIDA STATE UNIVERSITY

COLLEGE OF ENGINEERING

MORPHOLOGY AND PROPERTIES OF POLYMER/CARBON

NANOTUBE NANOCOMPOSITE FOAMS PREPARED BY

SUPERCRITICAL CARBON DIOXIDE

By

NEMAT HOSSIENY

A Thesis submitted to the Department of Industrial and Manufacturing Engineering in partial fulfillment of the requirements for the degree of Master of Science

Degree Awarded: Spring Semester, 2010

The members of the committee approve the thesis of Nemat Hossieny defended on December 01, 2009. ______Changchun Zeng

Professor Directing Thesis

______

Zhiyong Liang

Committee Member

______

Tao Liu

Committee Member

Approved: ______Chuck Zhang, Chair, Department of Industrial and Manufacturing Engineering ______Ching-Jen Chen, Dean, College of Engineering

The Graduate School has verified and approved the above-named committee members.

To the Almighty God for His guidance, protection and love. To my beloved parents, and brothers for their selfless support in every step of my endeavor. Lastly to Zhinus, my friend and soul mate, for her unswerving belief in me and giving me the opportunity to spend the rest of my life with her.

iii

ACKNOWLEDGMENTS

Sincerely, I would like to appreciate and thank the invaluable contributions, guidance and motivation of some wonderful people that helped me to accomplish this research work. I would like to thank Dr. Changchun Zeng, my major professor for his patience, support and motivation. I would also like to thank Dr. Richard Liang and Dr. Tao Liu for their continued support and guidance serving as my committee members. I appreciate the contributions of Dr. Joseph Pignatiello towards the statistical aspect of the research. Additionally, I would like to acknowledge Dr. Jin-Gyu Park, Dr. David Cheng, Mr. Jerry Horne, Mr. Jim Horne, Mr. Frank Allen and Ms. Judy Gardner for their assistance during various aspects of my research work. I would further like to extend my appreciation to the High Performance Materials Institute (HPMI) for providing the financial support needed to complete this work. I would also like to appreciate the entire faculty of the Department of Industrial and Manufacturing Engineering for their support.

I would like to express my deep appreciation to all my friends and colleagues at Florida State University, as well as my friends at the Baha’i Student Association at Florida State University and also the Baha’i community of Tallahassee for their kind support and love. Finally, I am thankful to all my family members for your support and help to accomplish my hearts desire.

iv

TABLE OF CONTENTS

LIST OF TABLES ...... VIII LIST OF FIGURES ...... IX ABSTRACT ...... XI CHAPTER 1 ...... 1 INTRODUCTION...... 1

1.1 BACKGROUND AND SIGNIFICANCE ...... 1 1.2 SCOPE AND OBJECTIVE OF THE RESEARCH ...... 2 CHAPTER 2 ...... 3 LITERATURE REVIEW ...... 3

2.1 POLYMER CNT NANOCOMPOSITES...... 3 2.1.1 Introduction ...... 3 2.1.2 CNT and CNT Nanocomposites ...... 4 2.1.3 Surface Functionalization ...... 6 2.1.4 Nanocomposite Preparation ...... 8 2.1.4.1 Solvent Blending ...... 9 2.1.4.2 Melt Blending ...... 9 2.1.4.3 In-situ polymerization ...... 10 2.2 POLYMERIC ...... 11 2.2.1 Polymeric Foams and the Foaming Process ...... 11 2.2.2 Microcellular Foams ...... 12 2.2.2.1 Synthesis of Microcellular Foams ...... 13 2.2.2.2 Microcellular Foam Properties ...... 15

2.3 SUPERCRITICAL CO2 (SCCO2) IN FOAMING ...... 16 2.4 FUNDAMENTALS IN PHYSICAL FOAMING PROCESS ...... 17 2.4.1 Formation of Polymer/Foaming Agent Homogeneous Solution ...... 17 2.4.2 Cell Nucleation ...... 19 2.4.3 Cell Growth and Cell Stabilization ...... 21 2.5 POLYMER NANOCOMPOSITE FOAMS ...... 22 2.5.1 Effects of Nanoparticles on Foam Morphology ...... 22 2.5.2 Effects of Nanoparticles on Foam Properties ...... 24

CHAPTER 3 ...... 26 EXPERIMENTAL PROCEDURE...... 26

v

3.1 MATERIALS ...... 26 3.2 SYNTHESIS OF PMMA/MWCNT NANOCOMPOSITES ...... 26 3.2.1 Solvent Casting Method ...... 27 3.2.2 Anti-solvent Precipitation Method (ASP) ...... 28 3.3 PREPARATION OF PMMA/MWCNT NANOCOMPOSITE FOAMS ...... 29 3.3.1 Design and Construction of High Pressure Batch Foaming System ...... 29 3.3.2 Batch Foaming ...... 30 3.4 STRUCTURAL AND MORPHOLOGICAL CHARACTERIZATION ...... 31 CHAPTER 4 ...... 32 RESULTS AND DISCUSSION ...... 32

4.1 PURE PMMA FOAMS: EFFECT OF PROCESSING PARAMETERS ...... 32 4.2 PREPARATION OF NANOCOMPOSITES ...... 33 4.3 NANOCOMPOSITE FOAMS: BIMODAL MORPHOLOGY ...... 37 4.3.1 Effect of CNT Concentration ...... 37 4.3.2 Effect of CNT Surface Functionalization ...... 42 4.4 NANOCOMPOSITE FOAMS: MONO-MODAL MORPHOLOGY ...... 47 4.5 DESIGN AND ANALYSIS OF EXPERIMENTS FOR THE CELL SIZE COMPARISON ...... 52 4.5.1 Response Selection ...... 52 4.5.2 Choice of Factors ...... 53 4.5.3 Choice of Experiments ...... 53 4.5.4 DOE Analysis Results and Discussion ...... 54 4.6 CONCLUSIONS ...... 55 CHAPTER 5 ...... 57 MECHANICAL PROPERTIES OF PMMA/MWCNT NANOCOMPOISTE FOAMS ...... 57

5.1 INTRODUCTION ...... 57 5.2 EXPERIMENTAL PROCEDURE ...... 57 5.2.1 Preparation of PMMA/CNT nanocomposite ...... 57 5.2.2 Foaming procedure ...... 58 5.2.3 Tensile testing ...... 59 5.3 RESULTS AND DISCUSSION ...... 59 5.3.1 Mold design for foaming ...... 59 5.3.2 Tensile properties ...... 60

CHAPTER 6 ...... 65 CONCLUSION AND RECOMMENDATIONS ...... 65

6.1 CONCLUSIONS ...... 65 vi

6.2 RECOMMENDATIONS ...... 66 REFERENCES ...... 67 BIOGRAPHICAL SKETCH ...... 75

vii

LIST OF TABLES

TABLE 4.1 COMPARISON OF CELL SIZE AND CELL DENSITY OF PMMA CNT NANOCOMPOSITE FOAMS ...... 44

TABLE 4.2 COMPARISON OF NORMALIZED CELL SIZE OF PMMA AND NANOCOMPOSITE FOAMS ..... 44

TABLE 4.3 COMPARISON OF NORMALIZED CELL DENSITY OF PMMA AND NANOCOMPOSITE FOAMS……………………………………………………………………………….. 45

TABLE 4.4 EFFECT OF NANOCOMPOSITES PROCESSING METHODS ON THE FOAM MORPHOLOGY .... 49

TABLE 4.5 COMPARISON OF CHARACTERISTICS OF PMMA AND PMMA/CNT NANOCOMPOSITE FOAMS PREPARED AT 120°C AND 13.8 MPA BY MASP USING FCNT ...... 52

TABLE 4.6 FULL FACTORIAL DESIGN GENERATED BY DESIGN EXPERT ...... 53

TABLE 5.1 TENSILE TEST DATA OF BULK SAMPLES ...... 61

TABLE 5.2 TENSILE TEST DATA OF FOAM SAMPLES ...... 63

viii

LIST OF FIGURES

FIGURE 2.1 DIFFERENT NANOPARTICLES ...... 4

FIGURE 2.2 CARBON NANOTUBES...... 5

FIGURE 2.3 TYPICAL SURFACE FUNCTIONALIZATION OF CNTS ...... 7

FIGURE 2.4 TYPICAL INTERACTION OF PYRENE-CNT ...... 8

FIGURE 2.5 MICROCELLULAR FOAMING PROCESS ...... 13

FIGURE 2.6 SCHEMATIC PRESSURE-TEMPERATURE PHASE DIAGRAM FOR A PURE COMPONENT SHOWING THE SUPERCRITICAL FLUID (SCF) REGION ...... 17

FIGURE 2.7 TENSILE MODULUS OF EXTRUDED PS AND PS/CNF NANOCOMPOSITE FOAMS ...... 24

FIGURE 2.8 COMPRESSIVE MODULUS OF PS AND PS/CNF NANOCOMPOSITE FOAMS ...... 25

FIGURE 3.1 SOLVENT CASTING METHOD ...... 28

FIGURE 3.2 SCHEMATICS OF BATCH FOAMING SETUP ...... 29

FIGURE 3.3 PHOTO OF IN-HOUSE BATCH FOAMING SETUP ...... 30

FIGURE 4.1 MORPHOLOGY OF PURE PMMA FOAM ...... 32

FIGURE 4.2 COMPARISON OF CNTS DISPERSION BY SOLVENT CASTING ...... 34

FIGURE 4.3 FTIR SPECTRA OF (A) PRISTINE AND (B) FUNCTIONALIZED CNTS ...... 35

FIGURE 4.4 MORPHOLOGY OF CNTS BEFORE ((A) AND (B) AND AFTER ((C) AND (D) ACID TREATMENT ...... 36

FIGURE 4.5 DISPERSION AND DISTRIBUTION OF FCNT IN PMMA BY ASP ...... 37

FIGURE 4.6 EFFECT OF PCNT CONCENTRATION ON PMMA FOAMS ...... 38

FIGURE 4.7 MORPHOLOGY OF PMMA-ASP-PCNT-1% NANOCOMPOSITE FOAMS PREPARED AT 13.8 MPA AND DIFFERENT TEMPERATURES ...... 39

ix

FIGURE 4.8 SCHEMATICS OF POLYMER-GAS-PARTICLE INTERFACE ...... 40

FIGURE 4.9 FOAM MORPHOLOGY OF PMMA AND NANOCOMPOSITE FOAMS PREPARED AT 120 °C AND 13.8 MPA ...... 42

FIGURE 4.10 FOAM MORPHOLOGY OF PMMA AND NANOCOMPOSITE FOAMS PREPARED AT 120 °C AND 16.5 MPA ...... 43

FIGURE 4.11 FOAM MORPHOLOGY OF PMMA AND NANOCOMPOSITE FOAMS PREPARED AT 100 °C AND 16.5 MPA ...... 43

FIGURE 4.12 COMPARISON OF NORMALIZED CELL SIZE AND CELL DENSITY OF PMMA AND PMMA NANOCOMPOSITE FOAMS AT A SERIES OF FOAMING CONDITIONS ...... 46

FIGURE 4.13 COMPARISON OF STABILITY OF NANOCOMPOSITE SUSPENSION VIA ASP (LEFT) AND MASP (RIGHT) ...... 47

FIGURE 4.14 SEM OF PMMA-MASP-PCNT-1% NANOCOMPOSITE (A) AND NANOCOMPOSITE FOAMS (B) PREPARED AT 120°C AND 13.8 MPA ...... 48

FIGURE 4.15 SEM MICROGRAPHS SHOWING UNIFORM CNTS DISPERSION VIA COMBINATION OF SURFACE FUNCTIONALIZATION AND MASP ...... 50

FIGURE 4.16 MORPHOLOGY OF PMMA FOAMS PREPARED AT 120 °C AND 13.8 MPA ...... 51

FIGURE 4.17 ONE FACTOR PLOT ON THE SMALL CELLS ...... 54

FIGURE 4.18 ONE FACTOR PLOT ON THE LARGE CELLS ...... 55

FIGURE 5.1 NANOCOMPOSITE FOR TENSILE PROPERTY MEASUREMENT ...... 58

FIGURE 5.2 INSTRON TENSILE TESTING SETUP ...... 59

FIGURE 5.3 TENSILE SAMPLE FOAMING MOLD ...... 60

FIGURE 5.4 TENSILE PROPERTIES OF BULK SAMPLES ...... 62

FIGURE 5.5 TENSILE PROPERTIES OF FOAM SAMPLES ...... 64

x

ABSTRACT

Poly(methyl methacrylate) (PMMA) multi-walled carbon nanotubes (MWCNTs) nanocomposites were synthesized by several methods using both pristine and surface functionalized carbon nanotubes (CNTs). Fourier transform infrared (FTIR) spectroscopy was used to characterize the presence and types of functional groups in functionalized CNTs, while the dispersion of CNTs in PMMA was characterized using scanning electron microscopy (SEM).

The prepared nanocomposites were foamed using carbon dioxide (CO2) as the foaming agent. The cell morphology was observed by SEM, and the cell size and cell density were calculated via image analysis. It was found that both the synthesis methods and CNTs surface functionalization affect the CNTs dispersion in the polymer matrix, which in turn profoundly influences the cell nucleation mechanism and cell morphology. The CNTs are efficient heterogeneous nucleation agents leading to increased cell density at low particle concentrations. A mixed mode of nucleation mechanism was observed in nanocomposite foams in which polymer rich and particle rich region co-exist due to insufficient particle dispersion. This leads to a bimodal cell size distribution. Uniform dispersion of CNTs can be achieved via synergistic combination of improving synthesis methodology and CNT surface functionalization. Foams from these nanocomposites exhibit single modal cell size distribution and remarkably increased cell density and reduced cell size. An increase in cell density of ~70 times and reduction of cell size of ~80% was observed in nanocomposite foam with 1% CNTs.

