i SYNTHESIS, CHARACTERIZATION AND DIELECTRIC PROPERTIES OF CADMIUM SULFIDE POLYMER NANOCOMPOSITES ERIAWAN RISMANA UNIVERSITI TEKNOLOGI MALAYSIA i SYNTHESIS, CHARACTERIZATION AND DIELECTRIC PROPERTIES OF CADMIUM SULFIDE POLYMER NANOCOMPOSITES ERIAWAN RISMANA A thesis submitted in fulfilment of the requirements for the award of the degree of Doctor of Philosophy (Chemistry) Faculty of Science Universiti Teknologi Malaysia FEBRUARY 2010 iii Specially dedicated to: My mother, father, wife, daughter and sons iv ACKNOWLEDGEMENT In the name of Allah, the most Gracious, the most Merciful Alhamdulillaahirabbil ‘alamin, Allaahumma sholli ala Muhammad wa ‘ala alihi wa shohbihi wa sallam. All praise must be to Allah, The supreme Lord of the world. Peace and blessing to Rasulullaah Muhammad Shollallahu ‘Alaihi Wasallam, all the prophets, his families, his close friends and all Muslims. I wish to thank my supervisors, Prof. Dr. Salasiah Endud and Assoc. Prof. Dr. Hadi Nur for their continual guidance, encouragement and patience. Their understanding and supervision are very much appreciated. A research grant from the Ministry of Science Technology and Innovation Malaysia (MOSTI) is gratefully acknowledged. I also would like to express my gratitude to all lecturers and researchers of the Department of Chemistry for their support and Ibnu Sina Institute for Fundamental Science Studies UTM for the nanocomposites characterizations. Thanks to all the laboratory assistants and friends in the Department of Chemistry for their kindness and wonderful cooperation. Last but not least, I am grateful to my father and mother: H. E. Suparman and Hjh. E. Reswati, my father and mother in law: Suparmo and Purniasih, my wife: Endang, my daughter: Fani, my sons: Faruq and Fathian, and my brothers for their prayer, love, understanding, encouragement and support. v ABSTRACT Semiconductor/polymer nanocomposites are of increasing importance with their tunable properties being used as dielectric materials. This thesis focused on cadmium sulfide (CdS)/polymer nanocomposites. CdS has been combined with three polymer matrices, i.e. poly(styrene-divinylbenzene) [P(S-DVB)], poly(methacrylic acid-ethyleneglycoldimethacrylic acid) [P(MAA-EGDMA)] and sulfonated poly(styrene-divinylbenzene) [SO3H-P(S-DVB)]. CdS/P(S-DVB) nanocomposite was synthesized by in-situ polymerization in a miniemulsion system using monomer as oil-phase. CdS/P(MAA-EGDMA) nanocomposite has been synthesized by ion exchange and precipitation processes. While, the CdS/SO3-P(S-DVB) nanocomposite has been prepared by sulfonation, ion exchange and precipitation. Agglomerated nanoclusters of CdS were obtained from the above in-situ preparation methods. The structure-dielectric property relationship of the nanocomposites is elucidated by various techniques such as UV – Vis, FTIR, UV-Vis DR, TEM, SEM, XRD, impedance analyzer, AAS, EDX, thermal conductivity analyzer and thermogravimetric analysis. Dielectric properties of the CdS/polymer nanocomposites have been studied at frequencies of 0.1 – 1,000 kHz. The decrease in dielectric constant was found in CdS/SO3H-P(S-DVB) nanocomposite. Considering that the SO3H-P(S-DVB) has very high dielectric constant due to its proton mobility, the replacement of the proton of SO3H-P(S-DVB) with the CdS particles caused the decrease in dielectric constant of the nanocomposite. Interestingly, an increase in the dielectric constant was also observed in CdS/P(S-DVB) and CdS/P(MAA-EGDMA) nanocomposites compared to that of CdS nanoparticles or pure polymers. It is also demonstrated that the unusual enhancement of dielectric constant of CdS/P(MAAEGDMA) depended on the concentration of CdS nanoparticles. The occurrence of strong interfacial interaction between CdS nanoparticles with P(MAA-EGDMA) polymer has been proved by FTIR spectra. One explains that an increase in dielectric constant is due to the increase of interfacial interaction among CdS nanoparticles and also CdS nanoparticles with polymer. These interactions increase the mobility of charge carriers and polarizability of electron. Based on the results of this study, it can be suggested that dielectric properties of CdS/polymer nanocomposites can be explained by the following unique properties i.e. nanosize of CdS particles, semiconductor property of CdS, the interfacial interaction between CdS nanoparticles and polymer and intrinsic properties of polymer. These conclusions lay the foundation for developing new synthetic strategies for designing new dielectric materials by varying the size, concentration and distribution of CdS nanoclusters in various polymer matrices. vi ABSTRAK Semikonduktor/polimer menjadi sangat penting kerana sifat – sifatnya yang boleh diselaraskan sebagai bahan dielektrik. Tesis ini memfokuskan kepada nanokomposit kadmium sulfida (CdS)/polimer. CdS telah digabungkan dengan tiga polimer matriks iaitu poli(stirena-divinilbenzena) [P(S-DVB)], poli(asid metakrilikasid etilenaglikoldimetakrilik) [P(MAA-EGDMA)] dan poli(stirena-divinilbenzena) tersulfonat [SO3H-P(S-DVB)]. Nanokomposit CdS/P(S-DVB) telah disintesis melalui kaedah pempolimeran in-situ dalam suatu sistem miniemulsi menggunakan monomer sebagai fasa minyak. Nanokomposit CdS/P(MAA-EGDMA) telah disintesis melalui kaedah penukaran ion dan proses pemendakan. Manakala, nanokomposit CdS/SO3-P(S-DVB) telah dihasilkan secara pengsulfonan, penukaran ion dan pemendakan. Nanokluster aglomerat CdS telah diperolehi daripada kaedah penyediaan in-situ. Hubungan struktur-sifat dielektrik nanokomposit tersebut telah dianalisis menggunakan pelbagai tenik seperti UV – Vis, FTIR, UV – Vis DR, TEM, SEM, XRD, penganalisis impendans, AAS, EDX, penganalisis kekonduksian terma dan analisis termogravimetri. Sifat dielektrik nanokomposit CdS/polimer telah diukur pada frekuensi 0.1 – 1,000 kHz. Pemalar dielektrik bagi nanokomposit CdS/SO3-P(SDVB) didapati telah berkurang. Memandangkan SO3H-P(S-DVB) mempunyai pemalar dielektrik yang sangat tinggi disebabkan oleh mobiliti protonnya, penggantian proton pada SO3H-P(S-DVB) dengan partikel CdS telah mengakibatkan penurunan pemalar dielektrik bagi nanokomposit. Yang lebih menarik, peningkatan pemalar dielektrik juga telah diperhatikan bagi nanokomposit CdS/P(MAAEGDMA) dan CdS/P(S-DVB) berbanding partikel CdS atau polimer tulen. Telah ditunjukkan juha bahawa peningkatan yang luar biasa pemalar dielektrik bagi CdS/P(MAA-EGDMA) bergantung pada kepekatan partikel CdS. Kehadiran interaksi antaramuka yang kuat antara partikel CdS dengan polimer P(MAAEGDMA) telah dibuktikan dengan spektrum FTIR. Dapat dijelaskan bahawa peningkatan pemalar dielektrik adalah disebabkan oleh pertambahan interaksi anataramuka di kalangan partikel – partikel CdS dan juga nanopartikel CdS dengan polimer. Interaksi tersebut telah meningkatkan mobiliti pembawa cas dan kebolehpengutuban elektron. Berdasarkan hasil kajian ini, adalah dicadangkan bahawa sifat dielektrik nanokomposit CdS/polimer boleh diterangkan dengan sifat – sifat unik berikut iaitu, saiz nano partikel CdS, sifat semikonduktor CdS, interaksi antaramuka antara nanopartikel CdS dengan polimer dan sifat intrinsik polimer. Kesimpulan ini menjadi asas kepada strategi sintesis yang baru bagi menghasilkan bahan dielektrik dengan cara mengubah saiz, kepekatan dan taburan nanokluster CdS dalam matriks pelbagai polimer. vii TABLE OF CONTENTS CHAPTER 1 2 TITLE PAGE DECLARATION ii DEDICATION iii ACKNOWLEDGEMENT iv ABSTRACT v ABSTRAK vi TABLE OF CONTENTS vii LIST OF TABLES xi LIST OF FIGURES xiii LIST OF SYMBOL / ABBREVIATIONS xix LIST OF APPENDICES xxii INTRODUCTION 1 1.1 Background of Study 4 1.2 Design and Strategy of Research 7 1.3 Objective of The Study 9 1.4 Scope of The Study 9 LITERATURE REVIEW 12 2.1 Dielectric Properties and Dielectric Constant 12 2.2 Embedded Capacitor 15 2.3 Design of High Dielectric Property Materials 16 2.3.1 Overview of Dielectric Properties of Polymer Nanocomposites 19 viii 2.3.2 CdS/polymer Nanocomposites 2.4 Cadmium Sulfide as Semiconductor Material 2.4.1 Methods of Synthesis of Cadmium Sulfide 25 26 29 Nanoparticles 32 2.5 Miniemulsion Polymerization 2.5.1 Introduction 33 2.5.2 Miniemulsion 34 2.5.3 The Miniemulsion Process 35 2.5.4 Preparation of Miniemulsion 36 2.5.4.1 Formulations 37 2.5.4.2 Method of Preparation 38 2.5.4.3 Homogenization Devices 39 2.5.4.4 Particle Nucleation 40 2.5.5 Applications 42 2.5.5.1 Encapsulation of Inorganic Material 3 42 EXPERIMENTAL 46 3.1 Synthesis and Preparation 46 3.1.1 Purification of Monomer 46 3.1.2 Preparation of CdS Nanoparticles by Reverse 46 Micelles Using n-Decane as Oil-phase 3.1.3 Preparation of CdS Nanoparticles by Reverse 47 Micelles Using Monomer as Oil-phase 3.1.4 Preparation of P(MAA-EGDMA) 3.1.5…Preparation of 48 CdS/P(MAA-EGDMA) 48 Preparation of P(S-DVB), SO3H/P(S- DVB) and 49 Nanocomposites 3.1.6 CdS/SO3H-P(S- DVB) 3.2 Characterization Techniques 50 3.2.1 UV-Vis Spectroscopy 50 3.2.2 X-Ray Diffraction 51 3.2.3 Fourier Transform Infrared Spectroscopy 51 3.2.4 UV-Vis Diffuse Reflectance Spectroscopy 52 ix 4 3.2.5 Field Emission Scanning Electron Microscope 52 3.2.6 Transmission Electron Microscopy 52 3.2.7 Dielectric Constant and Dissipation Factor 53 3.2.8 Ionic Conductivity 53 3.2.9 Thermal Conductivity 53 3.2.10 Thermogravimetric Analysis 54 3.2.11 Atomic Absorption Spectroscopy 54 RESULTS AND DISCUSSION 55 4.1 58 Optimization of Synthesis of CdS Nanoparticles and CdS/P(S-DVB) Nanocomposites 4.1.1 Structural and Morphological Properties of 67 CdS/P(S-DVB) Nanocomposites 4.1.2 Fourier Transform Infrared and UV – Vis Diffuse Reflectance Spectroscopy 71 of CdS/P(S-DVB) Nanocomposites 4.1.3 Dielectric and Electrical Properties of 74 CdS/P(S-DVB) 79 Properties 81 CdS/P(S-DVB) Nanocomposites 4.1.4 Thermal Properties of Nanocomposites 4.2 Synthesis and Physicochemical Cadmium Sulfide-Poly(methacrylic of acid-ethylene glycol dimethacrylic acid) [CdS/P(MAA-EGDMA)] Nanocomposites 4.2.1 Synthesis of P(MAA-EGDMA) and 81 CdS/P(MAA-EGDMA) Nanocomposites 4.2.2 Fourier Transform Infrared and UV – Vis Diffuse Reflectance Spectroscopy of 84 4.2.3 Structural and Morphological Properties of 88 CdS/P(MAA-EGDMA) Nanocomposites CdS/P(MAA-EGDMA) Nanocomposites 4.2.4 Dielectric and Electrical Properties of 92 x CdS/P(MAA-EGDMA) Nanocomposites 4.2.5 Influence of Temperature on Dielectric and Electrical Properties of 102 CdS/P(MAA- EGDMA) Nanocomposites 4.2.6 Thermal Properties of CdS/P(MAA-EGDMA) 104 Nanocomposites 4.3 Synthesis and Physicochemical Properties of 107 4.3.1 Synthesis of P(S-DVB), SO3H-P(S-DVB) and 107 CdS/Sulfonated-Poly(styrene-divinylbenzene) CdS/SO3-P(S-DVB) Nanocomposites 4.3.2 Fourier Transform Infrared and UV – Vis Diffuse Reflectance Spectroscopy 108 of CdS/SO3-P(S-DVB) Nanocomposites 4.3.3 Structural and Morphological Properties of 113 CdS/SO3-P(S-DVB) Nanocomposites 4.3.4 Dielectric and Electrical Properties of 118 4.3.5 Thermal Properties of CdS/SO3-P(S-DVB) 