The effect of CNT surface functionalization on the tensile properties of the PMMA/MWNT nanocomposite and nanocomposite foams were noticed. An increase of ~60% in elastic modulus and ~40% increase in the tensile strength was observed in nanocomposite foam with 0.5 % functionalized CNTs.

xi

CHAPTER 1

INTRODUCTION

1.1 Background and Significance

Polymer nanocomposites are materials with as the matrix phase with a small addition of nanoparticles acting as reinforcement. These materials have drawn a great deal of interest in recent years because they possess high potential of achieving great property improvement. Polymeric foams, on the other hand, represent a group of polymeric materials with gaseous voids in them. These are lightweight materials and have been widely used in various applications such as cushioning, insulation, packaging and absorbency [1]. However the applications of foam are limited due to their inferior mechanical strength, poor surface quality, and low thermal and dimensional stability.

Microcellular foams with cell sizes less than 10µm and cell densities larger than 109 cells/cm3 have been found recently opening new dimensions in achieving light weight materials with excellent mechanical properties [2-6]. However a very narrow operation window [7, 8] is available in the production of microcellular foams because a high pressure drop rate and a low foaming temperature is required. Hence nucleating agents are added to serve as nucleating sites facilitating bubble nucleation process. Various nucleants such as talc [9-11], zinc stearate[11-14], calcium carbonate[11] , and calcium stearate[10,11] have been successfully investigated to produce microcellular foams with a high cell density and a uniform cell size distribution. Recently, nanoparticles have been studied as the foaming nucleants as well. Nanoparticles offer unique advantages for controlling both the foam structures and properties compared to their micro-sized counterparts. The extremely small particle size assists with generation of very large number of nucleants with a relatively low particle loading. Furthermore, the nano-scaled dimension, the high aspect ratio, and the large surface area make those particles desirable as reinforcing elements for the cell walls. The uses of spherical and plate-like nanonparticles have been studied by various researchers and foams in micron range with a controlled cell density

1

have been achieved. Carbon nanotubes (CNTs) represent another type of nanoparticle that offer exceptional properties but the use of CNT in polymer foams have not been investigated and is the subject of this research.

This research is the first step to understand the processing-morphology-property relationship in novel polymer CNT nanocomposite foam. A successful understanding of this relationship may lead to innovation of new foam materials with enhanced mechanical properties and added multifunctional characteristic, increasing the applications of foam as structural materials.

1.2 Scope and Objective of the Research

The objectives of this research are: i) Test the hypothesis that CNTs will increase cell density, enhance mechanical properties and add other functionalities to polymeric foams. ii) If the hypothesis holds, quantify the improvements in terms of cell density and mechanical properties.

Poly (methyl methacrylate) (PMMA)/multi-walled carbon nanotubes (MWNTs) nanocomposites were synthesized using solvent casting and anti solvent precipitation. Both pristine and surface functionalized MWNTs were used to prepare the nanocomposites. The nanocomposite foams were prepared in a batch foaming process using environmentally benign

CO2. The scope of this research is to investigate the effect of CNTs and the processing conditions on the foam morphology and properties to establish the process-morphology-property relationship. The following tasks were be conducted,

i) Study the effect of CNT dispersion of foam morphology. ii) Study the effect of temperature and pressure as the processing conditions on the foam morphology. iii) Study the effect of CNT concentration and surface functionalization of the CNTs on the foam morphology. iv) Measure the mechanical properties of the prepared foams.

2

CHAPTER 2

LITERATURE REVIEW

2.1 Polymer CNT Nanocomposites

2.1.1 Introduction

Polymer composites are widely used in areas of automotive, aerospace, construction, and electronic industries because of their improved mechanical (e.g., stiffness, strength) and physical properties over pure polymers. These composites are made using micron-sized particulates and long fibers to reinforce the weak polymer matrices. In recent years polymer nanocomposites have drawn a great deal of interest because of a high potential of achieving property improvement by a small addition of nanoparticles in the polymer matrices. Furthermore, this significant improvement in variety of properties is achieved without sacrificing the lightweight of polymer matrices. The term nanocomposites usually refer to composites in which at least one phase (the phase) possesses ultrafine dimensions (order of few nanometers). Generally three different types of nanoparticles as shown in Figure 2.1[17] are used to make polymer nanocomposites. The first type of nanoparticles has one dimension in the nanometer scale giving them a platelet- like structure. The lateral dimensions may be in the range of several hundred nanometers to microns, while the thickness is usually less than a few nanometers. Clay and layered nanographites are a good example of this type of nanoparticles. Second types of nanoparticles have two of their dimensions in the nanometer scale, while the third is larger. Thus these particles have an elongated structure. Nanotubes and nanofibers belong to this group. The third type of nanoparticles has all three dimensions at the nanometer scale. Some good examples of these kinds of nanoparticles are spherical silica particles, nanocrystals, gold and other metal nanoparticles.

3

Figure 2.1 Different nanoparticles

2.1.2 CNT and CNT Nanocomposites

The discovery of carbon nanotubes [18] has opened a new dimension in the area of nanocomposite research. This tremendous interest in the research community is because of their remarkable mechanical, electrical, magnetic, optical and thermal properties [19].

CNT consist of graphitic sheets, which have been rolled up into a cylindrical shape. The length of CNT is in the size of micrometers with diameter up to 100 nm. CNT form bundles, which are entangled together in the solid state giving rise to a highly complex network. Depending on the factors such as the atomic arrangement, the diameter and length of the tubes, and the morphology the properties of nanotubes changes. The arrangement of the hexagonal rings along the tubular surface determines whether CNT is metallic or semiconducting. Carbon nanotubes (Figure 2.2) [20] exists as single-walled (Figure 2.2a) or multi-walled structures, and multi-walled carbon nanotubes (MWCNTs) are simply composed of concentric single-walled carbon nanotubes (SWCNTs) (Figure 2.2b). CNFs (Carbon nanofibers), on the other hand possess a larger diameter (50-200nm) lying between conventional carbon fibers (5-10µm) and CNTs (1-50 nm). The structure of CNFs is similar to that of MWCNT.

4

(a) (b)

Figure 2.2 Carbon nanotubes

Physical and mechanical properties of carbon nanotubes have been presented in various literatures by theoretical and experimental results. These reports have shown extremely high elastic modulus of carbon nanotubes, greater than 1 TPa [21] (the elastic modulus of diamond is 1.2 TPa) and have reported strengths 10-100 times greater than the strongest steel at a fraction of the weight. A past few years have shown a great deal of effort in the area of polymer CNT nanocomposites, spanning a wide range of polymers. Both thermoset and polymers have been used in making nanocomposite.

Nanotube-epoxy composites have been widely studied. Ajayan et al. [118] prepared nanocomposites using aligned arrays of MWNT within an epoxy resin matrix. The CNT material produced by arch-discharge technique was dispersed in the resin by mechanical mixing. The composites were cut into thin slices and the orientation of the nanotubes was observed. Jin et al. [5] reported a method to fabricate epoxy-based nanocomposites with mechanically aligned CNT. In this method the composites were prepared by casting a suspension of CNT in a solution of a thermoplastic polymer in chloroform. The samples were uniaxially stretched at 100°C and were found to remain elongated after removal of the load at room temperature. Many research groups have been studying the mechanical behavior of the nanotube-based composites. Cooper et al. [15, 16] detected a shift of the Raman 2600 cm-1 band to a lower wavenumber while studying the

5

stress transfer between the nanotubes and the epoxy matrix. The shift indicates that there is a stress transfer and hence reinforcement by the nanotubes. However, other investigators [22, 23] observed nearly constant value of Raman peak in tension.

PMMA/CNT nanocomposites were prepared by ultrasonication and melt blending. The composite films showed exceptional mechanical and electrical properties [24]. Alternatively, nanotube-PMMA composites were prepared by using coagulation method [25]. After mixing the components, the polymer chains were precipitated entrapping the nanotubes and preventing them from rebundling. A variety of hydrocarbon polymers, such as [26, 27], [28], and [29] have been used to make nanocomposites and CNTs have been successfully dispersed in these matrices. Polystyrene composites prepared by solution or shear mixing [30] have shown improved mechanical properties compared to the neat matrix. Barraza et al. [4] dispersed nanotubes in a styrene solution, and the mixture was subjected to polymerization under emulsion conditions. The final composite exhibited solubility in organic solvents, and the electrical resistivity dropped substantially due to the incorporation of the tubes.

2.1.3 Surface Functionalization

CNTs can be better utilized in composites applications if uniform particle dispersion is achieved in the polymer matrix. Furthermore, a strong interfacial bonding is required to obtain an efficient load transfer across the polymer-particle interface. However the strong intrinsic van der Waals attractions keeps the CNTs tightly entangled together and hence a uniform dispersion is not easy to achieve. Additionally the atomically smooth non-reactive surface precludes the formation of interfacial bonding between the CNTs and the polymer matrix. Various studies have revealed that via appropriate surface chemistry design, it is possible to achieve excellent nanotube dispersion as well as improved bonding between nanotube and polymer matrix.

There are two widely applied surface modifications: non-covalent [31] and covalent [32- 34]. Non-covalent functionalization relies on the intermolecular interaction. On the other hand

6

covalent functionalization is achieved by covalent attachment of certain functional groups onto the tube surfaces.

In covalent functionalization, functional groups are introduced to CNTs surface by covalent bonds. Oxidation of CNTs is one of the most commonly used strategy for covalent funtionalization. A typical oxidation (Figure 2.3) [35] of the CNTs is carried out by strong acid such as nitric acid. A variety of functional groups such as hydroxyl, ketone, carboxylic groups are generated through this approach. These groups can be utilized further for functionalization. Covalent functionalization can greatly improve the dispersion and the efficiency of load transfer due to the formation of covalent bonding. However, it creates defects on the carbon nanotube surface, lowering its own strength as a reinforcing material.

Figure 2.3 Typical surface functionalization of CNTs

Noncovalent functionalization on the other hand, relies on strong intermolecular interaction and will not disturb the intrinsic structures of the CNTs. The non-covalent interaction is based on van der Waals forces for π-π stacking, and it is controlled by thermodynamics. Surfactants are used to improve dispersion by adsorption onto the CNT surface, and reduction of

7

CNT-CNT interaction. Another typical approach is the use of polycyclic aromatic moiety such as pyrene [36, 37]. In this case a strong interaction of pyrene-CNT through π-π stacking (Figure 2.4) [38] is established leading to enhanced dispersion.

Figure 2.4 Typical interaction of pyrene-CNT

2.1.4 Nanocomposite Preparation

Generally there are three methods to fabricate polymer nanocomposites. These methods are solvent blending, melt blending and in-situ polymerization. Each method is discussed briefly in sections to follow. In general a well dispersed system is needed for any of above mentioned fabrication method. One of the widely used method is ultrasonication.

Ultrasonication creates localized packets of shear rates around the bundle of nanotubes. But this method is only suitable for very low viscous matrix materials and small volumes because ultrasonic devices have a high impact of energy, but introduce only low portion of shear forces. Also, the effectiveness of the dispersion is limited as the local introduction of the energy

8

leads to a rupture and damage of CNTs reducing the overall aspect ratio [33]. A proper way to apply the sonication technique would be to disperse CNTs in an appropriate solvent (i.e. chloroform, toluene etc), which allows the agglomerates to be separated due to vibrational energy. The suspension can later be mixed with the polymer and the solvent can be evaporated.

2.1.4.1 Solvent Blending

In solvent blending, the polymer is first dissolved in a solvent. This is followed by the addition of nanoparticles into the polymer solution, which is usually mixed together with the aid of ultrasonication. The interactions between the nanotubes and the solvent can be increased by the reduction of the van der Waals force among CNTs, and hence facilitating the polymer chains to diffuse into the spaces among nanotubes. The unique advantage with is method is that it is capable of dispersing CNTs in the polymer matrices while maintaining the aspect ratio of the original nanotubes. Additionally, the anisotropic nanotubes can be oriented with the aid of an electric or magnetic field. However, a major disadvantage related to this method is the large amount of solvent needed, resulting in higher purification costs. This method is especially attractive in preparing water soluble polymer-clay nanocomposites, such as polyvinyl alcohol (PVA) [39], poly ethylene oxide (PEO) [39], poly acrylic acid (PAA) [40]. The method has also been successfully applied with the aid of surfactant and surface modified nanotubes, to produce CNTs filled polymers. Some of these polymers include PVA [41, 42], PS [24], PMMA [43], polyvinylidene fluoride (PVDF) [119].

2.1.4.2 Melt Blending

Melt blending process is carried out by directly mixing the nanoparticles with the molten polymers. The mixing occurs either statistically or under high shear. It can be used to produce polymer composites on a large scale and is already the most widely used technique to synthesize polymer composites.

The most common method is of using a twin-screw extruder to blend the polymers with the particles. The dispersion of the nanoparticles is found to be affected by the screw design and

9

the residence time. However, excessive shear intensity or back mixing causes a poor mixing. This phenomenon is yet not clear. The complexity of the process itself further adds to the challenges of the studies on how the process conditions affect the formation of nanocomposites.

Up to date, a wide variety of nanocomposites has been prepared via melt blending. 6 [44, 45], PS [26, 27], polypropylene (PP) [28, 46], PEO [29, 47], ethylene-vinyl acetate copolymer (EVA) [47] are some made from nanoclay-filled systems. Application of CNTs as fillers using melt blending have challenges with uniform dispersion of nanotubes into the polymer domain due to their high entanglement and strong affinity among tube ropes and bundles. Instead, the combination of solvent blending and melt blending offers a more powerful dispersion capability [48]. In case of CNFs-filled system, the melt blending offers a feasible and low-cost method to incorporate nanoparticles into various polymers, e.g PP [49, 50], poly (ethylene terephthalate) (PET) [51], PC [52], etc.

The simplicity offered by the melt blending to synthesize nanocomposites is not very helpful to achieve a nano-sacled particle dispersion for composite parts composed of a large volume of nano-elements and high molecular weight polymers.

2.1.4.3 In-situ polymerization

In in-situ polymerization, nanoparticle is dispersed in monomer followed by polymerization. The low viscosity of the monomer (as compared to that of polymer melt or polymer solution) makes it easy to break the aggregated particles using either sonication or high shear mixing. Additionally, the diffusion of monomer into the CNTs bundles causes the expansion of the interparticle space due to the growth of the polymer chains. In addition with an appropriate surface chemistry modification, it is possible to grow polymer chains directly from the nanoparticles surface leading to a significant improvement in the dispersion and interfacial bonding between the particles and polymer.

Incorporation of nanoparticles in thermosets such as epoxy, in situ polymerization is the only viable method [53, 54]. For such as poly (methyl methacrylate) (PMMA)

10

[55, 56], [57], PS [58, 59] the particle dispersion and the wet ability between the particles and polymers has been improved by in-situ polymerization.

2.2 Polymeric Foam

2.2.1 Polymeric Foams and the Foaming Process

Polymeric foams [60, 61] are lightweight structures with a gas phase dispersed in the form of bubbles. They have been widely used in various applications such as cushioning, insulation, packaging and absorbency. Foams with interconnected pore structures are being studied recently for their applications in tissue engineering as scaffolds for cell attachment and growth.

Various polymers have been used for foam applications, e.g., (PU), polystyrene (PS), polyolefin (polyethylene (PE) and polypropylene (PP)), poly() (PVC), (PC), just name a few. In US market PU occupies the largest market share (53%) in terms of the amount consumed, while PS is the second (26%) [1].

Polymeric foams can be classified depending on their composition, cell morphology and physical properties into two categories, rigid or flexible foams. Rigid foams are used in applications such as building insulation, appliances, transportation, packaging, furniture, flotation and cushion, and food and drink containers, whereas flexible foams are used as furniture, transportation, bedding, carpet underlay, textile, gaskets, sports applications, shock and sound attenuation, and shoes.

Based on the size of the foam cells, polymer foams are classified as macrocellular (>100µm), microcellular (1-100 µm), ultramicrocellular (0.1-1 µm) and nanocellular (0.1- 100nm).

Polymer foams can also be defined as either closed cell or open cell foams. A closed cell has the foam cells isolated from each other by complete cell walls. Whereas, in open cell foams,

11

cell walls are broken and the structure consists of ribs and struts. Generally, closed cell foams have lower permeability, leading to better insulation properties. Open cell foams, on the other hand, provide better absorptive capability.

The foaming process consists of a system composing polymer (or monomer), blowing agents, nucleating agent, and other necessary additives (fire retardants, surfactant, catalyst etc). Blowing agent plays a vital role in the foam cell morphology.

Typically there are two types of blowing agents: physical blowing agents and chemical blowing agents. Chemical blowing agents produce gases by chemical reactions or thermal decomposition which are trapped within the polymer matrix to form foams. Physical blowing agents consists of volatile chemicals such as chloroflurocarbons (CFCs), hydrocarbons/alcohols, and inert gases (CO2, N2, argon). Current concerns with the ozone layer depletion has gradually reduced the use of CFCs. Inert gases especially CO2 has become a favorable choice due to its environmentally benign and supercritical fluid working properties.

2.2.2 Microcellular Foams

Plastic foams with cell sizes smaller than 10 µm and cell densities larger than 109 cells/cm3 are defined as microcellular foams [62, 63]. Nam Suh [64] was the first who proposed an idea of introducing small bubbles in solid polymers. The rationale is that if the cell size is smaller than the critical flaws, which already exist in the bulk polymer matrix and is generally introduced in sufficient numbers, then the material density could be reduced while maintaining the essential mechanical properties. Microcellular foams compared to conventional polymeric foams offer higher impact strength, increased toughness and longer fatigue life [8, 9, 62, 65, 66, 67].

Extensive research has been carried out in this area during the past several decades. A wide range of polymers such as PS [62, 64], PC [68], and PMMA [69, 70] have been successfully synthesized into microcellular parts.

12

2.2.2.1 Synthesis of Microcellular Foams

Microcellular foams can be produced by a batching, semi-continuous, and continuous process. Each process mentioned has three basic steps: mixing/saturation, cell nucleation and cell growth as shown in Figure 2.5 [71].

Figure 2.5 Microcellular foaming process

The batch foaming process [68, 72] of polymer materials is carried out by placing the polymer samples in a pressurized autoclave and saturating it with the blowing agent at certain saturation temperature and saturation pressure. If the temperature at which the polymer is saturated is higher than the glass transition temperature, Tg, of the polymer matrix, sudden release of pressure would result in super- saturation and cell nucleation and growth. Cell nucleation is usually fixed by cooling the materials below its Tg. However when the saturation temperature is

13

lower than Tg, the cell is not able to nucleate and grow after the release of pressure even if the gas is in the super saturation state. This is because of the glassy nature (high rigidity) of the polymer matrix. An increase in temperature above the Tg can cause foaming. Cell structure is again fixed by cooling. The latter method allows an independent manipulation of saturation and foaming condition, leading to higher process flexibility. However, diffusion of the gas is inevitable while transferring the gas-saturated material to the high temperature environment, leading to thick skin region.

Kumar et al. [73] developed the semi-continuous foaming process. It was used to produce polymer sheets in a solid state. In this method, a gas channeling material (gas permeable materials) is rolled by interleaving them between layered polymer sheets. Subsequently, the roll is saturated with the blowing agent at room temperature. Finally, the pressure is released and the saturated polymer sheets are separated from the channeling material. The bubble nucleation and growth is induced by pulling the sheets through a heating station.

The continuous extrusion foaming process is a attractive method because of its mass production features of foamed polymers leading to high productivity, easy control, and flexible product shaping [63, 74]. Extrusion foaming can be carried out on either a single-screw extruder or twin-screw extruder. During the extrusion, it is better to reduce the temperature profile from the hopper to the die. A homogeneous single-phase solution is achieved by mixing the blowing agent into the barrel with the polymer. Cell nucleation is induced by either a rapid, large pressure drop or a sudden temperature increase through the die. Cells will expand until the extrudate temperature is below the glass transition temperature of the polymer. The foam shape and expansion sis controlled by a shaping die. The two distinctive characteristics of extrusion foaming compared to a batch foaming process is that instead of saturated amount of gas a metered amount of gas is mixed with the polymer. Secondly, the driving force for bubble nucleation is controlled by the flow instead of the saturation pressure.

14

2.2.2.2 Microcellular Foam Properties

Many polymers have been synthesized as microcellular foams. However, very limited development has taken to understand their mechanical properties.

A brief status on the previous research on the mechanical properties of microcellular foams can be summarized as follows. In case of most polymers, microcellular foams exhibited superior impact strength, toughness and fatigue life compared to solid polymers. The extent of improvement differs among different polymers. Further, different research groups have reported different results for the same polymer-gas system. To conclude, a direct comparison of the mechanical properties between microcellular foams and macrocellular foams with the same density is very limited. The review to follow is focused on the impact, tensile, and compressive properties.

The microcellular foams prepared from PVC [75-77] and PC [3] showed an improvement in their impact strength. A void fraction of 80% increased the impact strength of PVC foams by four times compared to solid PVC [75]. Barlow et al. work on impact strength of PVC reported the strength to be a strong function of both the cell density and cell size [3]. There are some controversial results as well such as from Kumar et al. [78] reporting lowered impact strength by introducing microcellular structure in PVC compared to that of neat PVC. The reason is yet not clear.

Tensile strength and modulus of microcellular foams were studied for PC, ABS, PET and PVC [75, 79, 80]. Though not much improvement in these two properties is seen in microcellular foams over their bulk counterparts a marginal increase in the relative tensile strength is noticed. It was noticed that a linear relationship exists between the tensile strength and the foam density for the polymer systems examined. Waldman [79] reported a 400% increase in toughness of PS foams compared to solid samples. Additionally, the tensile toughness peaked at a relative foam density of 0.75. Arora [2] carried out a systematic study of the compressive behavior of microcellular PS foam. An anisotropic model was proposed to describe the effect of cell size and cell shape on the

15

compressive strength. It was reported that the compressive strength of PS foams increases as the size of the cell increases. The development of a stable neck in the polymer while subjected to a uniaxial tension correlated the phenomenon of heterogeneous, progressive buckling of the microcellular structures. From an energy balance consideration, a model was established describing the densification process of microcellular foams under compression. The fatigue life characterizing the behavior of materials under repeating external forces were studied in case of foams. PC foams with a relative density of 0.9 (10% of the weight of reduction) showed the same fatigue life as that of the PC solid. Furthermore, the PC foams exhibited a fatigue life one order of magnitude higher than that of solid with an increase of relative density to 0.97[72].

2.3 Supercritical CO2 (scCO2) in Foaming

Carbon dioxide is a clean and versatile solvent for the synthesis and processing of a wide range of materials. Supercritical (scCO2) as a processing fluid has made noticeable developments in the past decade and have been extensively used in a variety of applications such as polymerization, polymer fractionation and extraction, impregnation, polymer foaming and blending, surface modification, coating and microlithography [14, 81]. A supercritical fluid (SCF) as seen in Figure 2.6 [14] may be defined as a substance for which both temperature and pressure are both above the critical values. Under supercritical conditions the SCF exhibits gas like diffusivity and liquid like density with zero surface tension. The high solvation power and fast diffusion are especially beneficial to polymer processing and there is a great deal of research in using scCO2 in polymer processing and foaming technology. Additionally, the critical point of

CO2 is relatively low, 31° C and a pressure of 73.8 bar. Furthermore, CO2 is abundantly available at low cost; they are not-toxic, non-flammable, and environmentally benign. All these advantages make ScCO2 a promising blowing agent for polymeric foaming production.

16

Figure 2.6 Schematic pressure-temperature phase diagram for a pure component showing the supercritical fluid (SCF) region .

2.4 Fundamentals in Physical Foaming Process

The gas foaming process involves several important steps namely, formation of polymer- foaming agent homogeneous solution, cell nucleation, cell growth, and cell stabilization [106]. Each step is described in detail in sections below.

2.4.1 Formation of Polymer/Foaming Agent Homogeneous Solution

Formation of gas/polymer solution is one of the fundamental steps of the gas foaming process. Solubility and diffusivity are the two important factors that describe the gas absorption behavior into polymers. Solubility denotes the maximum concentration of the gas in the polymer which can be described by Henry’s law as,

C=H.P (1) where, P is the pressure, C is solubility of gas in the polymer and H is the Henry constant, which is dependent on the temperature. While diffusivity denotes how fast the gas can enter or disperse out of the polymer. The diffusivity can be described by Arrhenius relationship as,

17

E D D exp a (2) 0 RT

Where D is the diffusivity, D0 is the diffusion constant, Ea is the activation energy for diffusion of a gas in a polymer, R is the gas constant, and T is the absolute temperature. An ideal foaming condition is a condition with higher gas solubility in a polymer assisting in greater cell nucleation and growth. A higher diffusivity is sought in this step because of a shorten saturation time and better productivity. However, this may not assist in cell growth, discussed latter.

Both the solubility and diffusivity are highly dependent on the pressure and temperature. A lower temperature generally results in a higher solubility, a highly desirable situation. However, a decreased processing temperature decreases the diffusivity of gas in polymer reducing the productivity. In order to improve the productivity, a higher gas pressure is usually used thereby increasing the solubility. Wissinger et al [82, 83] and Zhang [84] reported that in a

PS-CO2 system there is a linear relationship between the solubility and the saturation pressure

(Henry’s law). Similar results were noticed in the PP-CO2 system. Handa et al [85] investigated 0 the solubility of CO2 in PMMA over a wide range of temperature (0-167 C) and pressure up to 61 atm. They reported that the linear relationship between the solubility and pressure only exists at high temperature regions. However, at lower temperature, the solubility was convex towards the pressure.

Gas solubility being affected by various other factors has been reported in recent research studies. Effect of nanoclay on the kinetics of CO2 gas in PMMA was studied by A.Manninen et al. [86]. It was reported that diffusivity increased with a higher nanoclay concentration while the solubility remained unchanged by the presence of nanoclay. Handa et al. [70, 85] reported that the diffusivity of highly pressurized CO2 in PMMA at a lower temperature may be higher than that at a higher temperature because of the shifting of the glass transition temperature (Tg). The change in crystallinity of semi-crystalline polymer was also found to change the solubility of gas in a polymer [87].

18

Polymers with electron donor groups such as ether, fluro, and carbonyl groups, usually exhibits a higher solubility of CO2. Kazarian et al [88] have shown that CO2 can participate in Lewis acid-base type interactions with polymers containing electron-donating groups such as carbonyls. In this case, CO2 is considered as Lewis acid and the polymer with those functional groups as the Lewis base.

2.4.2 Cell Nucleation

Formation of a gas/polymer solution is followed by a rapid drop in pressure and/or a increase in temperature. According to the Henry’s law the solubility decreases during this process. The resulting over saturation induces large number of cell nucleation’s because the gas tends to escape out of the polymer matrix. The morphology of the final foam product is determined by the cell formation in a polymer and hence cell nucleation is of great importance in the foaming process.

Classical nucleation [11] theory is commonly adopted to explain the nucleation process. The theory classifies the cell nucleation into two different types: homogeneous nucleation and heterogeneous nucleation. Homogeneous nucleation occurs in a pure gas/polymer solution. There are no additional impurities added to the solution. The rate of homogeneous nucleation is expressed as,

* Nhom Cf 00 exp( Ghom / kT) (3) where, f is the frequency factor for homogeneous nucleation a function of both the surface 0 tension and the mass of the gas molecule, C is the concentration of gas molecules, G* is the 0 hom free energy required for the homogeneous nucleation to form a nucleus with critical size, is the Boltzmann’s constant, T is the temperature in Kelvin. The critical nucleation energy is expressed as,

* 16 3 Ghom 2 bp (4) (3 P)

19

and the corresponding critical bubble size is,

2 r * (5) P

Here, is the liquid-gas surface tension, and P is the pressure difference between that of the inside of critical nuclei and the surrounding liquid. Assuming the polymer is fully saturated by the blowing agent and the partial molar volume of blowing agent in the polymer is zero, P can be taken as the saturation pressure.

In the presence of nucleating agents, heterogeneous nucleation takes place in the polymer matrix. It occurs at the interface between the polymer/gas solution and the nucleants. The heterogeneous nucleation rate is given by [11, 12]:

* Nhet Cf 11 exp( Ghet / kT) (6)

Where, f is the frequency factor, C is the concentration of the heterogeneous nucleation sites, 1 1 * which can be related to the particle concentration. The term Ghet is given by,

16 G* 3 f )( (7) het (3 P)2 bp

P where, bp is the surface energy of the gas bubble polymer interface, is gas saturation pressure, and f )( is wetting angle geometric factor.