127 CdS/SO3-P(S-DVB) Nanocomposites Nanocomposites 4.4 Comparison of Physicochemical and Electrical 131 Properties of Those CdS/polymer Nanocomposite 5 CONCLUSIONS 136 REFERENCES 138 Appendices A-H 149-156 xi LIST OF TABLES TABLE NO. TITLE PAGE 1.1 Outline of results and discussion 10 2.1 Dielectric constant of several materials 17 2.2 Dielectric constant and method of synthesis of several 23 polymer nanocomposites 2.3 Dielectric constant and methods of synthesis of CdS 26 nanoparticle and CdS/polymer nanocomposites 2.4 Several CdS/polymer nanocomposites, synthesis method, 31 applications and property of interest 2.5 Examples of formulation of miniemulsion polymerization 37 2.6 Somes applications of polymerization miniemulsion 42 4.1 The optimum conditions for synthesis of CdS nanoparticle 59 using n-decane as oil-phase 4.2 The optimum conditions for synthesis of CdS/P(S-DVB) 63 nanocomposite using monomer as oil-phase 4.3 The absorption onset wavelength and particle size of CdS 86 nanoparticles in CdS/P(MAA-EGDMA) nanocomposites with various CdS contents 4.4 Dielectric constants of P(MAA-EGDMA), CdS 93 nanoparticles, CdS/P(MAA-EGDMA) prepared by physical mixing and CdS/P(MAA-EGDMA) nanocomposites at various frequencies 4.5 Dissipation factors of P(MAA-EGDMA), CdS nanoparticles 100 and CdS/P(MAA-EGDMA) nanocomposites 4.6 The absorption onset wavelength and particle size of 110 xii CdS/SO3-P(S-DVB) nanocomposites 4.7 Dielectric constants of P(S-DVB), CdS nanoparticles, SO3H- 119 P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites 4.8 Meq H+/gram of SO3H-P(S-DVB) with degree of sulfonation 120 = 39.10 % and CdS/SO3-P(S-DVB) nanocomposites with various amount of CdS 4.9 Dissipation factors of P(S-DVB), CdS nanoparticles, SO3H- 122 P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites 4.10 Electrical conductivities of CdS nanoparticles, P(S-DVB), 125 Cd(SO3)2-P(S-DVB), NaSO3-P(S-DVB), SO3H-P(S-DVB), and Cd/SO3-P(S-DVB) with various amount of CdS 4.11 Physicochemical properties of CdS/polymer nanocomposites 132 4.12 Electrical properties of CdS/polymer nanocomposites 135 xiii LIST OF FIGURES FIGURE NO. 1.1 TITLE PAGE Schematic representation of embedded capacitance solution for 2 electronic devices 1.2 Schematic representation of the synthesis of CdS/polymer 8 nanocomposites 2.1 Schematic representation of capacitor 13 2.2 Schematic representation of types of polarizations 14 2.3 Frequency dependence of the several contributions to the 15 polarizability 2.4 Schematic changes in the density of states on going from a bulk 28 crystal to a nanocrystal to molecule/atom 2.5 Principle of miniemulsion polymerization 35 2.6 Schematic for monomer miniemulsion preparation methods 39 2.7 Schematic of sonication process 39 2.8 The principle of the encapsulation of inorganic materials in 43 polymer particles by the miniemulsion process 4.1 Schematic representation of mechanism of CdS nanoparticles 57 attachment and encapsulation on polymer surface 4.2 Schematic representation of the synthesis of CdS nanoparticles 61 by reverse micelles in miniemulsion system and CdS/P(S-DVB) nanocomposites via in-situ polymerization 4.3 Reaction of styrene and divinylbenzene by free radical 62 polymerization into P(S-DVB) 4.4 Schematic representation of the encapsulated CdS in P(S-DVB) matrices 62 xiv 4.5 UV – Vis spectra of CdS nanoparticles in miniemulsion system 64 using styrene-divinylbenzene as oil-phase at Po = 32.5 after 5 minutes formation and different Wo values (a) 5.50, (b) 4.0, (c) 2.75, and (d) 1.50 4.6 UV – Vis spectra of CdS nanoparticles in miniemulsion using 65 styrene-divinylbenzene as oil-phase at different Wo values and Po = 32.5 at different time intervals after formation (a) Wo = 1.5 (b) Wo = 2.75 (c) Wo = 4 and (d) Wo = 5.55 4.7 XRD patterns of pure P(S-DVB), CdS nanoparticles and 68 CdS/P(S-DVB) nanocomposites with various amounts CdS 4.8 SEM images of CdS/P(S-DVB) nanocomposites at (a) 2,500 and 70 (b) 10,000 magnifications 4.9 TEM images of CdS/P(S-DVB) nanocomposites 71 4.10 FTIR spectra of (a) pure P(S-DVB) and (b) CdS/P(S-DVB) 72 nanocomposites 4.11 UV-Vis DR spectra of pure P(S-DVB) and CdS/P(S-DVB) nanocomposites prepared by in-situ polymerization 73 in miniemulsion system at different Wo value (a) Wo = 4, (b) Wo = 5.55, and (c) Wo = 7.5 4.12 Dielectric constant for P(S-DVB) and CdS/P(S-DVB) 74 nanocomposites with different amount of CdS as a function of various frequencies 4.13 Dissipation factor of (a) pure P(S-DVB) and (b) 0.18 % 75 CdS/P(S-DVB) nanocomposites 4.14 Schematic representation of the mechanism of interfacial 76 interaction in CdS/P(S-DVB) nanocomposites 4.15 Influence of temperature on dielectric constant of 0.03 % 77 CdS/P(S-DVB) nanocomposites at various frequencies 4.16 Percentage decrease of dielectric constant of 0.03 % CdS/P(S- 77 DVB) nanocomposites at various frequencies after heat treatment at 150o C 4.17 Ionic conductivity of CdS/P(S-DVB) nanocomposites at different amount of CdS 78 xv 4.18 Thermal conductivity of (a) pure P(S-DVB), (b) 0.01 % 79 CdS/P(S-DVB), (c) 0.02 % CdS/P(S-DVB), (d) 0.03 % CdS/P(S-DVB) and (e) 0.18 % CdS/P(S-DVB) nanocomposites at room temperature 4.19 Thermogravimetric analyses curves of pure P(S-DVB), and 0.18 79 % CdS/P(S-DVB) nanocomposites 4.20 Schematic representation of synthesis of P(MAA-EGDMA) 82 particles 4.21 Schematic representation of preparation of CdS/P(MAA- 83 EGDMA) nanocomposites 4.22 Reaction of methacrylc acid and ethylene glycol dimethacrylic 84 acid by free radical polymerization into P(MAA-EGDMA) 4.23 UV-Vis DR spectra of (a) P(MAA-EGDMA), (b) 0.81 % 85 CdS/P(MAA-EGDMA), (c) 0.90 % CdS/P(MAA-EGDMA), (d) 1.08 % CdS/P(MAA-EGDMA), (e) 1.23 % CdS/P(MAAEGDMA), (f) 1.70 % CdS/P(MAA-EGDMA) and (g) 2.07 % CdS/P(MAA-EGDMA) 4.24 FTIR spectra of P(MAA-EGDMA), CdS/P(MAA-EGDMA) 87 nanocomposites and physically mixed CdS/P(MAA-EGDMA) 4.25 SEM images of 2.07 % CdS/P(MAA-EGDMA) nanocomposites 89 at magnification (a) 50,000 X and (b) 100,000 X 4.26 TEM images of (a) agglomerated and (b) fused samples 2.07 % 90 CdS/P(MAA-EGDMA) nanocomposites 4.27 XRD patterns of pure P(MAA-EGDMA) and CdS/P(MAA- 91 EGDMA) nanocomposites at various CdS amount 4.28 Dielectric constant of CdS/P(MAA-EGDMA) nanocomposite 93 with various amounts of CdS measured at different frequencies 4.28 Dielectric constant of (a) P(MAA-EGDMA), (b) CdS nanoparticles, (c) CdS/P(MAA-EGDMA) physical mixing, (d) 11.9 % Na-P(MAA-EGDMA), (e) 1.4 % Cd-P(MAA-EGDMA), (f) 0.81 % CdS/P(MAA-EGDMA), (g) 0.90 % CdS/P(MAAEGDMA), (h) 1.08 % CdS/P(MAA-EGDMA), (i) 1.23 % CdS/P(MAA-EGDMA), (j) 1.70 % CdS/P(MAA-EGDMA) and 94 xvi (k) 2.07 % CdS/P(MAA-EGDMA) at 100 Hz 4.30 FTIR spectra of P(MAA-EGDMA), CdS/P(MAA-EGDMA) 96 physical mixing, CdS nanoparticle and CdS/P(MAA-EGDMA) nanocomposites with various content of CdS 4.31 Schematic representation of interfacial interaction in the 97 CdS/P(MAA-EGDMA) nanocomposites 4.32 SEM images of CdS/P(MAA-EGDMA) at various amount of 99 CdS (a) 0.81 % , (b) 1.08 % and (c) 2.07 % 4.33 Dissipation factor of P(MAA-EGDMA) and CdS/P(MAAEGDMA) nanocomposite at various CdS content 100 and frequencies (a) P(MAA-EGDMA), (b) 0.81 % CdS/P(MAAEGDMA), (c) 0.90 % CdS/P(MAA-EGDMA), (d) 1.08 % CdS/P(MAA-EGDMA), (e) 1.23 % CdS/P(MAA-EGDMA), (f) 1.70 % CdS/P(MAA-EGDMA), and (g) 2.07 % CdS/P(MAAEGDMA) 4.34 Ionic conductivity of (a) P(MAA-EGDMA), (b) CdS 101 nanoparticles, (c) 11.9 % Na-P(MAA-EGDMA), (d) 1.4 % CdP(MAA-EGDMA), (e) 0.81 % CdS/P(MAA-EGDMA), (f) 0.90 % CdS/P(MAA-EGDMA), (g) 1.08 % CdS/P(MAA-EGDMA), (h) 1.23 % CdS/P(MAA-EGDMA), (i) 1.70 % CdS/P(MAAEGDMA) and (j) 2.07 % CdS/P(MAA-EGDMA) 4.35 Influence of temperature on dielectric constant of 1.70 % 103 CdS/P(MAA-EGDMA) nanocomposites at different frequencies 4.36 Percentage decrease of dielectric constant of 1.70 % 103 CdS/P(MAA-EGDMA) nanocomposites after heat treatment and at various frequencies 4.37 Influence of temperature on ionic conductivity of CdS/P(MAA- 104 EGDMA) nanocomposites 4.38 Thermal conductivity of (a) P(MAA-EGDMA), (b) CdS 105 nanoparticle, (c) 0.81 % CdS/P(MAA-EGDMA), (d) 0.90 % CdS/P(MAA-EGDMA) , (e) 1.08 % CdS/P(MAA-EGDMA), and (f) 2.07 % CdS/P(MAA-EGDMA) nanocomposites 4.39 Thermogravimectric analysis curve of (a) P(MAA-EGDMA), 106 xvii (b) 0.90 % CdS/P(MAA-EGDMA) (c) 1.23 % CdS/P(MAAEGDMA) and (d) 2.07 % CdS/P(MAA-EGDMA) nanocomposites 4.40 Reaction of styrene and divinyl benzene by free radical 107 polymerization and sulfonation process to form SO3H/P(S-DVB) 4.41 Schematic representation of the synthesis of CdS/SO3-P(S-DVB) 109 nanocomposites 4.42 UV-Vis DR spectra of (a) Pure P(S-DVB), (b) SO3H-P(S-DVB), 110 (c) 2.56 % CdS/SO3-P(S-DVB), (d) 3.34 % CdS/SO3-P(S-DVB), (e) 7.64 % CdS/SO3-P(S-DVB), (f) 14.65 % CdS/SO3-P(S- DVB) (g) 16.56 % CdS/SO3-P(S-DVB) and (h) CdS nanoparticles 4.43 FTIR spectra of P(S-DVB), SO3H-P(S-DVB) and CdS/SO3-P(S- 112 DVB) nanocomposites 4.44 XRD patterns of (a) CdS nanoparticles and (b) bulk CdS 113 4.45 XRD patterns of SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) 114 nanocomposites at various CdS content 4.46 SEM images of 7.64 % CdS/SO3-P(S-DVB) nanocomposites at 116 two magnifications (a) 2,000 X and (b) 5,000 X 4.47 SEM images of the surface of CdS/SO3-P(S-DVB) at 117 magnification 100,000 X 4.48 TEM images of (a) non agglomerated and (b) agglomerated 118 CdS/SO3-P(S-DVB) nanocomposites 4.49 Dielectric constant of (a) P(S-DVB), (b) CdS nanoparticles, (c) 119 Cd(SO3)2-P(S-DVB), (d) NaSO3-P(S-DVB), (e) SO3H-P(SDVB), (f) 2.56 % CdS/SO3-P(S-DVB), (g) 3.34 % CdS/SO3P(S-DVB), (h) 7.64 % CdS/SO3-P(S-DVB), (i) 14.65 % CdS/SO3-P(S-DVB) and (j) 16.56 % CdS/SO3-P(S-DVB) at 100 Hz 4.50 A graph of the relationship between amount of meq H+ and 121 dielectric constant of CdS/SO3-P(S-DVB) nanocomposites 4.51 Dissipation factor of (a) SO3H-P(S-DVB) and CdS/SO3-P(SDVB) at various CdS content and frequencies. (b) 2.36 % 122 xviii CdS/SO3-P(S-DVB), (c) 3.34 % CdS/SO3-P(S-DVB), (d) 7.64 % CdS/SO3-P(S-DVB), (e) 14.65 % CdS/SO3-P(S-DVB), (f) 16.56 % CdS/SO3-P(S-DVB) 4.52 Influence of temperature on the dielectric constant of 7.64 % 123 CdS/SO3-P(S-DVB) nanocomposites 4.53 Ionic conductivity of (a) Pure P(S-DVB), (b) CdS nanoparticles, 125 (c) Cd(SO3)2-P(S-DVB), (d) NaSO3-P(S-DVB), (e) SO3H-P(SDVB), (f) 2.36 % CdS/SO3-P(S-DVB), (g) 3.34 % CdS/SO3P(S-DVB), (h) 7.64 % CdS/SO3-P(S-DVB), (i) 14.65 % CdS/SO3-P(S-DVB) and (j) 16.56 % CdS/SO3-P(S-DVB) 4.54 A graph of the relationship between amount of meq H+ and ionic 126 conductivity of CdS/SO3-P(S-DVB) nanocomposites 4.55 Ionic conductivity of CdS/SO3-P(S-DVB) nanocomposites at 126 different temperatures 4.56 Thermal conductivity of P(S-DVB), SO3H-P(S-DVB), CdS 127 nanoparticles, and CdS/SO3-P(S-DVB) nanocomposites 4.57 A graph of the relationship between amount of CdS nanoparticle and thermal conductivity of 128 CdS/SO3-P(S-DVB) nanocomposites 4.58 Thermogravimetric analysis curves of (a) CdS nanoparticles, (b) 129 16.56 % CdS/SO3-P(S-DVB), (c) 7.64 % CdS/SO3-P(S-DVB), (d) 2.36 % CdS/SO3-P(S-DVB) and (e) SO3H-P(S-DVB) 4.59 Schematic representation of mechanism replacing proton with sodium and CdS nanoparticles attached on surface of polymer and their properties 130 xix LIST OF SYMBOL / ABRREVIATIONS FTIR - Fourier transform infrared UV - Vis - Ultraviolet – visible DR - Diffuse reflectance TEM - Transmission electron microscopy SEM - Scanning electron microscopy TGA - Thermogravimetric analysis AIBN - 2,2’-Azobisisobutyronitrile XRD - X-ray diffraction P(S-DVB) - Poly(styrene-divinyl benzene) P(MAA-EGDMA) - Poly(methacrylic acid – ethylene glycol dimethacrylic acid) CdS - Cadmium sulfide SO3H-P(S-DVB) - Sulfonated-poly(styrene-divinyl benzene) AOT - Sodium bis(2-ethylhexyl)sulfosuccinate SDS - Sodium dodecyl sulfate CTABr - Cetyltrimethylammonium bromide CVDAC - Cetyl-p-vinyl benzyldimethylammonium chloride PS - Polystyrene Mac - Maleic acid anhydride P(MAA-PMMA) - Poly(methacrylic acid-polymetylmethacrylic acid) MMA-BMA - Methyl methacrylate-butyl methacrylate PVA - Polyvinyl alcohol DMF - Dimethylformamide PEO - Polyethylene oxide mA - Milli ampere KV - Kilovolt AC - Alternating current xx PCB - Printed circuit board PZT - Lead zirconate titanate PANI - Polyaniline PTT - Poly(trimethylene terephtalate) M - Monomer AAS - Atomic absorption spectroscopy K - Dielectric constant εr - Relative permittivity of a material εo - Relative permittivity of vacuum C - Capacitance Q - Coulomb A - Area of the electrical conductor V - Volt PWB - Printed wiring board d - Thickness IC - Integrated circuit PVC - Polyvinyl chloride PI - Polyimide PVB - Polyvinyl butyral PMN - Lead magnesium niobate CB - Circuit board P(TMPTA) - Poly(trimethylolpropane triacrylate) PEN - Polyarylene ether nitriles S-SEBS - Sulfonated styrene-b-(ethylene-ran-butylene)-b-styrene block copolymers LDPE - Low density polyethylene DS - Diphenyl sulfoxide LUMO - Lowest unoccupied molecular orbital HOMO - Highest occupied molecular orbital CMC - Critical micelle concentration HD - Hexadecane SPS - Sodium persulfate MWD - Molecular weight distribution xxi KPS - Potassium persulfate VAc - Vinyl acetate EMA - Ethyl methacrylic acid PDA - Personal digital assistant Hi-Dk RCF - High dielectric constant resin coated foil VBTAC - Vinyl benzyl trimethylammonium chloride SLS - Sodium lauryl sulfate xxii LIST OF APPENDICES APPENDIX. A TITLE Comparison of stabilization time of CdS nanoparticles in PAGE 149 miniemulsion at Wo = 5 with different Po values B UV – Vis spectra of CdS nanoparticles in miniemulsion at 150 Wo = 5, Po = 65 and concentration of cadmium and sulfide ion of 3.0 x 10-4 M at different time after formation C UV - Vis spectra of CdS nanoparticles in miniemulsion 151 system at Po = 65 with Wo values = 5 and 10 at different time after formation D UV - Vis spectra of CdS nanoparticles in miniemulsion at 152 Wo = 5 and Po = 65 at different concentration of cadmium and sulfide ion E UV – Vis spectra of CdS nanoparticles in miniemulsion 153 system before and after addition styrene-divinyl benzene at different times after formation, (a) and (d) without monomer at 5 min and 24 h, respectively, (b) and (c) after mixing with monomer at 5 min and 24 h, respectively F Quantitative standard calibration plot of cadmium element 154 by using Perkin Elmer AA400 AAS G Quantitative standard calibration plot of sodium element 155 by using Perkin Elmer AA400 AAS H List of publications 156 1 CHAPTER 1 INTRODUCTION High dielectric materials have been actively explored and used, because the material has potential for application in microwave communication devices, artificial muscles, and embedded capacitor for micro-electromechanical system. Dielectric materials that can store large electric energy are highly desirable for many electronic and electric systems for energy pulse and power conditioning applications. Ceramic materials usually have large dielectric constant, but they are limited by their relative small breakdown strength. On the other hand, polymers usually enjoy higher breakdown strength but suffer from much lower dielectric constant. Nanotechnology has been the subject of increasing interest in recent years due to the optical, dielectric, electric, magnetic, biological, pharmaceutics and catalytic properties that are present in the metallic and inorganic nanoparticles. These properties are used in the development of different nanodevices, including microelectronic uses. The ever escalating speed, functionality and portability requirement for microelectronic products exert tremendous pressure interest for researchers and manufacturers to meet with the rapid growing demand for the miniaturization and performance. The miniaturization and high performance in electronic devices has driven research and development activities to produce embedded passive. Passives are non-active electrical elements and can be divided into resistive, capacitive and inductive components. In a typical electronic product today, more than 80 % of the surface area of the printed circuit board (PCB) is occupied by passive component (Lu et al., 2006). 2 By eliminating surface mount components and embedding it into the substrate board, embedded passive component offer various advantages over traditional discrete ones, such as higher component density, increased functionality, improved electrical performance, increased design flexibility, improved reliability and reduced unit cost. The architecture of passive component is one area with room for improvement due to the large and growing number of passive component in today’s increasingly functional devices. Discrete passives, especially capacitors, have already become the major barrier of the electronic systems miniaturization. Therefore, the development of embedded passives is desired, if not required. Among passives component, the development of embedded capacitor has been an area of significant activity because the capacitor use in multiple functions, such as decoupling, by-passing, filtering, and timing capacitors (Rao and Wong, 2004; Lu et al., 2007). Figure 1.1 shows a schematic of embedded capacitance solution for electronic devices. High speed computing boards (Servers, router, super computers) Power distribution improvement Ultra-thin substrate (used as embedded capacitor) Copper Foil 8 – 24 micron Polymer dielectric Module board Cell phones, PDA, note book Miniaturization/HDI Hi-Dk RCF (used as embedded capacitor) Copper Foil 16 micron polymer dielectric with Hi-Dk filler Figure 1.1: Schematic representation of embedded capacitance solution for electronic devices 3 Studies polymer/ceramic and syntheses nanocomposites, of polymer nanocomposites, polymer/inorganic including nanocomposites and polymer/metal nanocomposites with high dielectric constant have been actively explored, with the hope to substantially enhance the electric energy density of the resulting nanocomposite. The utilized polymer nanocomposites for electronic application have many advantages such as light weight, shape-flexibility, cost effectiveness, and good process ability of the material. Many studies were developed to produce polymer nanocomposites with high dielectric constant. Polymer nanocomposites are appealing for two reasons. First of all, they possess large interfacial exchange coupling through a dipolar interface layer and leading to enhancement in polarization and polarizability in polymer matrix near the interface. As a result, enhanced dielectric constant can be expected in the polymer matrix near the interfaces. Secondly, the nanoscale particles also make it possible to reduce the thickness of polymer matrix film to nano range, and thus increase its already high breakdown strength even further by avoiding avalanche effect. Nanoparticles are generally categorized as the class of materials that fall between the molecular and bulk solid limits, with an average size between 1 – 100 nm. Semiconductor nanoparticles, referred to as quantum dots, with dimensions in the order of nanometers have been the subject of intense research in the past two decades, due to their unique optical, electronic, physical and chemical properties (Alivisatos, 1996; Wang and Herron, 1991). Inorganic materials on nano-sized metals and semiconductors provide a potential solution to meet present and future technological demand in virtue of the novel properties and unique property combination of both metal and semiconductor nanoparticles. On the other hand, uniform dispersion of nanoparticles in the nanocomposites is required because clumps of particles inside the polymer matrix will not lead to desirable electrical or dielectric properties. However, uniformly dispersed ultrafine particles in polymer matrix may not be easily achieved by incorporating pre-made nano-size particles into a polymer. This is caused by the easy agglomeration of nanoparticles and high viscosity of polymers. The most promising way to decrease an obstacle of these 4 factors on the dielectric properties of nanocomposites is the in-situ or direct formation of nanoparticles in polymer matrix. The focus of this study is to synthesize CdS/polymer nanocomposites and characterization of the physicochemical and electrical properties. The materials were synthesize by in-situ polymerization and ion exchange-precipitation method. The nanocomposites were carried out by several techniques such as UV – Vis spectroscopy, UV - Vis Diffuse Reflectance (UV – Vis DR) spectroscopy, Fourier Transform Infrared (FTIR) Spectroscopy, X-ray Diffraction (XRD), Atomic Absorption Spectroscopy (AAS), Thermogravimetry Analysis (TGA), Transmission Electron Microscopy (TEM), Field Emmision Scanning Electron Microscopy (FESEM), AC Impedance Analyzer, and Thermal Conductivity Analyzer. The used of CdS nanoparticles as nanofiller have received great attention because of their unique electrical and optical properties. The selected approach of synthesis methods and the intrinsic of chemical properties of selected polymers will be produced on nanoscale CdS by encapsulated within polymer and attached on polymer matrices. The combination of CdS nanoparticles and the polymer have developed a CdS/polymer nanocomposite with a high dielectric constant and improved physicochemical properties. 