The homogeneous and heterogeneous nucleations are not different from each other. The mixed model describes the nucleation by,

' N Nhom Nhet (8)

20

where, N is the combined nucleation rate of both homogeneous and heterogeneous nucleation’s,

' N hom is the modified homogeneous nucleation rate, and N het is the heterogeneous nucleation

' rate. Modified homogeneous nucleation rate N hom can be given by,

G N ' f C ' exp hom (9) hom 00 kT

' where C0 is the concentration of gas molecules in solution after heterogeneous nucleation has occurred.

The free energy required for heterogeneous nucleation is generally much lower than that required for homogeneous nucleation. Therefore, additives such as talc, nano-clay or nanotubes can decrease the energy required to create bubbles and therefore promote the cell nucleation. However there are certain criteria to be fulfilled for being an ideal nucleant [89]. Three of the most important criterion are: first, highest nucleation efficiency can only be achieved when the nucleation on the nucleant surface is energetically favored and is relative to homogeneous and heterogeneous nucleation; secondly, ideal nucleants have uniform size and surface properties; thirdly, ideal nucleants are easily dispersible.

2.4.3 Cell Growth and Cell Stabilization

The process of cell growth involves mass, momentum and heat transfer of the fluid. The models describing the cell growth evolve from a basic model [90] used to describe the cell growth from a single bubble that is surrounded by an infinite sea of fluid with an infinite amount of available gas.

Cells come to close to each other as they grow. A solid wall of polymer separates the gaseous phase. The increased pressure inside the bubbles stretches the cell walls to become thinner. Ones the pressure inside a cell is high enough it ruptures the cell wall and two adjacent bubble becomes a single large bubble. This transformation is referred to as cell coalescence [91].

21

Cell coalescence adversely affects the cell sizes and hence should be avoided. Decreasing the flexibility of the polymer by cooling down the polymer is common way to prevent cell coalescence. A drop in temperature below the glass transition temperature (Tg) or the crystallization temperature (Tc) fixes the foam morphology.

2.5 Polymer Nanocomposite Foams

Polymer nanocomposite foams are a new class of materials that are lightweight, having high strength added with multifunctionality. This standard of materials can be achieved based on the combination of implementing functional nanoparticles and polymeric foaming technology together in achieving nanocomposite foams. The following discussion briefly discusses the impact of nanoparticles on foam morphology and their impact on foam properties.

2.5.1 Effects of Nanoparticles on Foam Morphology

Nanoclay [92, 93], nanosilica [94] and nano-scaled block copolymers [89, 95] are some of the widely studied nanoparticles in foaming technology in earlier years. A very intimate contact is established between the particles, the polymer matrix and the gas due to their extremely fine dimension. Furthermore, a significantly high effective particle concentration can be achieved at a low nominal particle concentration. These two considerations make nanoparticles potentially highly efficient as nucleants for polymeric foaming. Various studies have reported an improved nucleation efficiency for different polymer/nanoparticle systems at very low particle concentrations, e.g., PP/clay [92], PS/clay [96], Nylon/clay [97], PLA/clay [98], PC/clay [99], PU/clay [6], PVC/clay [100], and PS/CNFs and PS/CNTs [101,102]. Studies on the effect of particle concentration on the foam nucleation were also investigated [100, 103]. The cell density was found to increase linearly with the clay concentration at low clay concentrations, indicating the dominance of heterogeneous nucleation [100]. The effect of particle dispersion and surface chemistry on the foam morphology was studied in detail [17]. It was reported that by proper manipulation of surface chemistry can cause decrease in free energy contributing to the increase in cell density and reduction in cell size

22

[104]. Compared to intercalated clay dispersion the exfoliated dispersion exhibited the smallest cell size and the largest cell density. In addition, once the organoclay surface was grafted with

PMMA, the strong affinity between CO2 and the carbonyl groups in PMMA may reduce the interfacial tension and further promote bubble nucleation. Similarly thermoset polymers have been studied using nanoparticles and it was found in PU nanocomposite foams at 5 wt% clay content that the efficiency of the nanoclay on the size reduction was not as strong as that in thermoplastic foams [105]. Furthermore, the nanoparticles tend to result in a reduction in cell size, because more bubble start to nucleate, there is a less amount of gas available for the bubble growth. In addition, nanoparticles can significantly increase the melt viscosity. Strain induced hardening was observed under elongation, as result of the nanoclay alignment [106]. Both hinder growth and further contribute to the reduction of cell size.

In addition to nanoclay, CNFs and CNTs have also been utilized to prepare foams. A system of PS nanocomposite foams [107] incorporating a small amount of nanoparticles showed a microcellular cell structure. Compared to the pure PS foam, the addition of 1% wt of CNFs leads to an increase of cell density by more than two orders, while the cell size decreases from 20 to 2.64 µm. CNTs also exhibit a good nucleation effect however there are greater challenges of dispersing them into the polymer matrix. J.Lee et al. [104] reported heterogeneous nucleation comparing the nucleation efficiencies of CNFs, CNTs and exfoliated nanoclay as listed in table 1. It was observed that the potential nucleation sites with only 0.1 % weight CNTs (1.59 x 1015 #/cm3) is significantly higher than those of 1 % weight CNFs (1.41 x 1012 #/cm3) and 5 % weight MHABS (5.45 x 1013 #/cm3). However due to poor dispersion of CNTs compared to CNFs the measured cell density was lower. The nucleation efficiency, defined by the ratio of the measured cell density to the potential nucleants density is higher in case of CNFs (1.97%) compared to CNTs (9.06 x 10-5 #/cm3).

23

2.5.2 Effects of Nanoparticles on Foam Properties

Various properties including mechanical strengths, barrier performance, fire retardancy, dimensional and thermal stabilities, etc. can be effectively improved due to the high aspect ratio and the large surface area of nanoparticles. The nanometer dimension of nanoparticle is especially beneficial for reinforcing the foam materials, taking into consideration that the thickness of cells walls is in the micrometer or sub-micrometer regime.

Polymer nanocomposite foams have exhibited substantial improvement in properties compared to neat polymer foams. Lee M et al. [100] reported tensile properties of PVC/3%

Cloisite 30 B nanocomposite foams blown by three parts by weight of CO2. The foams showed 17.9 % increase of tensile strength, 25.9 % increase of bending strength and 250 % increase of the elongation ratio compared to pristine PVC foam. Lee L et al [104] investigated the tensile properties of extruded PS/CNFs nanocomposite foams. Figure 2.7 [104] shows the foams exhibiting a similar foam density of 0.6-0.7 g/cm3, indicating a similar weight reduction compared to the bulk PS. However in the neat PS foam, a weight reduction of 37 % sacrifices the tensile modulus by 40% from 1.26 GPa to 0.74 GPa. Addition of 1 wt% CNFs improved the tensile modulus by 28% from 0.74 GPa to 0.94 GPa. Further addition of fiber of 5 wt% increases the tensile modulus to 1.07 GPa, which is comparable to that of the bulk PS (1.26 GPa).

Figure 2.7 Tensile modulus of extruded PS and PS/CNF nanocomposite foams

24

It was reported that the compressive strength of PS foams increases as the size of the cell increases. As observed in Figure 2.8 [104], in the presence of CNFs (both 1% and 5%), PS foams show a higher compressive modulus than that of the PS solid. The result indicates that integration of CNFs into the PS foams have a great potential to bridge the gap between the lightweight and high strength requirements. The nanofibers can effectively induce the nucleation of a large amount of bubbles, providing a comparable weight reduction to the neat PS foams.

Figure 2.8 Compressive modulus of PS and PS/CNF nanocomposite foams

25

CHAPTER 3

EXPERIMENTAL PROCEDURE

3.1 Materials

The PMMA (Trade name Acrylite) used in the study is from Evonik Industries. Multi- walled carbon nanotubes (MWCNTs) were purchased from Sigma-Aldrich with more than 95% purity. These nanotubes have an inner and outer diameter of 5- 10 nm and 10- 30 nm. The nanotubes have a length ranging from 0.5 to 500 µm. Chloroform (HPLC Grade, EMD Chemicals Inc.) is used as a solvent in the processing of PMMA. The blowing agent used was bone dry grade CO2 (i.e. 99% pure, Air Gas). Concentrated nitric acid and sulphuric acid were purchased from Fisher Scientific.

3.2 Synthesis of PMMA/MWCNTs Nanocomposites

The nanocomposites were prepared using two methods: solvent casting and anti-solvent precipitation (ASP). Both pristine (pCNTs) and functionalized (fCNTs) were used for the nanocomposite synthesis. fCNTs was prepared by oxidization treatment of pCNTs with strong acid. In a typical procedure, 2 g of MWCNT was suspended in a mixture of 150 ml of concentrated HNO3 and 50 ml of concentrated H2SO4 (3:1) and sonicated for 2 hr. After the treatment the MWCNTs are separated by filtration on 0.2 µm PTFE membrane and washed with deionized water to remove all the traces of acid. The resulting CNTs were then dried under vacuum at 80°C.

26

3.2.1 Solvent Casting Method

PMMA and MWCNT nanocomposites were prepared by solvent casting method. Samples containing three different concentrations (1%, 2% and 5%) of MWCNTs with PMMA were prepared and used in foaming. Initially solvent casting technique was used to prepare composite samples. These samples were further used to make rods using Mixing Extruder (LE- 075, Custom Scientific Instruments).

In a typical solvent casting process, 10 g of PMMA pellets was dissolved in 90 g (approx. 60 ml) of Chloroform and mixed on a magnetic stirrer until PMMA is completely dissolved and a homogenous solution is achieved. Three separate samples each with the above mentioned polymer solvent solution were made in separate beakers.100 mg of MWCNT was added to PMMA solution for 1%, 200 mg was added for 2 % and 500 mg was added for 5 %. After mixing the MWCNT for further 10 minutes each sample was sonicated (Misonix Sonicator 3000) for a period of 1 hr. The power output of the sonicator was set to 10 for a 100 % output. The pulse on and pulse off time were 10 sec and 2 sec respectively. The samples were then poured in flat pans and kept in hood for a period of 48 hours for the solvent evaporation. This is followed by vacuum drying at 110 ° C for 24 hours.

The dried sample was extruded using Mixing Extruder (LE-075, Custom Scientific Instruments) at 240° C to make rods for the subsequent foaming. The final sample had a mean diameter of 1.5 mm and a length of approximately 7 inches. The step by step solvent casting method is shown in Figure 3.1.

27

A) Stirrer Mixing B) Sonication C) Casting D) Extruding

Figure 3.1 Solvent casting methods

3.2.2 Anti-solvent Precipitation Method (ASP)

Multi-walled carbon nanotubes (MWCNT) were added in chloroform at a concentration of 1.66 g/ml achieving a homogeneous solution. Next, an appropriate amount of PMMA was added to the solution and mixed on magnetic stirrer until PMMA is completely dissolved. The sample was further sonicated for 1hour. The power output of the sonicator was set to 10 for a 100 % output. The pulse on and pulse off time were 10 sec and 2 sec respectively. The solution was then added drop wise into a large volume of methanol (the solution is continuously stirred using mechanical stirrer) to enable co- precipitation of PMMA and the nanoparticles (~ 10 ml of methanol is used for every 1ml of solution). The solids were filtered from methanol and dried in vacuum at 80° C for 48 hours to ensure complete solvent removal. The process is denoted as ASP. A modified ASP process was conducted in a slightly different manner and denoted as mASP. CNT was suspended and sonicated in chloroform for extended period of time to form a stable suspension. PMMA solution is prepared by in a separate container by dissolving PMMA in chloroform under magnetic stirrer. The PMMA solution is then added to the CNTs suspension and sonication continues for another 20 minutes. The mixture is then added drop-wise to large amount of methanol. The remaining steps are similar to the original ASP.

28

3.3 Preparation of PMMA/MWCNT Nanocomposite Foams

3.3.1 Design and Construction of High Pressure Batch Foaming System

Figure 3.2 [17] and Figure 3.3 show the schematic of high pressure batch foaming system and the in-house designed setup. The setup accessories are designed for a pressure ranges as high as 103 MPa (15000 psi). Temperature for the system is provided by heat tapes having controller with outputs ranging from 0-600 °C (32-1112 °F). The ISCO high pressure syringe pump can supply and maintain a constant pressure upto 34.45 MPa (5000 psi). The setup consists of two pressure vessels that can be utilized and controlled seperately.

Figure 3.2 Schematics of batch foaming setup

29

Figure 3.3 Photo of in-house batch foaming setup

3.3.2 Batch Foaming

Batch foaming was conducted in a high pressure vessel that housed the polymer rod. Constant temperature was maintained by wrapping heat tape around the entire cell controlled by thermocouple and a temperature controller. The temperature accuracy inside the cell was ± 1ºC.

Gaseous CO2 was fed to the cell from and ISCO-high pressure pump at a fixed pressure.

Polymer, MWCNT composite rod with a mean diameter of 1.5 mm and length of approximately 7 inches was placed in the pressure vessel. CO2 was fed to the pressure vessel using an ISCO high pressure syringe pump and maintained at constant pressure. The system was kept at constant temperature (foaming temperature) for 20 hours to allow saturation and equilibrium. The pressure was then rapidly released and the foam cells were fixed by cooling with an ice and water mixture.

30

3.4 Structural and Morphological Characterization

Nanocomposite and nanocomposite foamed samples were fractured in liquid nitrogen and mounted on stubs with carbon tape for Au/Pd sputtering in a Sputter coater. The cylindrical shaped samples were pulse-coated for 2 minutes to have enough coating for the scanning electron microscope (SEM) analysis. Scanning electron micrograph was taken using JEOL (JSM-7401 F) microscope. Images were acquired at an accelerating voltage of 10 kV and a working distance of 8 mm. Image analysis on the SEM micrographs is conducted to obtain the average cell size and cell density, using Image J. Typically, a micrograph showing more than 50 bubbles is chosen, and the number of bubbles, n, in the micrograph is determined by the software. If the area of the micrograph is A cm2 and the magnification factor is M, the cell density can be estimated as,

23 nM 2 N (10) f A

FT-IR spectra of both pCNTs and fCNTs in KBr pellets were recorded by using a Nicolet 4700 FT-IR spectrometer (Thermo Electron, Co., USA). The scan range was 400-4000 cm-1.

31

CHAPTER 4

RESULTS AND DISCUSSION

4.1 Pure PMMA Foams: Effect of Processing Parameters

Pure PMMA extruded using mixing extruder and were foamed in a batch process over a range of pressure and temperature conditions. Figure 4.1 shows the foam morphology of the pure polymer foam at different temperatures and a constant pressure of 13.8 MPa. Closed cell morphology is seen from the micrographs. As temperature increases, the cell size increases and cell density decreases, along with a change of the shape of the bubbles. Under low temperature the bubble possesses a spherical shape, which changes to polygonal as temperature increases.

(a) (b)

(c) (d)

Figure 4.1 Morphology of Pure PMMA foam; (a) 80°C, (b) 100°C, (c) 120°C, (d) 140°C

32

The solubility of CO2 is a strong function of temperature. A high solubility is seen at low temperature, leading to higher nucleation rate and cell density. While nucleation rate determines the cell density, cell size and shape are affected by the growth of cells, which in turn is strongly influenced by rheological properties of polymer-gas system. Under low temperature, which has a higher solubility, the viscosity is more prominent. The viscosity of the system exhibits an exponential decay as temperature increases. Thus cell growth is more favored under higher temperature, leading to cell impingement and coalescence, which result in polygonal cell shape.

4.2 Preparation of Nanocomposites

PMMA MWCNT nanocomposites were prepared by both solvent casting and anti-solvent precipitation (ASP) using pristine nanotubes (pCNTs) and were denoted as PMMA-sc-pCNT-x% and PMMA-asp-pCNT-x%, respectively, in which sc stands for solvent casting, asp stands for anti-solvent precipitation, p indicates pristine CNT, and x denotes the weight concentration of CNT in the polymer. For example, PMMA-asp-pCNT-2% would be the PMMA CNT nanocomposite prepared using anti-solvent precipitation and pristine CNT with a weight concentration of 2 %.

Due to their large surface area and strong van der Waals interaction, CNTs exist as bundles or agglomerates. An efficient dispersion method is of critical importance for achieving good dispersion of CNTs in polymers. High power sonication, one of the most commonly used methods for nanoparticle dispersion, is often employed in the nanocomposite preparation. However, in the solvent casting process, CNT may re-aggregate as solvent evaporated due to thermodynamic reasons to reduce the surface free energy. This process may be slowed down or prevented by kinetic mechanism, provided the timescale to lock down (freeze) the structure is much smaller than that of re-aggregation. Thus, the anti-solvent precipitation process was employed. Instead of allowing solvent to evaporate, small drops of the suspension were added to an anti-solvent (methanol) in which the polymer is insoluble while the original solvent (chloroform) is soluble. It is envisaged that the rapid diffusion of chloroform into methanol leads to fast precipitation of small polymer/CNTs particles and slowing down of CNTs re-aggregation.

33

The effect of processing methods on CNT dispersion is investigated and the results are shown in Figure 4.2. CNTs aggregates are observed in each case. Apparently neither process result in adequate CNT dispersion in the polymer.

(a) (b)

(c) (d)

Figure 4.2 Comparison of CNTs dispersion by solvent casting ((a) and (b) and anti-solvent precipitation ((c) and (d)). Arrows indicate regions with CNTs aggregates. (a), (c) 10,000x; (b), (d) 20,000x

To improve CNTs dispersion, functionalization is routinely conducted to impart functionality to enhance the interaction between CNTs and polymers. In this study surface functionalization is conducted by using mixed acid as described in the experimental section. The functionalized CNTs are denoted as fCNT.

34

The FTIR spectra of both pCNT and fCNT are shown in Figure 4.3. Compared to the featureless spectrum of pristine CNTs, the fCNTs show several peaks that are characteristics of several types of functional groups (hydroxyl, carboxyl and carbonyl) as previously reported [108]. The peak at 1634 cm-1 is assigned to the stretching vibration of the carbonyl group present in carboxylate (COO-). A broad peak at 3435 cm-1 corresponds to the stretching vibration of hydroxyl group (OH).The peak at 1021 cm-1 is assigned to the stretch vibration of C-O group, which may come result from either the carboxyl or hydroxyl group attached to CNTs. A small peak at 2920 cm-1 would result from stretching vibration of corresponds to methyl, consistent with the breaking of sp2 bond to form sp3 bond during acid treatment. The increased polarity and the π-π interactions between the nanotubes and constitutive carbonyl unities of PMMA is expected to promote CNT dispersion [109]. In addition, the favorable interaction between the carbonyl group and CO2 by electron donor acceptor interaction [110], is beneficial for reducing contact angle and heterogeneous nucleation free energy [111].

1021 1634 2849 2920 Transmittance 3435

a

b

3500 3000 2500 2000 1500 1000 Wavenumber (cm-1)

Figure 4.3 FTIR spectra of (a) pristine and (b) functionalized CNTs .

Figure 4.4 shows a comparison of the morphology of both pCNTs and fCNTs. The CNTs possess similar morphology indicating the condition employed did not cause any significant damage or truncation of the CNTs. 35

(a) (b)

(c) (d)

Figure 4.4 Morphology of CNTs before ((a) and (b) and after ((c) and (d) acid treatment; magnification: (a), (c) 5000x; (b), (d) 30,000x

The nanocomposites were prepared using fCNT and ASP. Figure 4.5 shows the SEM micrograph of the nanocomposites. While CNT are better dispersed (the bright spots in the micrograph), the distribution of CNT is not uniform. Both CNT rich and polymer rich (or CNT deficient) regions are present.

36

(a) (b)

Figure 4.5 Dispersion and distribution of fCNT in PMMA by ASP; bright spots are fCNTs, while arrow indicates polymer rich region. Magnification (a) 10,000x and (b) 20,000x

4.3 Nanocomposite Foams: Bimodal Morphology

Both the pure polymer and the nanocomposites were foamed in a batch process over a range of temperature and pressure conditions.

4.3.1 Effect of CNT Concentration

The PMMA/MWCNT nanocomposites with different CNT concentrations (1%, 2%, 5%) were foamed at 120° C and 13.8MPa, and the respected cell morphology are shown in Figure 4.6. While the pure PMMA foam exhibits relatively uniform cell size, the nanocomposite foams show a bimodal cell size distribution in the presence of pCNT: a large number of small cells are interspersed between a small number of big cells. Similar morphology has also been observed in EVA/CNT nanocomposite foams [112]. As pCNT concentration further increases to 5% (Figure 4.6 (d)), cell coalescence and interconnection becomes more prominent. This is similar to the results previously reported [113], where the presence of large number of anisotropic

37

nanoparticles may facilitate directional cell growth, leading to enhanced cell impingements and wall rupture and cell coalescence.

(a) (b)

(c) (d)

Figure 4.6 Effect of pCNT concentration on PMMA foams; (a) 0%; (b) 1%; (c) 2%; (d) 5 %. Foams were prepared at 13.8 MPa and 120°C

To ascertain and confirm the observed bimodal distribution of cell size, the nanocomposite PMMA-asp-pCNT-1% was foamed under 13.8 MPa and a series of temperatures, and the cell morphologies are shown in Figure 4.7. Bimodal cell size distribution is observed in each case, further confirming the observation.

38

(a) (b) (c)

Figure 4.7 Morphology of PMMA-asp-pCNT-1% nanocomposite foams prepared at 13.8 MPa and different temperatures, (a) 100 °C; (b) 120 °C; (c) 140 °C.

The cell density, or number of bubbles per unit volume, is largely determined by the nucleation step in the foaming process. Nucleation is a classical phenomenon and exists in many processes, e.g., vapor condensation, crystallization. During nucleation, molecules overcome an energy barrier and gather together (via local density and energy fluctuation) to form embryos of the new phase. When the sizes of the embryos are smaller than a critical size, an increase of embryo size is accompanied by an increase of free energy. On the other hand, if the size exceeds the critical size, further increase of embryo sizes lead to a reduction in free energy. Thus stable nuclei are generated. Classical nucleation theory [114, 115] was widely used in foaming to illustrate experimental phenomena due to its simplicity . According to the classical nucleation theory, the homogeneous nucleation rate is expressed as [11].

N Cf exp( G* / kT) hom 00 hom (11)

Where f 0 is the frequency factor representing the frequency that gas molecules joining the embryo nucleus, C is the concentration of the gas molecules, G* is the Gibbs free energy 0 hom associated with the formation of a nucleus and is given by

39

* 16 3 Ghom 2 bp (3 P) (12)

Where γbp is the interfacial tension at the gas bubble interface and P is the pressure exerted by the gas on the cell walls.

In the presence of particles, bubbles form at the polymer-particle interface following heterogeneous nucleation mechanism. The nucleation rate is expressed as [11, 12]:

N Cf exp( G* / kT) (13) het 11 het

Where f1 has the similar physical meaning as f0 in homogeneous nucleation; C1 is the concentration of heterogeneous nucleation sites, which is directly related to the particle

* concentration. The Gibbs free energy for heterogeneous nucleation Ghet is given by

* 16 3 Ghet 2 bp f )( (14) (3 P)

The presence of an interface thus greatly reduces the free energy for nucleation leading to an increase in nucleation rate.

Figure 4.8 Schematics of polymer-gas-particle interface

40

The observed bimodal cell morphology may arise from a mixed mode nucleation mechanism in which simultaneous homogeneous and heterogeneous nucleation take place. In the PMMA/CNT nanocomposites, the CNTs are not well dispersed and form large aggregates (Figure 4.2). CNT rich and CNT deficient (polymer rich) region co-exist. In particle rich region, the more energetically favored heterogeneous nucleation dominates and the homogeneous nucleation is substantially suppressed, as both processes compete for gas molecules to form nuclei. On the other hand, in polymer rich region, the fate of dissolved gas molecules is two-fold. Nuclei may form via homogeneous nucleation mechanism, or the dissolved gas may diffuse to the formed nuclei (from the heterogeneous nucleation mechanism) to support the growth. Despite the high nucleation free energy barrier, it is plausible that homogenous nucleation still takes place in polymer rich regime where particles are deficient, but at significantly lower rate.

On the other hand, majority of the dissolved CO2 may serve as reservoir to supply the growth of the limited number of nuclei formed through homogeneous nucleation, rather than forming new nuclei. Hence considerably fewer bubbles grow in the polymer rich region with ample supply of gas molecules, resulting in larger bubble size. On the other hand, in particle rich region simultaneous growth a large number of bubbles leads to substantially smaller bubble size. This leads to cells with much larger size. It should be noted that mixed mode nucleation have been reported in the literature for polymer foams in the presence of micron sized particles [10]. These effects may be more prominent in foaming of nanocomposites because of more technically challenging nature of achieving sufficient dispersion of nanoparticles.

Despite the bimodal cell size distribution result from insufficient CNTs dispersion, increased cell density in nanocomposite foams compared to that of the pure PMMA foam (Figure 4.6) demonstrates the potential of the nucleation efficiency of CNT, provided a strategy for better CNT dispersion can be developed.

41

4.3.2 Effect of CNT Surface Functionalization

To improve CNT dispersion in PMMA, surface functionalization was conducted. The FTIR results (Figure 4.3) clearly suggests the introduction of polar groups enhancing the polymer CNT interaction. As discussed earlier, the enhanced interaction is beneficial for achieving a more uniform CNTs dispersion. Additionally, similar to the previous research on polystyrene nanoclay nanocomposite foams [111], a favorable interaction between the CO2 and the CNTs through the carbonyl group attached to the CNT surface potentially leads to a reduction in heterogeneous nucleation free energy and significantly increases nucleation rate. Thus nanocomposites were prepared with the fCNTs using anti-solvent precipitation. A series of foams were prepared at various processing conditions and the results are shown in Figure 4.9, Figure 4.10 and Figure 4.11. Again, a bimodal cell size distribution was observed in each case.

(a) (b) (c)

Figure 4.9 Foam morphology of PMMA and nanocomposite foams prepared at 120 °C and 13.8 MPa; (a) pure PMMA; (b) PMMA-asp-pCNT-1%; (c) PMMA-asp-fCNT-1%

42

(a) (b) (c)

Figure 4.10 Foam morphology of PMMA and nanocomposite foams prepared at 120 °C and 16.5 MPa; (a) pure PMMA; (b) PMMA-asp-pCNT-1%; (c) PMMA-asp-fCNT-1%

(a) (b) (c)

Figure 4.11 Foam morphology of PMMA and nanocomposite foams prepared at 100 °C and 16.5 MPa; (a) pure PMMA; (b) PMMA-asp-pCNT-1%; (c) PMMA-asp-fCNT-1%

To obtain more quantitative information on the cell morphology, image analysis was conducted to calculate the cell size and cell density of each mode. The results are shown in Table 4.1.

43

Table 4.1 Comparison of cell size and cell density of PMMA CNT nanocomposite foams

Experiment Sample Distribution Cell size 1 Cell density 1 Cell size 2 Cell density condition ID b Type c (µm) (cells/cm3) (µm) 2 (cells/cm3) ID a I A Mono 21.0 8.0x107 N/A N/A B Bi 16.8 1.4x108 33.0 8.0x106 C Bi 12.8 2.7x108 36.8 8.4x105 II A Mono 12.7 3.4x108 N/A N/A B Bi 7.6 7.6x108 18.6 3.4x107 C Bi 6.9 1.9x109 15.2 6.8x106 III A Mono 10.3 5.3x108 N/A N/A B Bi 5.7 2.3x109 17.1 1.5x107 C Bi 4.7 3.8x109 13.8 9.3x106 a: experimental conditions: I: 120 C, 13.8 MPa, II: 120 C, 16.5 MPa; III: 100 C, 16.5 MPa b: sample ID: A: pure PMMA; B: PMMA-asp-pCNT-1; C: PMMA-asp-fCNT-1 c: Mono: mono-size distribution; bi: bimodal distribution

To further examine the role of CNTs, the size and density of model 1 cells are normalized using values of pure PMMA foam at respective foaming conditions as the reference, and the results are shown in Table 4.2, Table 4.3 and Figure 4.12.