1.1 Background of Study Many researchers have developed methods for synthesis of high dielectric properties polymer nanocomposites. The first approach process for enhancing the dielectric constant of a polymer nanocomposites is to disperse a high dielectric constant insulating ceramic (ferroelectric materials), namely nanoceramic using barium titanate (BaTiO3) (Devaraju et al., 2005; Pant et al., 2006), lead titanate (PbTiO3), lead zirconate titanate (PZT) (Dong et al., 2006), strontium titanate (SrTiO3) (Prijamboedi et al., 2005), into polymers. In order to obtain a high value of 5 dielectric constant, large amount of the filler has to be loaded into polymer matrix, resulting in loss of flexibility and inhomogeneous nanocomposites. The second approach to obtain high dielectric constant polymer nanocomposites was achieved by dispersing conductive fillers into polymers. The common conductive materials used to produce the polymer nanocomposites are carbon nano-fibers, metals such as silver (Ag) nanoparticles (Lu et al., 2006), aluminium (Al) and nanoparticle aluminium (Xu and Wong, 2007), cadmium oxide (CdO) (Pant et al., 2006), zink sulfide (ZnS) (Ghosh et al., 2005b), polymers such as polyaniline (PANI) (Lu et al., 2007), and organic acids such as sulfamic acid (Ameen et al., 2007) and copper pthalocyanine. The polymer nanocomposites have successfully increased the dielectric constant of the polymer. The increased dielectric constant observed in such composites arises from conducting particles isolated by very thin dielectric layers to form micro-capacitors. However, the dielectric loss is very high and difficult to control, because the particles can easily form a conductive path in the composite as the filler concentration nears the percolation threshold. The third approach to increase dielectric properties of polymer was to utilize an inorganic and organic materials such as, titanium oxide (TiO2) (Li et al., 2006; Yang and Kofinas, 2007; Mo et al., 2008; Dey et al., 2004), aluminium oxide (Al2O3) (Li et al., 2007), clay (Zhang et al., 2005), zinc oxide (ZnO2) (Hong et al., 2005) and also other polymers such as poly(trimethylene terephtalate) (PTT) (Kalakkunnath and Kalika, 2006). The results showed that the dielectric constant of the polymer did not increase significantly. The challenge to produce dielectric polymer nanocomposites relies on the ability to manipulate the fraction, characteristic length, and arrangement of dielectric component inside the engineered nanocomposites. No systematic studies to date have been published on the influence of effective volume fraction, characteristic length, and arrangement of dielectric components on the effective dielectric constant of the composites. This is due the intrinsic incompatibility between inorganic particles and organic matrices. Most of the previous studies have utilized the conventional method of blending a high dielectric constant material into polymer, which has no real 6 control on particles size and distribution within the polymer matrix (Yang and Kofinas, 2007). The study, synthesis and characterization of semiconductor nanoparticles have attracted intense research lately owing to the unique chemical and physical properties of the nanoparticles and also the vast potential for practical application of the composite system incorporating the nanoparticles. Semiconductor nanoparticles have physical and chemical properties that may differ significantly from those of the bulk material. Such deviations are attributed to the small particle size and the accompanying surface structure effects. By controlling the particle size and surface structures of the semiconductor materials, electronic, optical, magnetic, mechanical, and chemical properties can be modified to suit a wide range of device application in various fields (Zhao et al., 2001). Semiconductor nanoparticles have interesting applications in electronic, optical, electro-optical devices, catalysis (Yanagida et al., 1990; Graetzel and Graetzel, 1979), and optics (Wang, 1991). More particularly, cadmium sulfide (CdS) nanoparticles have made a great impact in applications including pigments, battery, optoelectronic devices, photocatalyst (Yanagida et al., 1990; Graetzel et al., 1979; Hirai et al., 2002a, 2001, Hirai and Bando, 2005 and Yang, 2005), photosensitive matrices, and resistor for light detecting laser. Dielectric properties of CdS nanoparticles and bulk CdS have also been studied and reported (Zhou, 2003; Tiwari and Tiwari, 2006). Polymers as matrices have several advantages for producing CdS nanoparticles and CdS/polymer nanocomposites. On the other hand, polymers with a type of spatial conformation can be used as template to make nanoparticles in solution with narrow size distribution and uniform confinement throughout their periodic micro-domain. Besides that, polymers are able to achieve surface passivation, prevent particles from agglomeration and maintain the particles degree of dispersion of particle, which are well-known for their ability in controlling the particle size and size distribution effectively (Yang et al., 2003). 7 Based on these above explanations and reasons, studies on the synthesis, characterization, and elucidation of the dielectric and electrical properties of CdS/polymer nanocomposites are undertaken to develop a high dielectric constant material. The CdS/polymer nanocomposites were synthesized by combining several methods i.e. in-situ polymerization, ion exchange and precipitation. The polymer matrices employed are poly(styrene-divinylbenzene) [P(S-DVB)], poly(methacrylic acid-ethyleneglycol dimethacrylic acid) [P(MAA-EGDMA)] and sulfonated poly(styrene-divinylbenzene) [SO3H-P(S-DVB)]. 1.2 Design and Strategy of Research The synthesis strategy for preparation of high dielectric constant CdS/polymer nanocomposites with selected polymer matrices is shown schematically in Figure 1.2. In this study, the CdS/polymer nanocomposites were synthesized using three kinds of polymer i.e. P(S-DVB), P(MAA-EGDMA) and SO3H-P(S-DVB). Polystyrene and polymethacrylic acid are common polymers used in preparation of encapsulated metal or metal oxide and solid phase extraction, respectively. The choice of the polymers is based on their different properties and degree of polarity. Among the polymers employed in the present study polystyrene has no functional group, polymethacrylic acid has a weak acid functional group (COOH), while sulfonated-polystyrene has a strong acid functional group (-SO3H). Based on the different chemical properties of the polymers, hence different procedure of synthesis was applied in order to disperse CdS nanoparticles on the polymer surface or encapsulated within the polymer matrix. 8 Nano/micron polymer matrix CdS nanoparticles Intramolecular interaction Intermolecular interaction CdS/polymer nanocomposites with high dielectric constant Figure 1.2: Schematic representation of the synthesis of CdS/polymer nanocomposites In this study, CdS/poly(styrene-divinylbenzene) [CdS/P(S-DVB)] nanocomposites were synthesized by in-situ polymerization in miniemulsion system. First, the CdS nanoparticles were prepared by reverse micelles in miniemulsion system using CTABr as the surfactant, styrene-divinyl benzene as the oil-phase, 2propanol as the co-stabilizer, and also water. While, the CdS/poly(methacrylic acidethyleneglycol dimethacrylic acid) [CdS/P(MAA-EGDMA)] nanocomposites were synthesized by combination of three methods i.e. in-situ polymerization to produce P(MAA-EGDMA) nanoparticles, ion exchange and precipitation process to produce CdS/P(MAA-EGDMA) poly(styrene-divinyl nanocomposites, respectively. The CdS/sulfonatedbenzene) [CdS/SO3-P(S-DVB)] nanocomposite were synthesized by combining four methods i.e. in-situ polymerization to produce P(SDVB) nanoparticles, sulfonation reaction with fuming sulfuric acid (H2S2O7) to 9 produce SO3H-P(S-DVB), ion exchange and precipitation processes to produce CdS/SO3-P(S-DVB) nanocomposites, respectively. 1.3 Objective of The Study The main goals of this research are to synthesize, characterize and evaluate the dielectric and electrical properties of the CdS/polymer nanocomposites. The objectives of the research are: 1. To synthesize and characterize CdS nanoparticles prepared by reverse micelles in miniemulsion system using monomer as the oil-phase 2. To synthesize and characterize pure P(S-DVB) and P(MAAEGDMA) prepared by in-situ polymerization in miniemulsion system. 3. To synthesize and characterize CdS/P(S-DVB) nanocomposites prepared by in-situ polymerization in miniemulsion system using monomer as the oil-phase 4. To synthesize and characterize CdS/P(MAA-EGDMA) nanocomposites prepared by ion exchange and precipitation processes. 5. To synthesize and characterize CdS/SO3-P(S-DVB) nanocomposites prepared by sulfonation reaction, ion exchange and precipitation processes. 6. To evaluate and to compare the dielectric properties, electrical properties, thermal properties and thermal conductivity of these CdS/polymer nanocomposites. 1.4 Scope of The Study The scope of the study is focused on the following aspects i.e. synthesis of CdS nanoparticles, synthesis of pure P(S-DVB) and P(MAA-EGDMA), preparation 10 of CdS/P(S-DVB) nanocomposites, preparation of CdS/P(MAA-EGDMA) nanocomposites, preparation of CdS/SO3-P(S-DVB) nanocomposites, and evaluation of dielectric properties, electrical properties, thermal properties and thermal conductivity. Table 1.1: Outline of results and discussion Outline of results and discussion CdS/P(S-DVB) nanocomposites • Section in chapter 4.1 Optimization of CdS synthesis and synthesis of CdS/P(S-DVB) • Physicochemical properties of CdS/P(S-DVB) • Dielectric properties of CdS/P(S-DVB) CdS/P(MAA-EGDMA) nanocomposites • Physicochemical properties of CdS/P(MAA-EGDMA) • Dielectric properties of CdS/P(MAA-EGDMA) CdS/SO3-P(S-DVB) nanocomposites • Physicochemical properties of CdS/SO3HP(S-DVB) • Dielectric properties of CdS/SO3-P(S-DVB) Comparison of physicochemical and electrical 4.2 4.3 4.4 properties of CdS/polymer nanocomposites This research involves the synthesis of CdS nanoparticles by reverse micelles in miniemulsion using styrene-divinyl benzene as oil-phase, followed by the synthesis and characterization of CdS/P(S-DVB) nanocomposites via in-situ polymerization in the miniemulsion system. The miniemulsion system is prepared using CTABr as surfactant, monomer as oil-phase, 2-propanol as co-stabilizer and water. The formation, stabilization and particle size of CdS nanoparticles in the miniemulsion system were characterized using UV – Vis spectroscopy. Outcomes of this work are reported in Chapter 4. Chapter 4 also reports the synthesis and characterization of dielectric properties of CdS/P(MAA-EGDMA) nanocomposites 11 and observation on the preparation and characterization of dielectric properties of CdS/SO3-P(S-DVB) nanocomposites. The outline of the results and discussions section are shown in Table. 1.1. Chapter 5 presents the conclusions of the synthesis, characterization and dielectric properties of all the nanocomposites of the prepared CdS/polymer nanocomposites. 