Table 4.2 Comparison of Normalized cell size of PMMA and nanocomposite foams

Experimental I II III conditions ID

Sample ID A 1 1 1

B 0.80 0.60 0.55

C 0.62 0.54 0.46

44

Table 4.3 Comparison of Normalized cell density of PMMA and nanocomposite foams

Experimental I II III conditions ID Sample ID A 1 1 1 B 1.7 2.2 4.3 C 3.4 5.6 7.2

Several observations were made on the role of CNT on the reduction of cell size and concurrent increase of cell density. First the nucleation effect is more prominent when solubility of CO2 in the polymer is higher. For example, at 120 C and 13.8 MPa (condition I), samples B and C exhibit .8x and 2.4x increase of cell density compared to the pure PMMA foam. Increasing the saturation pressure from 13.8 MPa (I) to 16.5Mpa (II) leads to 1.2x and 4.6x, both values higher than those at condition I. Further increase of CO2 solubility by lowering temperature from 120 C (II) to 100 C (III) leads to additional enhancement of nucleation efficiency (3.3x for sample B and 6.2x for sample C).

Furthermore the PMMA-fCNT nanocomposite foams consistently show higher cell density and more reduction in cell size than PMMA-pCNT foams. These would be due to the following reasons. First the functionalized CNTs are better dispersed (albeit not uniformly) in PMMA, leading to increase number of potential nucleation sites. Moreover, the favorable interaction between CO2 and the carbonyl group on CNT may result in reduction in the wetting angle and nucleation free energy and higher nucleation efficiency. Similar phenomena were observed in the previous study where induction of carbonyl group on nanoclay surface leads to similar behavior [111].

45

(a)

(b)

Figure 4.12 Comparison of normalized cell size and cell density of PMMA and PMMA nanocomposite foams at a series of foaming conditions .

Regarding the other mode, the cell size of the nanocomposites foams were larger than that of pure PMMA foam, while the cell densities are order of magnitude lower. These are again in good agreement with the hypothesis that the large cells result from homogeneous nucleation

46

mechanism, which has much lower rate due to the much higher nucleation free energy and therefore being significantly suppressed in the presence to nanoparticles.

The fundamental reason for the mixed mode nucleation mechanism that leads to the bimodal cell size distribution is presence of CNTs rich and CNTs deficient regime due to insufficient dispersion of CNTs. To truly utilize the nucleation efficiency of the CNTs, a more effective strategy is required for uniform CNTs dispersion.

4.4 Nanocomposite Foams: Mono-modal Morphology

A modified ASP process (mASP) is employed to further improve dispersion. The major difference between mASP and the original ASP is that CNT is first dispersed in the solvent by sonication followed by addition of the dissolved polymer solution. The low viscosity of solvent assists with improved dispersion of CNT than in polymer solution (original ASP). Once dispersed and mixed with dissolved polymer, the polymer chains were able to wrap around the CNTs and stabilize the dispersion. Similar approaches have been adopted by other researchers [104] to prepare PMMA single walled CNT nanocomposites and good dispersion was reported.

Figure 4.13 Comparison of stability of nanocomposite suspension via ASP (left) and mASP (right). Material is PMMA-pCNT-1%.

47

The uniformity of the CNTs dispersion was significantly improved via the modified process. Figure 4.13 shows the comparison of the stability of PMMA-pCNT-1% suspension prepared by ASP and mASP. The former process resulted in precipitates exhibiting a grayish color indicating possible macroscopic phase separation, which completely settled down the bottom of the vial in a matter of minutes. On the other hand, the nanocomposite suspension from mASP shows a dark black color, and was extremely stable for an extended period of time. Shown in the figure is a sample that was still stable 12 weeks after preparation.

In spite of the great improvement of CNT distribution, the CNT rich region is nevertheless still present. Figure 4.14 shows SEM micrographs of the nanocomposites and the nanocomposite foams. CNTs aggregates were observed in the nanocomposites, and the nanocomposite foam exhibited a bimodal distribution of cell size.

(a) (b)

Figure 4.14 SEM of PMMA-masp-pCNT-1% nanocomposite (a) and nanocomposite foams (b) prepared at 120°C and 13.8 MPa

The cell size and cell density of the foam were calculated via image analysis and compared to those of the nanocomposite foams by the original ASP. The results are show in Table 4.4. Albeit still possessing bimodal distribution, the nanocomposite from mASP process yields foam with slight reduction of cell size and increase in cell density.

48

Table 4.4 Effect of nanocomposites processing methods on the foam morphology

Sample Cell size Cell size 1 Cell density Cell size 2 Cell density distribution (µm) 1 (#/cc3) 2 (#/cc3) (µm) type

Pure PMMA mono 21.0 8.0x107 N/A N/A

PMMA-ASP- bimodal 16.8 1.4x108 33.0 8.0x106 pCNT-1%

PMMA-mASP- bimodal 11.6 1.7x108 24.7 1.7x107 pCNT-1%

CNT dispersion can be improved by either surface functionalization or refinement of nanocomposite synthesis methods; however, neither is adequate to produce a nanocomposite with uniform CNT dispersion. Further improvement of CNT dispersion may result from the combination of the two aspects in nanocomposite preparation. Thus nanocomposites were synthesized using fCNT via mASP processes. Two types of fCNT are utilized: one being treated with acid for 2 hours and the other for 6 hours. Both suspensions are very stable for extended periods of time. The SEM micrographs of the two nanocomposites are shown in Figure 4.15. Uniform dispersion of CNTs was observed for both nanocomposites. More CNTs are observed in nanocomposite in which CNTs undergo 6 hr acid treatment, suggesting possible breakage and shortening of CNTs.

49

(a) (b)

(c) (d)

Figure 4.15 SEM micrographs showing uniform CNTs dispersion via combination of surface functionalization and mASP; (a) and (b): PMMA-mASP-fCNT-1%-2hr, (c) and (d): PMMA- mASP-fCNT-1%-6hr

Both nanocomposites were foamed at 13.8 MPa and 120 °C and the cell morphology is shown in Figure 4.16 and compared with that of pure PMMA foam at the same processing conditions. Both foams show mono sized cell distribution with significantly higher cell density and smaller cell size.

50

(a) (b) (c)

Figure 4.16 Morphology of PMMA foams prepared at 120 °C and 13.8 MPa; (a) pure PMMA; (b) PMMA-mASP-fCNT-1%-2hr; (c) PMMA-mASP-fCNT-1%-6hr

A detailed analysis of the cell size and cell density is shown in Table 4.5. Compared to the pure PMMA foams, the cell size of the nanocomposite foam (PMMA-mASP-fCNT-1%-2hr) exhibit a reduction of cell size by ~44% and increase of cell density by 5.6 times. In spite of the observed nanotube breakage and shorting by prolonged surface functionalization by acid treatment for 6 hours, the CNTs show remarkable nucleation efficiency in the nanocomposite foam. At 1% concentration, the cell size is reduced by ~80% (to ~4.5 µm) and cell density increases by ~70 times to 5.4x109. Microcellular foams with uniform cell size distribution were successfully prepared under rather mild processing condition. The exceptional nucleation efficiency may results from a combination of the following two factors: i) the breakage of CNTs lead to an increase in effective particle concentration and increase in number of nucleation sites; ii) extended functionalization leads to increase of density of the functional groups, which improve both particle dispersion and facilitate interaction with CO2 to reduce heterogeneous nucleation free energy.

51

Table 4.5 Comparison of characteristics of PMMA and PMMA/CNTs nanocomposite foams prepared at 120°C and 13.8 MPa by mASP using fCNT(number in parenthesis show quantitatively the reduction of cell size and increase of cell density)

Sample Cell size distribution Cell size (µm) Cell density (#/cc3) type Pure PMMA mono 21.0 8.0x107 PMMA-mASP-fCNT- mono 11.6 4.5x108 1%-2hr (~ 44 %) (5.6x) PMMA-mASP-fCNT- mono 4.5 5.4x109 1%-6hr (~ 78 %) (67.5x)

4.5 Design and Analysis of Experiments for the Cell Size Comparison

The major goal of the designed experiment is to conduct a statistical analysis procedure of the cell size distribution (bimodal morphology). The objective is to distinguish between the cell size distributions achieved at the polymer rich region as compared to those formed with CNTs as the nucleating agent. The data generated from the image analysis of the SEM images are incorporated to carry out the analysis. Design Expert version 6 is used to conduct the analysis.

4.5.1 Response Selection

Two parameters are selected as response: (1) Small cell size (cell size 1), (2) Large cell size (cell size 2). Based on the discussion in section 4.3 a bimodal cell size distribution is seen in the final foam morphology with the addition of CNTs. This is caused by contrast effect of CNT rich and CNT deficient (polymer rich) regions. The small cell size is generated at CNT rich

52

region whereas the large cell size is at the polymer rich region. Thus the small cell and large cell sizes are considered as the two responses.

4.5.2 Choice of Factors

The factors in this experiment consist of the following quantitative factors: (1) Pressure (13.8 and 16.5 MPa), (2) Temperature (100°C and 120°C) and (3) Formulation (0, +1, -1). The formulation consists of three levels: Pure PMMA (0), Pristine CNT (-1) and Acid treated CNT (+1).

4.5.3 Choice of Experiments

A full factorial 22 x 3 design was employed in this research (Table 4.6). Factors pressure and temperature are set at two levels while formulation is set at three levels as described in section 4.6.2.

Table 4.6 Full factorial design generated by Design Expert

Std Run Block Factor 1 Factor 2 Factor 3 Response 1 Response 2 Pressure Temperature Formulation Small cell size Large cell size (MPa) (°C) (µm) (µm) 5 1 Block 1 13.8 100 -1 5.19 16.67 2 2 Block 1 16.5 100 0 10.34 0 10 3 Block 1 16.5 100 1 4.74 13.78 6 4 Block 1 16.5 100 -1 5.69 17.11 12 5 Block 1 16.5 120 1 6.86 15.18 3 6 Block 1 13.8 120 0 21 0 11 7 Block 1 13.8 120 1 13.01 36.84 9 8 Block 1 13.8 100 1 4.61 32.15 4 9 Block 1 16.5 120 0 12.65 0 7 10 Block 1 13.8 120 -1 16.78 33.03 1 11 Block 1 13.8 100 0 13.8 0 8 12 Block 1 16.5 120 -1 7.55 18.56

53

4.5.4 DOE Analysis Results and Discussion

The effect of individual factors on the responses was analyzed using the one factor plot generated by Design Expert. The one factor plot on the small cells (cell size 1) is as shown in Figure 4.17.

DESIGN-EXPERT Plot One Factor Plot DESIGN-EXPERT Plot One Factor Plot DESIGN-EXPERT Plot One Factor Plot

Small Cell Size 21 Small Cell Size 21 Small Cell Size 21

X = A: Pressure X = B: Temperature X = C: Formulation

Design Points Design Points Design Points 16.0096 16.9025 16.7419 Actual Factors Actual Factors Actual Factors B: Temperature = 100 A: Pressure = 13.8 A: Pressure = 13.8 C: Formulation = -1 C: Formulation = -1 B: Temperature = 100

11.0193 12.805 12.4839 Small Cell Size Size Cell Small Size Cell Small Size Cell Small

6.02891 8.7075 8.22578

1.03855 4.61 3.96771

13.8 16.5 100 120 -1 0 1

A: Pressure B: Temperature C: Formulation (a) (b) (c)

Figure 4.17 One factor plot on the small cells

Statistically, the effect of pressure leading to an increase in CO2 solubility can be seen with a drop in small cell size with an increase in pressure from 13.8 MPa to 16.5 MPa (Figure 4.17a). The increase in temperature from 100 °C to 120 ° C leads to an increase in size of the small cells (Figure 4.17b). This would be attributed to change in viscocity of the polymer causing free motion of the CO2 molecules in the enclosed cells leading to increase in the cell size. However comparing the effect of addition of 1 % CNTs (both pristine and functionalized) to those cell size formed in pure PMMA show smaller cell sizes (Figure 4.17c). This supports the hypothesis set earlier suggesting the formation of small bubbles from the CNTs rich region with the heterogeneous nucleation mechanism the dominant mode of nucleation. Similarly, the one factor plot of the large cells (cell size 2) is shown in Figure 4.18. The large cells seen in the bimodal cell morphology are hypothesized to arise from the polymer rich

54

region initiated by homogeneous nucleation mechanism. The effect of increase in pressure from

13.8 MPa to 16.5 MPa leads to higher solubility of CO2 in PMMA resulting in reduced size of the large cell size (Figure 4.18a). However the increase in temperature from 100°C to 120°C doesn’t cause significant growth in the size of the large cells (Figure 4.18b). This effect supports the hypothesis that the excess CO2 gas surrounding the few formed nuclei act as a reservior leading to their growth into large cells. Finally, in presence of 1 % CNTs (both pristine and functionalized) the cell size are large then the cells formed in pure PMMA (Figure 4.18c). This suggest the formation of large cells in the polymer rich via homogeneous nucleation mechanism.

DESIGN-EXPERT Plot One Factor Plot DESIGN-EXPERT Plot One Factor Plot DESIGN-EXPERT Plot One Factor Plot

Large Cell Size 36.84 Large Cell Size 36.84 Large Cell Size 36.84

X = A: Pressure X = B: Temperature X = C: Formulation

Design Points Design Points Design Points 27.63 27.63 26.4757 Actual Factors Actual Factors Actual Factors B: Temperature = 100 A: Pressure = 13.8 A: Pressure = 13.8 C: Formulation = -1 C: Formulation = -1 B: Temperature = 100

18.42 18.42 16.1115 Large Cell Size Size Cell Large Size Cell Large Size Cell Large

9.21 9.21 5.74719

0 0 -4.61708

13.8 16.5 100 120 -1 0 1

A: Pressure B: Temperature C: Formulation (a) (b) (c)

Figure 4.18 One factor plot on the large cells

4.6 Conclusions

In this study, PMMA MWCNT nanocomposites were synthesized and novel PMMA MWCNTs nanocomposite foams were prepared in a batch foaming process using carbon dioxide as the foaming agent over wide range of foaming conditions. It was observed that CNT has profound influence on the foam morphology. It was observed that the CNTs are efficient nucleation agent and the nucleation efficiency greatly depends on the dispersion of CNT in the polymer matrix. Bimodal cell size distribution is observed in the nanocomposite foams as a

55

result of mixed mode of nucleation, which in turn results from the co-existence of CNT rich and CNT deficient (polymer rich) region due to insufficient CNT dispersion. By appropriate CNT surface functionalization and process improvement in nanocomposite synthesis, uniform CNT dispersion in PMMA is achieved and outstanding nucleation efficiency of CNT is observed in the nanocomposites foams (~80 times increase in cell density in the presence of 1% CNT). Microcellular nanocomposite foams were prepared.