12 CHAPTER 2 LITERATURE REVIEW 2.1 Dielectric Properties and Dielectric Constant Dielectric constant (K), also known as relative permittivity (εr), is the ratio of the permittivity of a material (εr) to that of vacuum (εo). The relationship between the capacitance (C) and dielectric constant is given by Equation 2.1: C = Q/V (coulomb/volt) = εr εo A/d (Farad) Equation (2.1) Where εo is the dielectric constant of the free space (8.854 x 10-12 F/m); A, the area of the electrical conductor; d, the thickness of material; and εr, the dielectric constant of material layer. In general, all materials with dielectric constant higher than 4.0 (or the dielectric constant of SiO2) are called high dielectric, while those with dielectric constant lower than 4.0 are called low dielectric and those with dielectric below 2.0 are called ultra low dielectric. Any material that contains polar components is represented as dipoles (e.g. polar chemical bond). The dipoles align with an external electric field, adding the electric field of every dipole of the external field. As a result, a capacitor with a dielectric medium other than vacuum will hold more electric charge at certain applied voltage. In other words, its capacitance and dielectric constant will be higher. 13 Plate area A + + + + + + + + + + + + + + _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ d : thickness Figure 2.1: Schematic representation of capacitor There are several molecular mechanism associated with these polarizations: (a) Electronic polarization (displacement of electron). The polarization is the result of dipole moments induced by electric field. (b) Atomic polarization or distortion polarization (displacement of ions or atoms). This polarization is the result from an unsymmetrical sharing of electrons when two different atoms combine to form a molecule. (c) Dipole or orientation polarization (displacement or reorientation) of polar molecules. Some molecules have a permanent dipole that can react with an external electric field and for dipole or orientation polarization. (d) Space – charge polarization (long range charge migration). The polarization is different from the previous three types of polarization because it is not due to the charges that are locally bound. Instead, it is produced by charge carriers, such as free ions, that can migrate for some distance throughout the dielectric. Space charge transfer polarization is extrinsic to any crystal lattice and arises as a result of charges on the surface, grain and phase boundaries. The polarization produces the highest dielectric constant value. The schematic representations of these various polarizations are shown in Figure 2.2 14 Insulator (a) Electronic polarization occurs in all insulators Atomic (b) Molecular polarization occurs in all insulating molecules, such as oils, polymer, water Dipole (c) Ionic polarization occurs in all ionic (d) Space charge polarization solid, such as NaCl, MgO Figure 2.2: Schematic representation of types of polarizations The schematic representations of these frequencies dependence contributing to the polarizability are shown at Figure 2.3. 15 Molecular Polarization Ionic Polarization Electronic Polarization Figure 2.3: Frequency dependence of several contributions to the polarizability 2.2 Embedded Capacitor Electronic technologies that allow for a reduction in size, shape, weight, and cost while improving functionality and performance are highly desired for military and commercial applications, which includes telecommunications, network systems, automotives, and computer electronic devices. Electronic systems are composed of active and passive components. Passive components include resistor, capacitor and inductor. The architecture of passive component is one with room for improvement due to the large and growing number of passive components in today’s increasingly functional devices. Today, more than 80 % of the area of printed wiring board (PWB) is occupied by passive components. Discrete passives, especially capacitors, have already become the major barrier of the electronic system miniaturization. Therefore, the development of embedded capacitor has been an area of significant activities because capacitors are used as multiple functions, such as decoupling, bypassing, filtering and timing capacitor (Rao and Wong, 2004). 16 Therefore, integration of passive component inside the substrate, so namely embedded passives, is the key for the next generation electronics. Embedded passives offer many advantages over discrete component. Discrete surface mounted passives require large board area and can introduce additional parasitic into the system, which may limit system performance. With integral embedded passives, the size of the board can be reduced, especially if integrated circuit (IC) chips can be place above the embedded passives component. Also, embedding these devices leave free board space for mounting component that cannot be integrated. Integral passives are defined as functional elements either embedded in or incorporated on the surface of an interconnecting substrate. With increased production, emphasis towards efficient electronic packaging, integral embedded passives technology may satisfy such demand. The several main advantages of embedded passives component are (a) no separate interconnect to the substrate, (b) improved electrical performance, (c) lower cost and (d) ease of processing. For producing the embedded passives component a material with high dielectric constant are required because of the development of advanced decoupling capacitor in order to reduce the power distribution (Rao and Wong, 2004). 2.3 Design of High Dielectric Property Materials Table 2.1 shows the dielectric constant of several substances including polymer, ceramic, metal, metal oxide, etc. Polymers generally have low dielectric constant and it is difficult for them to meet the capacitor requirements. On the other hand, ceramic or metal have high dielectric constant but requires high processing temperatures. It is already known that polymer has a low dielectric properties compared to those of inorganic materials. Due to the processibility of polymer, it is of interest to increase its dielectric properties by combining with inorganic materials by an appropriate approach to obtain polymer nanocomposites. Nanomaterials combine one or more separate components in order to obtain the best properties of each component (composite). In nanocomposite, nanoparticles 17 act as fillers in a matrix, usually polymer matrix. Polymer nanocomposites combine these two concepts, i.e., composites and nanometer-size materials. Table 2.1: Dielectric constant of several materials (Sears et al., 1982) Material Inorganic material Polymer Dielectric constant Potassium tantalate niobate 6000 Barium titanate 4000 Potassium niobate 700 Rochelle salt 170 TiO2 rutile 50.0 Silicon 11.8 GaAs 10.9 Magnesium titanate 10.0 Magnesium oxide 9.0 Epoxy 4.0 Fused silica 4.0 Nylon 6,6 4.0 PVC 3.5 Polystyrene 2.5 Polyethylene 2.3 Polymer nanocomposites (PNCs) are materials composed of a polymeric host in which particles (buckyballs, nanotubes, semiconductor of metallic nanocrystals, clays) of nanoscale dimensions are incorporated. Because the particles and the polymers posses different functionalities, the range of applications is diverse. Applications include membranes, coatings, solar cells, electronic devices, sensing, medicine and structural applications, such as automobile parts. PNCs which contain layered silicates (clays) have been successfully used for different thermal and mechanical applications. Conjugated polymers mixed with various semiconducting particles have received considerable attention because of applications involving solar cells. PNCs are especially interesting because they exhibit properties that are superior to conventional composites with relatively low concentrations of the 18 inorganic component. The challenges are associated with controlling the composition and spatial distribution of the inorganic component within the polymer. Prediction of the properties is often challenging because the result of a range of complex and collective intermolecular interactions. There has been enormous interest in the commercialization of polymer nanocomposites for a variety of applications, and several of these applications will be successful in the near future. Mineral fillers, metals, and fibers have been added to polymer material for decades to form composites. Compared to neat resins, these composites have a number of improved properties including tensile strength, heat distortion, temperature, and modulus. Thus, for structural applications, composites have become very popular and are sold in billion pounds quantities. These filled thermoplastics are sold in even larger volumes than neat thermoplastics. Furthermore, the volume of fillers sold is roughly equal to the volume of thermoplastic resin sold. Clearly, the idea of adding fillers to thermoplastics and thermosets to improve properties, and in some cases decrease costs, has been very successful for many years. More recently, with advances in synthetic techniques and the ability to readily characterize materials on an atomic scale has lead to interest in nanometer-size materials. Since nanometer-size grains, fibers and plates have dramatically increased surface area compared to their conventional-size materials, the chemistry of these nanosized materials is altered compared to conventional materials. Polymer nanocomposites were intensively studied and synthesized to produce high dielectric constant materials. The materials combine the high dielectric constant of ceramics, metals, metal oxides, semiconductors, inorganic and organic substances and the ease to process ability and flexibility of polymers. One of the main advantages of polymer nanocomposites is that their dielectric properties can be changed over a wide range by choosing the shape, size and fractions of the filler into polymer matrices. Dispersion of inorganic phases into polymer matrix is an issue of both scientific and technological interest. The unique properties and processing flexibility of these polymer nanocomposites have found a 19 number of technological applications. Many researchers have studied and developed methods for the synthesis of high dielectric constant polymer nanocomposites. (Dang et al., 2007; Devaraju et al. 2005). 2.3.1 Overview of Dielectric Properties of Polymer Nanocomposites Many researchers have studied and synthesized the polymer/ceramic nanocomposites using common ceramic materials such as barium titanate (BaTiO3), lead zirconate titanate (PZT) and lead titanate (PbTiO3) (Dang et al., 2007). Devaraju et al. (2005) have studied and synthesized BaTiO3/polyimide (BaTiO3/PI) nanocomposites film by in-situ imidization in n-methyl pyrrolidinone as the solvent. The result of the study showed high dielectric constant nanocomposites at various BaTiO3 content (in the range 40 – 90 wt.%) and frequency 1 kHz to 1 MHz. The single layer capacitors were formed and a dielectric constant as high as 125 have been attained at a frequency of 1 kHz (about 117 at 1 MHz), with a 90 wt.