The DOE analysis carried by Design Expert statistical software package further confirmed the nucleation efficiency of CNTs leading to increase in cell density and reduction in cell size. The analysis also provided statistical insights regarding the formation of large bubbles via homogeneous nucleation mechanism confirming the mixed nucleation mechanism due to insufficient dispersion of CNTs.

56

CHAPTER 5

MECHANICAL PROPERTIES OF PMMA/MWCNT NANOCOMPOISTE FOAMS

5.1 Introduction

Microcellular morphologies have been of great interest to researchers in recent years because of their unique microstructure leading to enhanced mechanical properties. These microcellular foams have a great potential for applications such as packaging, insulation, automotive and aircraft industries, and structural components. Nanoparticles have been added to further improve the mechanical properties of the microcellular foams. In recent years, nanoclays and various other nanoparticles [111,113] have been used as reinforcements for polymeric foams. This is achieved due to the fact that the nanoparticles can be completely embedded in the polymer matrix without disrupting its cellular structure. However CNF- or CNT-reinforced polymeric foams have received minimal attention even though initial investigations are reported on CNT/polystyrene [104,107], CNF/PEEK systems [117]. The excellent mechanical properties and high aspect ratios of the carbon nanostructures have shown a great potential both in modifying the foaming processes and to enhance the mechanical properties of the final cellular composites.

5.2 Experimental Procedure

5.2.1 Preparation of PMMA/CNT nanocomposite

The PMMA/CNT nanocomposites were prepared using the modified ASP process as discussed in section 3.2.2. The prepared nanocomposites were molded to final shape and size for measuring the tensile properties. A 0.5 mm thick mold made from aluminum plate was used to

57

make the nanocomposites in dog bone shape. The dimension of the final nanocomposite was determined based on ASTM D 638 standard for specimen thickness less than 4 mm. The nanocomposites were hot pressed in the mold between two stainless steel plates to the final shape (Figure 5.1).

Figure 5.1 Nanocomposite for tensile property measurement

5.2.2 Foaming procedure

The PMMA/CNT nanocomposite foams were produced using the one-step batch process method. The batch foaming was conducted in a high pressure system as shown in Figure 3.2. Samples were placed in respective mold and the molds were placed in the high pressure vessel. CO2 was fed to the pressure vessel using an ISCO high pressure syringe pump and maintained at constant pressure of 20.68 MPa. The system was kept at constant temperature of 120°C for 20 hours to allow CO2 saturation and equilibrium. The pressure was then rapidly released and the foam morphology was fixed by cooling with an ice and water mixture.

58

5.2.3 Tensile testing

The tensile property of the foam was determined using an Instron tensile testing machine (Figure 5.2) and ASTM D638 was used as a guide for testing. Due to long duration of processing a sample, three samples of each kind were tested. Both bulk and foam samples were tested. The dimensions of the sample are detailed in section 5.3.1 based on ASTM standard. The gauge length of 10 mm was set and the samples were tested at a constant speed of 1 mm/min.

Figure 5.2 Instron tensile testing setup

5.3 Results and discussion

5.3.1 Mold design for foaming

Foaming a bulk sample desired for mechanical testing is not a trivial task. One of the challenges is to control the expansion to achieve the final sample to the desired shape and size. To overcome this challenge special mold are designed (Figure 5.3a). For foam samples to be used for tensile property measurement a cylindrical stainless steel rod 3.59 inches in length and 0.67 inches in diameter was cut into half and the desired dog bone shaped mold was cut on one of the flat surface. The length and width of the mold was 2.90 inches and 0.37 inches

59

respectively and the depth of the mold was 1 mm giving a 100 % expansion to the final foam sample. The other half of the rod was mirror finished to create a very smooth surface. Both these semi-circular rods were tapped and bolted to each other with four bolts creating a strong clamping force. This prevents them from being separated during expansion process caused by the enormous force of expansion. The final foam sample is as per the designed shape and dimensions (Figure 5.3b).

(a) (b)

Figure 5.3 Tensile sample foaming mold

5.3.2 Tensile properties

The measured properties of the PMMA/MWNT nanocomposite and nanocomposite foams included the elastic modulus and tensile strength. The detailed tensile properties of the bulk specimens are shown in Table 5.1. For the bulk samples, the modulus of the nanocomposites increased slightly with addition of both 0.5 % and 1 % weight concentration of nanotubes. The nanocomposite PMMA-fCNT-0.5%-6hr showed the best with 7 % increase (Figure 5.4a) in the modulus. Similarly the strength of the bulk samples increased with addition

60

of both 0.5 % and 1 % weight concentration of nanotube. The nanocomposite PMMA-fCNT-1%- 6hr had a 14 % increase (Figure 5.4b) in strength compared to pure PMMA.

Table 5.1 Tensile test data of bulk samples

Sample Modulus (GPa) Tensile strength (MPa)

Pure PMMA 2.94±0.22 53.21±2.36

PMMA-pCNT-1% 3.01±0.17 53.93±1.85

PMMA-fCNT-1% 3.12±0.11 60.71±4.66

PMMA-pCNT-0.5% 2.86±0.11 60.83±0.22

PMMA-fCNT-0.5% 3.14±0.21 55.89±3.65

4 3.14 3.5 2.94 3.01 3.12 2.86 3 2.5 2 1.5 1 0.5 Modulus (GPa) Modulus 0

(a)

61

70 60.71 60.83 60 53.21 53.93 55.89 50 40 30 20 Strength (MPa) Strength 10 0

(b)

Figure 5.4 Tensile properties of bulk samples; (a) modulus (GPa), (b) strength (MPa)

However the PMMA nanocomposite foams showed a contrasting effect with the addition of 0.5 % and 1 % CNTs. The detailed tensile properties of the foamed specimen’s are shown in Table 5.2. Both PMMA-pCNT-1% and PMMA-fCNT-1%-6hr nanocomposites showed a drop in modulus and strength. The modulus decreased by ~ 13 % in both the cases, while the strength dropped by ~ 20 % and ~ 7 % respectively (Figure 5.5a). On the other hand both PMMA-pCNT- 0.5% and PMMA-fCNT-0.5%-6hr showed an increase in modulus and strength of the nanocomposite foams. The nanocomposite PMMA-fCNT-0.5%-6hr showed an increase of ~ 60 % and ~ 38 % (Figure 5.5b) in the modulus and strength compared to the pure PMMA foam. Therefore, the effects of CNTs on the tensile properties were substantially influenced by the CNTs dispersion. While the lower weight concentration (0.5%) of nanotubes with their functionalization had a better chance of dispersion compared to the higher concentration (1%). The CNTs can be very efficient reinforcements provided they are properly dispersed. On the other hand poor dispersion may actually serve as flaws and stress concentrators, leading to a reduction in properties.

62

Table 5.2 Tensile test data of foam samples

Sample Modulus (GPa) Tensile strength (MPa)

Pure PMMA 0.91±0.06 17.83±1.27

PMMA-pCNT-1% 0.79±0.07 14.34±1.80

PMMA-fCNT-1% 0.79±0.15 16.52±2.46

PMMA-pCNT-0.5% 0.95±0.08 18.84±0.32

PMMA-fCNT-0.5% 1.45±0.30 24.94±0.44

2.00 1.80 1.60 1.45 1.40 1.20 0.91 0.95 1.00 0.79 0.79 0.80 0.60 0.40 Modulus (GPa) Modulus 0.20 0.00

(a)

63

30 24.56 25 17.83 16.52 18.84 20 14.337 15 10

Strength (MPa) Strength 5 0

(b)

Figure 5.5 Tensile properties of foam samples; (a) modulus (GPa), (b) strength (MPa)

64

CHAPTER 6

CONCLUSION AND RECOMMENDATIONS

6.1 Conclusions

In this study, PMMA MWCNT nanocomposites were synthesized and novel PMMA MWCNTs nanocomposite foams were prepared in a batch foaming process using carbon dioxide as the foaming agent over wide range of foaming conditions. It was observed that CNT has profound influence on the foam morphology. It was observed that the CNTs are efficient nucleation agent and the nucleation efficiency greatly depends on the dispersion of CNT in the polymer matrix. Bimodal cell size distribution is observed in the nanocomposite foams as a result of mixed mode of nucleation, which in turn results from the co-existence of CNT rich and CNT deficient (polymer rich) region due to insufficient CNT dispersion. By appropriate CNT surface functionalization and process improvement in nanocomposite synthesis, uniform CNT dispersion in PMMA is achieved and outstanding nucleation efficiency of CNT is observed in the nanocomposites foams (~80 times increase in cell density in the presence of 1% CNT). Microcellular nanocomposite foams were prepared.

The tensile properties of the bulk and foam samples were measured. The surface functionalization of CNTs leads to improved dispersion and a probable good bonding between the PMMA and CNT. The effectiveness of this functioalization could be seen with improved tensile properties of both the bulk and foam samples. By adding in only 0.5 % fCNTs the modulus of the foam increased by ~60 % and tensile strength by ~40%. On the other hand addition of 1 % fCNTs showed a slight drop in the modulus and tensile strength of foam samples. Though higher concentration of CNTs should work as better reinforcement to enhance the tensile property, dispersion becomes more challenging with increase in concentration. Furthermore well dispersed CNTs can be very efficient reinforcement; poorly dispersed nanoparticles may actually serve as flaws and stress concentrators, leading to a reduction in properties.

65

6.2 Recommendations

The current study conducted was limited in its range of processing conditions of pressure and temperature. Future work could prepare a detailed study and expand the property and morphology database at a range of pressure and temperature conditions. Furthermore, other polymers such as polystyrene (PS) and polycarbonate (PC) can be studied to prepare CNTs nanocomposite foams. This study can be extended further to create polymer CNTs nanocomposite foams in specialty applications such as EMI shielding, acoustic damping, barrier for gas and UV, optical applications, flame resistance, electrical and thermal applications.

66

REFERENCES

1. Klempner D, Frish KC. Handbook of polymeric foams and foam technology. New York: Oxford University Press; 1991.

2. Arora, K. A., Lesser, A. J., & McCarthy, T. J. Macromolecules 1998, 31(14), 4614-4620.

3. Barlow, C., Kumar, V., Flinn, B., Bordia, R. K., & Weller, J. Journal of Engineering Materials and Technology-Transactions of the Asme 2001, 123(2), 229-233.

4. Barraza, H. J., Pompeo, F., O'Rear, E. A., & Resasco, D. E. Nano Letters 2002, 2(8), 797-802.

5. Jin L.;Bower, C.;Zhou, O. Appl. Phys. Lett 1998, 73, 1197.

6. Cao, X., Lee, L. J., Widya, T., & Macosko, C. Polymer 2005, 46(3), 775-783.

7. Chiou, J. S., Barlow, J. W., & Paul, D. R. Journal of Applied Polymer Science 1985, 30(6), 2633-2642.

8. Collias, D. I., & Baird, D. G. Polymer Engineering and Science 1995, 35(14), 1178-1183.

9. Collias, D. I., & Baird, D. G. Polymer Engineering and Science 1995, 35(14), 1167-1177.

10. Colton, J. S., & Suh, N. P. Polymer Engineering and Science 1987, 27(7), 500-503.

11. Colton, J. S., & Suh, N. P. Polymer Engineering and Science 1987, 27(7), 485-492.

12. Colton, J. S., & Suh, N. P. Polymer Engineering and Science 1987, 27(7), 493-499.

13. Condo, P. D., Sanchez, I. C., Panayiotou, C. G., & Johnston, K. P. Macromolecules 1992, 25(23), 6119-6127.

14. Cooper, A. I. Journal of Materials Chemistry 2000, 10(2), 207-234.

15. Cooper, C. A., & Young, R. J. Journal of Raman Spectroscopy 1999, 30(10), 929-938.

16. Cooper, C. A., Young, R. J., & Halsall, M. Composites Part A-Applied Science and Manufacturing 2001, 32(3-4) 401-411.

67

17. Zeng C, Dissertation, Polymer nanocomposites: synthesis, structure and processing, The Ohio State University, Chemical Engineering; 2003.

18. Iijima, S. Nature 1991, 354(6348), 56-58.

19. Baughman, R.H., A.A. Zakhidov, and W.A. de Heer. Science 2002. 297(5582), 787-792.

20. Bal S., Samal S.S. Bull. Mater. Sci. 2007, Vol. 30, No. 4, 379-386.

21. Treacy MMJ, Ebbesen TW, and Gibson JM. Nature (London) 1996; 381 (6584); 678- 680.

22. Ajayan, P. M., Schadler, L. S., Giannaris, C., & Rubio, A. Advanced Materials 2000, 12(10), 750-+.

23. Schadler, L. S., Giannaris, S. C., & Ajayan, P. M. Applied Physics Letters 1998, 73(26), 3842-3844.

24. Haggenmueller, R., Gommans, H. H., Rinzler, A. G., Fischer, J. E., & Winey, K. I. Chemical Physics Letters 2000, 330(3-4), 219-225.

25. Du, F.; Fischer, J. E.; Winey, K. I. J. Polym. Sci., Part B 2003, 41, 3333.

26. Vaia R, Dissertation, A polymer melt intercalation in mica-type layered silicates. Cornell University, Department of Materials Science and Engineering; 1995.