% loaded film. The nanocomposites showed high stability dielectric constant in the frequency range of 1 kHz to 1 MHz. Pant et al. (2006) have investigated BaTiO3/polyaniline (BaTiO3/PANI) and BaTiO3/maleic acid resin (BaTiO3/Mac). Both nanocomposites were synthesized by mechanical grinding process. Dielectric properties of both nanocomposites were investigated for microwave studies in the frequency range of 8.2 – 12.4 GHz. They observed that pure BaTiO3 showed quite high dielectric constant value in the range of 100 – 125. However, it is found that on mixing with 10 wt.% of PANI the dielectric constant values of BaTiO3/PANI drastically change and reduce to 22 over the whole frequency range. Different result was obtained for BaTiO3/Maleic acid resin nanocomposites at the same percentage and frequency in which the dielectric constant of the nanocomposites was at about 70 – 80. Dong et al. (2006) have synthesized polyvinyl-butyral/lead zirconate titanates (PVB/PZT) by a simple mixing method. The highest dielectric constant of 155 can be achieved at a PVB volume fraction of 0.15 at 100 Hz. Rao et al. (2002) incorporated lead magnesium niobate – lead titanate/BaTiO3 (PMN-PT/ BaTiO3) into 20 epoxy matrix. The PMN-PT/BaTiO3/epoxy nanocomposites with 70 wt.% ceramic loading and showed a dielectric constant value of 110 at 1 kHz. Several researchers have also studied and synthesized the polymer nanocomposites using many inorganic and organic materials as filler such as, nano aluminium, conducting polymer (polyaniline) and sulfamic acid. Xu and Wong (2007) have reported dielectric properties of aluminum/epoxy and nanoaluminum/epoxy composite with coupling and non-coupling treatment using γglycidoxypropyl-trimethoxysilane as sylane coupling agent. The polymer matrix used was diglycidyl ether of bisphenol-A with methyl-hexahydrophtalic anhydride as hardener agent. The dielectric constant before coupling treatment of 70 wt.% microaluminum (3 micrometer)/epoxy and 40 wt.% nanoaluminum (100 nm)/epoxy which were measured at 10 kHz were 82.1 and 52.3 respectively. After the addition of coupling agent, the dielectric constants of both nanocomposites increased to 96 and 61, respectively. The improvement of dielectric constant came from the better dispersion of coupling agent treated aluminum particles in the polymer matrix. Ghosh, et al., (2005b) have studied and synthesized zinc sulfide (ZnS) nanobelt in pores of polyvinylalcohol (PVA) by chemical bath deposition. The dielectric constant of the ZnS/PVA obtained lies in the range of 30 – 130 at 100 – 1000 kHz, and higher than the dielectric constant of pure PVA, which was about 28. Lu et al. (2007) have studied and synthesized conducting polyaniline/epoxy resin (PANI/epoxy resin). The nanocomposites were prepared via in-situ polymerization of mixed aniline salt and epoxy (using EPON 828) matrix. Dielectric properties of PANI/epoxy resin nanocomposites with different PANI contents at 10 kHz have been examined. The dielectric constant at 8 wt.%, 15 wt.%, 20 wt.% and 25 wt.% PANI contents are 10, 192, 916 and 2980, respectively. The results also suggested that the in-situ polymerization was useful to achieve good dispersion and high compatibility of PANI with the epoxy matrix in which it was effective to eliminate agglomeration surrounded by insulating matrix and the formation of fine network led to enhanced dielectric properties of the nanocomposites. 21 Dielectric properties of polyarylene ether nitriles/titanium oxide (PEN/TiO2) hybrid films prepared by sol-gel method have been studied by Li et al. (2006). The results showed that the dielectric constant was slightly increased with increasing content of TiO2 in the polymer matrix. Dielectric constant of 4 wt.% TiO2/PEN at temperature of 200 oC and 250 oC and 1 kHz were 3.7 and 4.0, respectively. Yang and Kofinas (2007) have reported the dielectric properties of TiO2/sulfonated styrene-b-(ethylene-ran-butylene)-b-styrene block copolymers (TiO2/S-SEBS) by simple mixing and gradualy hydrolyzed method. The dielectric constant of uncrosslinked pure S-SEBS at 1, 100, 1,000 and 10,000 Hz were approximately 165,000, 18,000, 10,000 and 6,000 respectively. Dielectric properties of the nanocomposites were decreased with dispersion of TiO2 into polymer matrix and also the dielectric constant of 0.26 to 6.4 wt.% TiO2/S-SEBS nanocomposites lies in the range of 2 – 3.2 at 100 Hz. Li et al. (2007) have reported dielectric properties of nano-aluminum oxide/polyimide (nano-Al2O3/PI) nanocomposites prepared by in-situ polymerization. The result obtained showed that the dielectric constant of pure polyimide and 20 wt.% nano-Al2O3/PI nanocomposites at 100 Hz were 2.4 and 3.2, respectively. This finding was clearly affected by Al2O3 filler because pure alumina has higher dielectric constant compared to PI. The number of polar group increased due to the incorporation of filler into the polymer matrix and it grew bigger with the increase of the filler concentration which caused an extra polarization under an electric field. Zhang et al. (2005) have studied the dielectric properties clay/polyimide (clay/PI) prepared by thermal imidization from the polyimide precursors PMDAODA and organo-modified clay in N-N-dimethylacetamide. The results displayed the dielectric constant decreased with the increase of the clay content and the film nanocomposites showed low dielectric constant when the clay content was over 1 wt.%. The dielectric constant of pure PI and 10 wt.% clay/PI nanocomposites at 1 kHz and temperature 50 oC were 3.3 and 3.0, respectively. 22 Hong et al. (2005) have investigated the dielectric properties of micron-zinc oxide/low density polyethylene (micron-ZnO/LDPE) and nano-zinc oxide (49 nm)/low density polyethylene (nano-ZnO/LDPE) prepared by melt mixing with either homogeneous or controlled inhomogeneous dispersion of ZnO. The dielectric constant of homogeneous 14 wt.% micron-ZnO/LDPE and inhomogeneous nanoZnO/LDPE at 10 kHz were 5 and 6 respectively. The result concluded that the small particles size of the ZnO nanoparticles did not affect the dielectric constant of the nanocomposites. Therefore, the interfaces between ZnO and LDPE did not appear to contribute to dielectric properties of the nanocomposites. Scorcik et al. (2005) have reported dielectric properties of polymethylmethacrylate (PMMA) solution in acetone doped with 20 wt.% of diphenyl sulfoxide (DS) prepared by three consecutive steps, namely sputtering of bottom, spinning film and sputtering of upper. The result showed that the presence of DS dopant will slightly increase the dielectric properties of the nanocomposites. The dielectric constant of 20 wt.% DS/PMMA and PMMA which were measured at 1 kHz and temperature of 30 oC were 7.0 and 3.0, respectively. Podgrabinski et al. (2006) also studied dielectric properties of doped polystyrene (PS) and polymethylmethacrylate (PMMA) films with diphenyl sulfoxide (DS). These composites were prepared from solutions using spin-coating method. The results showed that the DS doping leads to an increase of dielectric constant of the PS and PMMA films. The dielectric constant of 20 wt.% DS/PMMA and DS/PS at 40 oC and 1 kHz were 5.5 and 2.5, respectively. Tanwar et al. (2006) studied dielectric properties of doped polymethylmethacrylate (PMMA) with iodine (I2), FeCl3, and benzoic acid prepared by solution cast method. Dielectric properties of those nanocomposites were increased by a little with increasing content of doped materials. The dielectric constant of pure PMMA, 8 wt.% I2/PMMA, 20 wt.% benzoic acid/PMMA, and 10 wt.% FeCl3/PMMA at 8.92 GHz were 2.47, 2.88, 3.00 and 2.50, respectively. The different results were attributed to the different properties of the dopant materials i.e. iodine an acceptor dopant, benzoic acid has a phenyl group which can function as an 23 electron withdrawing group and FeCl3 is also an acceptor dopant and selected for the dielectric studies. Table 2.2: Dielectric constant and method of synthesis of several polymer nanocomposites Polymer/nanocomposites Dielectric constant Method of synthesis (Reference) 90 wt.% BaTiO3/PI 125 (1 kHz) in situ imidization (Devaraju et al., 2005) 10 wt.% PANI/BaTiO3 22 (8.2 – 12.4 GHz) mechanically grinding 70 – 80 (8.2 – 12.4 GHz) mechanically grinding 155 (100 Hz) simply mixing method 110 (1 kHz) simply mixing method (Pant et al., 2006) 10 wt.% maleic resin/BaTiO3 (Pant et al., 2006) 15 wt.% PVB/PZT (Dong et al., 2006) 70 wt.% PMN-PT/ BaTiO3/epoxy (Rao et al., 2001) 70 wt.% microaluminum 82.1 (10 kHz) simply mixing (3 µm)/epoxy (Xu and Wong, 2006) ZnS/PVA (Ghost et al. 2005b) 15 wt.% PANI/ 30 – 130 chemical (100 – 1000 kHz) deposition 192 (10 kHz) in-situ polymerization epoxy resin (Lu et al., 2007) 3 wt.% TiO2/PEN ( Li et al.,2006) 4.0 (250 oC, 1 kHz) bath sol-gel 24 Table 2.2 (continue): Dielectric constant and method of synthesis of several polymer nanocomposites Polymer/nanocomposites Dielectric constant Method of synthesis (Reference) 0.26 to 6.4 wt.% 2 – 3.2 (100 Hz) simply TiO2/S-SEBS mixing and gradually hydrolyzed (Kofinas, 2007) 20 wt.% nano-Al2O3/PI 3.2 (100 Hz) in-situ polymerization (Li et al., 2007) 3.0 (1 kHz,50 oC) 10 wt.% clay/PI thermal imidization (Zhang et al., 2005) 14 wt.% micron- 5 (10 kHz) melt mixing ZnO/LDPE (Hong et al., 2005) 7.0 (1 kHz, 30 oC) 20 wt.% DS/PMMA sputtering (Scorcik et al., 2005) spinning film 8 wt.% I2/PMMA 2.88 (8.92 GHz) solution cast (Tanwar et al., 2005) 20 wt.% and solution cast benzoic 3.00 (8.92 GHz) solution cast acid/PMMA (Tanwar et al., 2005) 10 wt.% Fe Cl3/PMMA 2.50 (8.92 GHz) (Tanwar et al., 2005) 5.5 (40 oC, 1 kHz) 20 wt.% DS/PMMA spin-coating method o 20 wt.% DS/PS 2.5 (40 C, 1 kHz) spin-coating method (Podgrabinski et al., 2006) In previous studies as mentioned above, the polymer/metal or conductive nanocomposites have demonstrated that the dielectric constant of the polymer significantly increased. One disadvantage of polymer/ceramic and polymer/conducting nanocomposites is that in order to obtain a high dielectric constant, large amount of the filler has to be loaded into polymer matrix, resulting in loss of flexibility and inhomogeneous nanocomposites. The increased dielectric 25 constant observed in such nanocomposites arises from conducting particles isolated by very thin dielectric layers to form micro-capacitors. However, the dielectric loss is very high and difficult to control, because the particles can easily form a conductive path in the nanocomposites as the filler concentration nears the percolation threshold. On the other hand, several polymer nanocomposites have demonstrated that the dielectric constant of the polymers were increased by a little or was not affected at all. The summary of dielectric constants and methods of synthesis of polymer nanocomposites are shown in Table 2.2. 