27. Vaia RA, Giannelis EP. Macromolecules 1997; 30(25):8000.

28. Svoboda P, Zeng C, Wang H. Lee LJ, and Tomasko DL. Journal of Applied Polymer Science 2002; 85, 1562-1570.

29. Vaia RA, Sauer BB, Tse OK, Giannelis EP. J Poly Sci: Part B: Polym Phys 1997;35(1):59

30. Andrews, R., Jacques, D., Qian, D. L., & Rantell, T. Accounts of Chemical Research 2002, 35(12), 1008-1017.

31. Ying Z, Du J-H, Bai S, Li F, Liu C, and Cheng H-M. International Journal of Nanoscience 2002; 1 (5 & 6): 425-430.

32. Hill, D. E., Lin, Y., Rao, A. M., Allard, L. F., & Sun, Y. P. Macromolecules 2002, 35(25), 9466-9471.

68

33. Zhu, J., Kim, J. D., Peng, H. Q., Margrave, J. L., Khabashesku, V. N., & Barrera, E. V. Nano Letters 2003, 3(8), 1107-1113.

34. Mitchell CA, Bahr JL, Arepalli S, Tour JM, and Krishnamoorti R. Macromolecules 2002; 35 (23): 8825-8830.

35. Schulte K, Gojny F II, Fiedler B, Sandler J and Bauhofer W (2005). Polymer composites- from nano to macro scale (eds).

36. Zhu, J., Yudasaka, M., Zhang, M. F., & Iijima, S. Journal of Physical Chemistry B 2004, 108(31), 11317-11320.

37. Zhu, J., Yudasaka, M., Zhang, M. F., Kasuya, D., & Iijima, S. Nano Letters 2003, 3(9), 1239-1243.

38. Chen, R.J.; Zhang, Y.; Wang, D.; Dai, H.J. Am. Chem. Soc. 2001, 123, 3838.

39. Ogata N, Kawakage S, Ogihara T. J. Appl. Polym. Sci. 1997; 66(3):573-581.

40. Severini F, Formaro L, and Pegoraro M. Carbon 2002; 40 735.

41. Shaffer MSP and Windle AH. Adv. Mater. (Weinheim, Ger.) 1999; 11(11):937-941.

42. Lee BO, Woo WJ, and Kim MS. Marcomolecular Materials and Engineering 2001; 286(2): 114-118.

43. Safadi D, Andrews R, and Grulke EA. J. Appl. Polym. Sci. 2002; 84(14): 2660-2669.

44. Liu L, Qi Z, and Zhu X. Journal of Applied Polymer Science 1999; 71 1133-1138.

45. Lincoln DM, Vaia RA, Wang ZG, and Hsiao BS. Polymer 2001; 42 1621-1631.

46. Kato M, Usuki A, and Okada A. Journal of Applied Polymer Science 1997; 66 1781- 1785.

47. Vaia RA, Vasudevan S, Krawiec W, Scanlon LG, and Giannelix EP. Adv. Mater. (Weinheim, Ger.) 1995; 7(2): 154-156.

48. Cooper CA, Ravish D, Lips D, Mayer J, and Wagner HD. Composites Science and Technology 2002; 62 (7-8): 1105-1112.

49. Lozano K, Ph.D. Dissertation, Development and characterization of a nanofibers reinforced thermoplastic composite. Rice University, Department of Materials Scinence; 1999.

69

50. Lozano K and Barrera EV. J. Appl. Polym. Sci. 2000; 79(1): 125-133.

51. Ma H, Zeng J, Realff ML, Kumar S, and Schiraldi DA. Polymeric Materials Science and Engineering 2002; 86 411-412.

52. Hammel E, Tang X, Trampert M, Schmitt T, Mauthner K, Eder A, and Potschke P. Carbon 2004; 42(5-6): 1153-1158.

53. Sandler, J., Shaffer, M. S. P., Prasse, T., Bauhofer, W., Schulte, K., & Windle, A. H. Polymer 1999, 40(21), 5967-5971.

54. Gong X, Liu J, Baskaran S, Voise RD, and Young JS. Chemistry of Materials 2000; 12(4): 1049-1052.

55. Jia Z, Wang Z, Xu C, Liang J, Wei B, Wu D, and Zhu S. Materials Science and Engineering, A: Structural Materials: Properties, Microstructure and Processing 1999; A271 (1-2): 395-400.

56. Velasco-Santos C, Martinez-Hernandez AL, Fisher FT, Ruoff R, and Castano VM. Chemistry of Materials 2003; 15 (23): 4470-4475.

57. Park C, Ounaies Z, Watson KA, Crooks RE, Smith J, Lowther SE, Connell JW, Siochi EJ, Harrison JS, and St. Clair TL. Chemical Physics Letter 2002; 364 (3,4): 303-308.

58. Shaffer, M. S. P., & Koziol, K. Chemical Communications 2002, (18), 2074-2075.

59. Zeng C and Lee LJ. Macromolecules 2001, 34(12): 4098-4103.

60. Landrock AH, Handbook of Foams: Types, Properties, Manufacture and Applications.1991, New Jersey: Noyes Publications.

61. Khemani KC. ACS Symp. Ser. 1997.

62. Martini-Vvedensky JE, Suh NP, Waldman FA. US Patent 4,473,665; 1984.

63. Park CB, Suh NP, and Baldwin DF, 1999.

64. Martini JE, Waldman FA, and Suh NP in SPE ANTEC. 1982.

65. Matuana, L. M., Park, C. B., & Balatinecz, J. J. Polymer Engineering and Science 1998, 38(11), 1862-1872.

66. Matuana, L. M., Park, C. B., & Balatinecz, J. J. Cellular Polymers 1998, 17(1), 1-16.

70

67. Seeler, K. A., & Kumar, V. Journal of Reinforced and Composites 1993, 12(3), 359-376.

68. Kumar, V., & Weller, J. Journal of Engineering for Industry-Transactions of the Asme 1994, 116(4), 413-420.

69. Goel, S. K., & Beckman, E. J. Polymer Engineering and Science 1994, 34(14), 1137- 1147.

70. Handa, Y. P., & Zhang, Z. Y. Journal of Polymer Science Part B-Polymer Physics 2000, 38(5), 716-725.

71. S Siripurapu, Dissertation, Blend-and surface-assisted foaming of polymers with supercritical carbon dioxide. North Carolina State University, Department of Chemical Engineering; 2003.

72. Doroudiani, S., Park, C. B., & Kortschot, M. T. Polymer Engineering and Science 1998, 38(7), 1205-1215.

73. Kumar V and Schirmer HG, 1997.

74. Park, C. B., Baldwin, D. F., & Suh, N. P. Polymer Engineering and Science 1995, 35(5), 432-440.

75. Matuana, L. M., Park, C. B., & Balatinecz, J. J. Cellular Polymers 1998, 17(1), 1-16.

76. Juntunen RP, Kumar V, and Weller JE. Journal of Vinyl and Additive Technology 2002; 6 (2): 93.

77. Barlow C, Kumar V, and Flinn B. MD Porous, Cellular and Microcellular Materials, ASME 2001.

78. Kumar V. Annual Technical Conference - Society of Plastics Engineers 2002; 60th (2): 1892.

79. Waldman FA, Thesis, The Processing of Microcellular Foam. MIT, Mech. Eng. Dept. 1982.

80. Kumar V, Vanderwel M, and Well J. J. Eng. Mater. Tech 1994.

81. Kirby, C. F., & McHugh, M. A. Chemical Reviews 1999, 99(2), 565-602.

82. Wissinger, R. G., & Paulaitis, M. E. Journal of Polymer Science Part B-Polymer Physics 1987, 25(12), 2497-2510.

71

83. Wissinger, R. G., & Paulaitis, M. E. Industrial & Engineering Chemistry Research 1991, 30(5), 842-851.

84. Zhang Q, Xanthos M, and Dey SK. MD (American Society of Mechanical Engineers) 1998; 82 (Porous, Cellular and Microcellular Materials) 75-83.

85. Handa, Y. P., Zhang, Z., & Wong, B. Cellular Polymers 2001, 20(1), 1-16.

86. Manninen, A. R., Naguib, H. E., Nawaby, A. V., Liao, X., & Day, M. Cellular Polymers 2005, 24(2), 49-70.

87. Koros, W. J., & Paul, D. R. Journal of Polymer Science Part B-Polymer Physics 1978, 16(11), 1947-1963.

88. Kazarian, S. G., Vincent, M. F., Bright, F. V., Liotta, C. L., & Eckert, C. A. Journal of the American Chemical Society 1996, 118(7), 1729-1736.

89. Peter Spitael, Christopher W. Macosko, Richard B, Mcclurg. Macromolecules, 2004, 37 (18), 6874-6882.

90. Shimoda, M., Tsujimura, I., Tanigaki, M., & Ohshima, M. Journal of Cellular Plastics 2001, 37(6) 517-536.

91. Pop-Iliev R, Liu F, G Liu, Park CB. Advances in Polymer Technology 2003, 22, 4, 280- 296.

92. Nam, P. H., Maiti, P., Okamoto, M., Kotaka, T., Nakayama, T., Takada, M., et al. Polymer Engineering and Science 2002, 42(9), 1907-1918.

93. Ray, S. S., & Okamoto, M Macromolecular Materials and Engineering 2003, 288(12), 936-944.

94. Siripurapu, S., DeSimone, J. M., Khan, S. A., & Spontak, R. J. Advanced Materials 2004, 16(12), 989-+.

95. Siripurapu, S., DeSimone, J. M., Khan, S. A., & Spontak, R. J. Macromolecules 2005, 38(6), 2271-2280.

96. Strauss, W., & D'Souza, N. A. Journal of Cellular Plastics 2004, 40(3) 229-241.

97. Yuan, M., Turng, L., Gong, S., Caulfield, D., Hunt, C., & Spindler, R. (2004). Study of injection molded microcellular -6 nanocomposites

72

98. Ray, S. S., & Okamoto, M. Macromolecular Materials and Engineering 2003, 288(12), 936-944.

99. Mitsunaga, M., Ito, Y., Ray, S. S., Okamoto, M., & Hironaka, K. Macromolecular Materials and Engineering 2003, 288(7), 543-548.

100. Lee M, Lee B-K, Choi K-D. Foam compositions of nanocomposites, PCT Int Appl., Wo, 2004074357;2004.

101. Shen J, Han X, Lee LJ. Society of Plastics Engineers; 2005. p. 1896. ASAP.

102. Shen J, Lee LJ. Society of Plastics Engineers; 2004. p. 1836.

103. Zeng C, Han X, Lee LJ, Koelling KE, Tomasko DL. Society of Plastics Engineers; 2002. p. 1504.

104. Lee, L. J., Zeng, C. C., Cao, X., Han, X. M., Shen, J., & Xu, G. J. Composites Science and Technology 2005, 65(15-16), 2344-2363.

105. Cao X, Lee LJ, Widya T, Macosko C. Polymer 2005; 46(3):775.

106. Okamoto, M., Nam, P. H., Maiti, P., Kotaka, T., Nakayama, T., Takada, M., et al. Nano Letters 2001, 1(9), 503-505.

107. Shen J, Zeng C, Lee LJ. Polymer 2005; 46: 5218-5224.

108. Tasis, D., et al., Chemical Reviews, 2006. 106(3):1105-1136

109. Thomassin, J.M., Lou X., Pagnoulle C., Saib A., Bednarz L., Huynen I., Jerome R., and Detrembleur C.; The journal of Physical Chemistry C, 2007,111, 11186-11192.

110. Kazarian S.G.; Applied Spectroscopy Reviews, 1997, 32, 301-348.

111. Zeng C, Han X, Lee J, Koelling K.W., Tomasko D.L.; Adv. Mater., 2003, 15, 1743- 1747.

112. Park K.W., Kim G.H.; J. Appl. Polym. Sci., 2009, 112, 1845-1849.

113. Han X, Zeng C, Lee LJ, Koelling KW, Tomasko DL. Polym Eng Sci 2003;43: 1261-75.

114. Abraham F.F., in Advances in Theoretical Chemistry, Suppl. 1, Academic Press, New York, 1974.

115. Laaksonen A., Talanquer V., Oxtoby D.W., Annu. Rev. Phys. Chem., 1995, 46, 489-524.

73

116. Kashiwagi T., Fagan J., Douglas J.F., Yamamoto K., Heckert A.N., Leigh S.D., Obrzut J., Du F., Lin-Gibson S., Mu M., Winey K.I., Haggenmueller R.; Polymer, 2007, 48, 4855-4866.

117. Werner P, Verdejo R. Wöllecke F, Altstädt V, Sandler JKW, Shafer MSP, 2005, Adv Mater 17(23):2864.

118. Ajayan, PM., Stephan, O., Colliex, C., Trauth, D. Science 1994, 265(5176), 1212-1214.

119. Lee BO, Woo WJ, Song HS, Jung SW, Lee JC, and Kim MS. Carbon’01, An International Conference on Carbon, Lexington, KY, United States, July 14-19, 2001; 895-896.

74

BIOGRAPHICAL SKETCH

Nemat Jalall Hossieny was born in Kolhapur, Maharashtra, India to Jalall Hossieny Nejad and Rezwan Jalall Hossieny on October 7, 1979. He spent most of his early life at Kolhapur, where he had his primary education at Saint Xavier’s High School and secondary education at Vivekanand Junior College. Nemat had his B.Sc in Mechanical Engineering from D.Y.Patil College of Engineering and Technology affiliated to Shivaji University located in Kolhapur. After completion of his bachelor’s degree, he joined the Baha’i World Centre, a non-profit based organization located in Haifa, Israel and served on a volunteer basis, as an HVAC system engineer for a period of three years and six months. He then joined the Department of Mechanical and Industrial Engineering at Texas A & M University, Kingsville as a graduate student and worked as a Lab Assistant. He decided to move to Florida and joined the Department of Industrial and Manufacturing Engineering at Florida State University (FSU). Since January 2007 he has been working as Research Assistant with the High Performance Materials Institute (HPMI) in the area of polymer nanocomposite foaming technology until completion of his master’s degree. While studying for his master degree at FSU he got married to his life partner, Zhinus Vafai. Nemat intends to proceed for higher education in area of mechanical and industrial engineering.

75