2.3.2 CdS/polymer Nanocomposites The dielectric properties of CdS nanoparticles, bulk CdS, and CdS/polymer nanocomposites have also been investigated and reported. Zhou (2003) studied dielectric properties of conventional CdS powder, cubic and hexagonal CdS nanoparticles at various frequencies. The CdS nanoparticles were prepared by sonochemistry, while the conventional CdS powder was prepared by ion exchange. The results showed that the CdS nanoparticles have higher dielectric constant compared to conventional CdS powder measured at low frequency. The dielectric constant of cubic CdS nanoparticles and hexagonal conventional CdS powder at 100 Hz are 156.34 and 10.7, respectively. The high dielectric constant of CdS nanoparticles was suggested to appear from interfaces with a large volume fraction in the nanoparticles. Also the grain size of the sample has a great influence on the dielectric behaviour of CdS nanoparticles, which was explained via the theory of dielectric polarization, including space-charge polarization. Ghosh et al. (2005a) studied the dielectric properties of CdS nanoparticles growth at porous poly(vinyl alcohol) (PVA) matrix by chemical bath process. The dielectric constant of CdS/PVA nanocomposites was examined at various frequencies and dispersibility. The dielectric constants on dispersibility at 0.8, 0.6, and 0.3 lies in the range 120 – 250. El-Tantawi et al. (2004) also studied dielectric properties of CdS/PVA nanoconducting composite prepared by organosol technique. 26 It was shown that the dielectric constant of 20 wt.% and 40 wt.% CdS/PVA at 1 kHz were 10 and 20, respectively. Mondal et al. (2007) have investigated optical and dielectric properties of CdS nanocomposites embedded in PVA using simple chemical deposition bath. In their report, it was suggested that the dielectric properties of CdS and CdS/PVA were dependent on the particle size of CdS, interface (inter-intraparticle) interaction, and disperse stability of CdS nanoparticle at the polymer matrix. The summary of dielectric constant and methods of synthesis of CdS nanoparticles and CdS/polymer nanocomposites are tabulated in Table 2.3 Table 2.3: Dielectric constant and methods of synthesis of CdS nanoparticle and CdS/polymer nanocomposites Dielectric Method of constant synthesis CdS nanoparticles 156.34 (100 Hz) sonochemistry Zhou, 2003 Conventional CdS 10.7 (100 Hz) ion exchange Zhou, 2003 120 – 250 chemical bath Ghost et al., 2005b Material CdS/PVA Reference process 40 wt.% CdS/PVA 2.4 20 (1 kHz) organosol El-Tantawi et al. technique (2004) Cadmium Sulfide as Semiconductor Nanoparticles Nanoparticles are generally categorized as the class of materials that fall between the molecular and bulk solid limits, with an average size between 1 – 50 nm. Semiconductor nanoparticles, referred to as quantum dots, with dimensions on the order of nanometers have been the subject of intense research in the past two decades, due to their unique optical, electronic, physical and chemical properties (Alivisatos, 1996; Wang and Herron, 1991). 27 Compared to corresponding bulk semiconductor material, semiconductor nanoparticles have two distinguishing characteristics, both related to the size of individual nanoparticles. First, nanoparticles have a large ratio of surface atoms to those located in the crystal lattice, which make the physical and chemical properties of semiconductor nanoparticles particularly sensitive to the surface structure. Second, nanoparticles have size dependent optical properties, namely “quantum size effect”. As dimension of the particle decreased the bandgap energy of the nanoparticles increases. “Quantum size effect”, the bandgap energy of semiconductor nanoparticles increases as their size decreases. The most notable size effect is the significant blue shift of electronic absorption and emission spectrum with decrease in size. Brus has developed a popular effective mass model that relates particle size to the bandgap energy of semiconductor nanoparticles (Chestnoy et al., 1986; Brus, 1986; Steigerwald and Brus, 1990). As illustrated in Figure 2.4, the quantum size effect causes the continuous band of the crystal to split into discrete, quantized levels and the bandgap of semiconductor nanoparticles to increase. As the nanoparticles decreases in size, the bandgap increase, approaching the energy difference between LUMO and HOMO for the individual molecule. Due to the large percentage of surface atoms for nanoparticles, even a small amount of surface defects can introduce a high density of “surface trap”, shown in Figure 2.4. Therefore, the emission of semiconductor nanoparticles is often used as a sensitive probe to reveal their surface structure. The bandgap of semiconductor nanoparticles is predicted to be inversely proportional to the particle size (radius, R2), shown in Equation 2.2 (Brus, 1986; Brus, 1984). Eg (quantum dot) = Eg (bulk) + (h2) ( 1 + 1) – 1.8e2 8R2 mc mh 4πεoεR Equation (2.2) Where Eg is the bandgap energy of the corresponding bulk crystal, mc is the effective mass of the electron in the crystal, mh is the effective mass of the hole in the 28 crystal, and is the electric constant of the crystal. The middle term on the right side on the equation is a particle-in-a-box-like term for the exciton, while the third term represents the electron-hole pair coloumb attraction, mediated by the crystal. Equation 2.2 is based on two assumptions. One is that the nanoparticles are in spherical shapes; the other is that the effective mass of charge carriers and the dielectric constant of the crystal are constant as a function of particle size. BULK NANOCRYSTAL MOLECULE Conduction band LUMO Energy bandgap HOMO Valence band Figure 2.4: Schematic changes in the density of states on going from a bulk crystal to a nanocrystal to molecule/atom For CdS nanoparticles, Brus equation (Equation 2.3) is used to estimate the particle size where the bandgap energy is approximately equal to the band edge of optical absorption spectra. The Brus model works well for estimating the size of larger nanoparticles, but often shows significant deviation for very small particles. Eg (R) = 2.43 + 2.446 – 0.3031 R2 R Equation (2.3) Nanoparticle semiconductors are currently under intense investigation because of their enhanced photoreactivity and photocatalytic properties. Nanoparticles of CdS 29 are by far the most studied system among all the semiconductor nanoparticles as has been used as many applications e.g. photocatalyst (Bard, 1980; Yanagida, et al., 1990; Graetzel and Graetzel, 1979; Hirai, et al., 2002a, 2002b, 2001, 2006, Hirai and Bando, 2005; Hirai and Ota, 2006; Yang, et al., 2005), paint industry, battery, optoelectronic devices, photosensitive matrices, resistor for light detecting, laser, electrochemical cell (Haram, et al., 2001; Hagfeldt and Graetzel, 1995), fluorescent or luminescent labelling (Bruchez et al., 1998; Harruf et al., 2003), pharmaceutical and medical, especially for drug delivery system (Lai et al., 2003), biosensing (Chen et al., 2005), optics (Wang, 1991; Hines, 1998), and unique electronics (Henglein, 1989; Steigerwald and Brus, 1990; Wang, 1991). 2.4.1 Methods of Synthesis of Cadmium Sulfide Nanoparticles A wide range of nanomaterials including metals, semiconductors, inorganic oxides, and polymers have been produced using various methods and nanoreactor systems. Many methods and nanoreactors have been developed and studied including wide variety of techniques, such as micelles or reverse micelles (Bunker et al., 2004; Dutta and Fendler, 2002; Hirai et al., 2000a, 2000b, 2001, 2002a, 2005; Harruf et al., 2003; Zhao et al., 2001) or salt induced micelles (Zhao et al., 2001), aggregation (Libert et al., 2003), liquid crystal (Dellinger and Braun, 2001), miniemulsion or microemulsion (Agostiano et al., 2000; Khiew et al., 2003, 2004), microgels, amphiphylic polymers or block copolymer (Dellinger and Braun, 2001), liposomes (Graff et al., 2001), dendrimes (Crooks et al., 2001), thermally evaporated aerosol (Shankar et al., 2001), UV or microwave irradiation (Yang et al., 2005), precipitation (Banarjee et al., 2000) have been used in the preparation of semiconductor CdS nanoparticles. Several researchers have used a variety of surfactants to prepare the miniemulsion systems for preparation of CdS nanoparticles, such as sodium bis (2ethylhexyl) sulfosuccinate (AOT) (Hirai et al., 2001, 2002a), sodium dodecyl sulfate (SDS), cetyl trimethyl ammonium bromide (CTABr) (Agostiano et al., 2000), tritonX series, nonionic surfactant (Khiew et al., 2004), sucrose ester (Khiew et al., 2003) 30 and polymerizable surfactant such as cetyl-p-vinylbenzyldimethylammonium chloride (CVDAC) (Hirai et al., 2000c) and didecyl dimethylammonium methacrylate. Hirai and Bando (2005) synthesized CdS/aluminosilicate particle (AS-SH) composites. The CdS nanoparticles were prepared in reverse micelles system using sodium bis (2-ethylhexyl) sulfosuccinate (AOT), water and isooctane at Wo = [H2O]/[AOT] = 6. Ni, Ge, and Zhang (2006) have been reported and synthesized of CdS/polyacrylonitrile nanocomposites in an ethanol solution by γ-irradiation technique at room temperature and atmospheric pressure. The XRD result showed that the crystallinity of CdS was affected by several factors such as the amount of the organic compound in the composites and the dose of irradiation. Xu and Akins (2004) synthesized of CdS nanoparticles by reverse micellar in microemulsion system and self-assembly into superlattice. The microemulsion system was prepared using AOT-water-isooctane with Wo = [H2O]/[AOT] = 9, Cd(OAT)2 as cadmium ion source and thioacetamide in DMF as H2S source. The UV-Vis spectra of CdS nanoparticles displayed a blue-shifted band-edge when compare to bulk CdS (< 5 micron) i.e. 489 nm vs. 540 nm, in line with expectation associated with formation of quantum dots of semiconductor materials. But, the baseline of the spectra is high which possibly signal the presence of CdS aggregates. While, the TEM micrograph showed that the CdS nanoparticles are essentially monodispersed with particles size about 13 nm and the two dimensional superlattice hexagonally oriented with interparticles spacing of ca. 3 nm. The result also concluded that the CdS nanoparticles are capped and surrounded by AOT surfactant. Zhang et al. (2001) have described a controlled synthesis of CdS nanoparticles using functionalized MCM-41 (Phe-MCM-41) by ion exchange reaction. The absorption onset wavelength of CdS nanoparticles was close to 450 nm, which is significantly blue-shifted relative to that compare bulk CdS. The TEM and XRD analyses showed that the particles size of CdS nanoparticles were 2.5 and 3.6 nm, respectively. 31 Hirai et al. (2000c) prepared CdS nanoparticles-polymer composites by direct reverse micelle polymerization using cetyl-p-vinylbenzyldimethylammonium chloride (CVDAC) as polymerize surfactant by photopolymerization method. The result observed that the CdS nanoparticles were incorporated successfully into polymer (PCVDAC) and retained their size and quantum size effect. The onset wavelength UV – Vis spectra of CdS in miniemulsion at Wo = 1.5 and CdS/PCVDAC in ethanol were observed significantly blue-shifted compared to bulk CdS with particles size of 4.8 nm and 4.9 nm. The result concluded that the photopolymerization of CdS with PCVDAC was effective in incorporating the CdS nanoparticles into PCVDAC and also in maintaining their quantum size effect. Table 2.4: Several CdS/polymer nanocomposites, synthesis method, applications and property of interest Material Method of synthesis (Reference) Applications and property interest CdS/PVA chemical bath process, dielectric and (Ghosh et al.,2005a) organosol technique electrical properties CdS/PVA organosol technique conducting properties (El-Tantawy et al., 2004) CdS/PEO simply mixing, optical properties (Yang et al., 2003) precipitation CdS/PAA precipitation, (Chen et al., 2006) polymerization determination of protein CdS/SBA-15 acid base reaction, optical properties (Xu et al., 2002) ion exchange, calcination CdS/Aluminosilicate reverse micelles Photocatalytic γ-irradiation optical properties simply mixing optical properties in-situ quantitative (Hirai and Bando, 2005) CdS/PAN (Ni et al., 2006) CdS/polyphosphazene (Olshasky and Allcock, 2007) 32 Nanometer-sized semiconductors of lead sulphide (PbS) and CdS prepared by microemulsion using sucrose monoester (S-1670) as surfactant, 1-butanol as costabilizer, cyclohexane as oil-phase and water were studied by Khiew et al. (2003). The TEM micrograph exhibited that the particle size of CdS and PbS were in the range 2 – 7 nm and 3 – 8 nm, respectively. While, the particle size of the CdS and PbS nanoparticles estimated from the absorpstion onset wavelength were 5.1 nm and 4.5 nm, respectively. Yang et al. (2005) studied luminescence and photocatalytic properties of CdS nanoparticles. The CdS nanoparticle was prepared by microwave irradiation from cadmium chloride (CdCl2) solution and sodium sulfide (Na2S) solution. The particle size of CdS nanoparticle obtained from TEM was about 10 nm. The red photoluminescence of CdS nanoparticles also displayed a strong peak at 602 nm, which may be attributed to the recombination of an electron trapped in a sulfur vacancy with a hole in the valence band of CdS. The summary of several CdS nanocomposites materials, method of synthesis, application and property interest were tabulated in Table 2.4. 2.5 Miniemulsion Polymerization A conventional emulsion polymerization is one of the most common techniques employed for encapsulation of inorganic materials, such as titanium dioxide, magnetite, silicon dioxide, etc. In this method, the principal locus for particle nucleation is either in the aqueous phase or in the monomer swollen micelles. The dominant site of particle nucleation depends on the degree of water solubility of monomer(s) and the amount of surfactant used. The complexity of the particle nucleation mechanism and the difficulties in controlling the dispersion stability of inorganic particles in the aqueous phase during emulsification and encapsulation polymerization are the biggest obstacles to success of the encapsulation method and need to be investigated. 33 In contrast, the characteristic features of the miniemulsion polymerization technique may present several advantages as an encapsulation method. Potential advantages include the ability to control the droplet size, having particles directly dispersed in the oil phase, the ability to nucleate all the droplets containing the particles, and a faster rate of polymerization (Erdem et al., 2000a, 2000b). 2.5.1 Introduction Polymeric dispersions are used in a wide variety of application such as synthetic rubber, paints, adhesives, binder for non-woven fabrics, additive in paper and textiles, leather treatment, impact modifiers for plastic matrices, additive for construction materials and flocculants. They are also used in biomedical and pharmaceutical applications such diagnostic tests and drug delivery systems. The rapid increase of this industry is due to environmental concerns and governmental regulations to substitute solvent-based systems by water borne products, as well as to the fact that polymeric dispersions have unique properties that meet a wide range of market needs (Asua, 2002). Commonly, these products are produced by means of conventional emulsion polymerization. In this process, monomer is dispersed in an aqueous solution of surfactant with a concentration exceeding the critical micelle concentration (CMC) and polymerization is started by means of an initiator system. In principle, polymer particles can be formed by entry of radicals into the micelles (heterogeneous nucleation), precipitation of growing oligomers in the aqueous phase (homogeneous nucleation) and radical entry monomer droplets. However, monomer droplets are relatively large (1 – 10 micron) compared to the size of monomer-swollen micelles (10 – 20 nm), and hence the surface area of the micelles is in orders of magnitude greater than that of the monomers droplets. Consequently, the probability for a radical to enter into monomer droplets is very low, and most particles are formed by either homogeneous or heterogeneous nucleation. 34 Once they are nucleated, the polymer particles undergo substantial growth by polymerization. The monomer required for the polymerization must be transported from the monomer droplets by diffusion through the aqueous phase. In some cases, this represents a severe limitation of the conventional emulsion polymerization. Thus, water resistance of coating prepared from dispersed polymers is significantly improved if very hydrophobic monomers, e.g. lauryl and stearyl methacrylate are incorporated into polymer backbone. However, mass transfer of this monomer from monomers droplets to polymer particles through the aqueous phase is diffusionally controlled, and hence they cannot be readily incorporated into the polymer in conventional emulsion polymerization (Asua, 2002). 2.5.2 Miniemulsion The formulation and application of polymeric nanoparticles enjoy great popularity in academia and industry. The techniques of macroemulsion and microemulsion polymerization which are usually used for the preparation of polymer particle are based on a kinetic control during preparation, the particles are built from the centre to the surface, and the particle structure is governed by kinetic factors. Because of the dictate of kinetics, serious disadvantages such as lack of homogeneity and restrictions in the accessible composition have to be accepted. The concept of nanoreactors for the formation of the nanoparticles can take advantage of a potential thermodynamic control for the design of nanoparticles. This means that the droplets have to become the primary locus of the nucleation of the polymer reaction. The polymerization in such nanoreactors should take place in a highly parallel fashion, i.e. the synthesis is performed in 1018 – 1020 nano-compartments which are separated from each other by a continuous phase (Asua, 2002). In miniemulsion polymerization, the principle of small nanoreactor is realized as demonstrated in Figure 2.5. 35 Surfactant Water Monomer Polymer Monomer Water Miniemulsification (ultrasonification) Figure 2.5: Principle of miniemulsion polymerization (Asua, 2002) Miniemulsions are dispersions of critically stabilized oil droplets with a size between 50 and 500 nm prepared by shearing a system containing oil, water, a surfactant, and a hydrophobe agent. Polymerization in such miniemulsions, where carefully prepared, result in latex particles which are about the same size as the initial droplets. This means that the appropriate formulation of a miniemulsion suppresses coalescence of droplets or Ostwald ripening. The polymerization of miniemulsions extends the possibilities of the widely applied emulsion polymerization and provides advantages with respect to copolymerization reactions of monomers with different polarity, incorporation of hydrophobic materials or with respect to stability of the formed latexes. 2.5.3 The Miniemulsion Process A system where small droplets with high stability in a continuous phase are created by using high shear is classically called a “miniemulsion”. One of the tricks for obtaining stability of the droplets is the addition of an agent which dissolves in 36 the dispersed phase, but is insoluble in the continuous phase. The small droplets can be hardened by either a subsequent polymerization by decreasing the temperature (if dispersed phase is a low-melting-point material). For a typical oil-in-water miniemulsion, an oil, a hydrophobic agent (or several), an emulsifier and water are homogenized by high shear to obtain homogenous and monodisperse droplets in the size range of 30 – 500 nm. In the first step of the miniemulsion process, small stable droplets in the size range between 30 and 500 nm are formed by shearing a system containing the dispersed phase, the continuous phase, a surfactant, and an osmotic pressure agent. In a second step, these droplets are polymerized without changing their identity. On the basis of the principle on miniemulsion, the preparation of new nanoparticle which could not be prepared in heterophased processes is now possible. That also includes, in particular, the encapsulation of inorganic material such as CdS semiconductor. For creating a miniemulsion, the step of homogenization is of great importance since fairly monodisperse small droplets have to be achieved. The homogenization can be achieved by using an ultrasonicator for the miniemulsification of small quantities in a laboratory scale batch process. 2.5.4 Preparation of Miniemulsion Monomer miniemulsions suitable for miniemulsion polymerization are submicron monomer-in-water dispersion stabilized against both diffusional degradation and droplet coagulation by using a water-insoluble low-molecular weight (co-stabilizer) compound and an efficient surfactant. The key issues in the preparation of the monomer miniemulsions are the formulation and the method of preparation. 37 2.5.4.1 Formulations A typical formulation includes water, a monomer mixture, a co-stabilizer and the surfactant and initiator system. Table 2.5 shows that miniemulsions of monomers with a wide range of water solubility have been prepared including vinyl chloride, vinyl acetate, methyl methacrylic, butyl acrylic, styrene, vinyl hexanoate, dodecyl methacrylic , and stearyl methacrylic. In addition, multimonomer formulations have been prepared, including miniemulsions in which some amounts of completely water-soluble monomers such as acrylic acid and methacrylic acid have been used. Table 2 5: Examples of formulation of miniemulsion polymerization (Asua, 2002) Monomer Costabilizer Surfactant Initiator Vinyl acetate Hexadecane Polyvinyl Sodium persulfate alcohol Methyl Polymethyl Sodium lauryl sulfate Potassium persulfate methacrylate metacrylate Methyl Hexadecane Sodium lauryl sulfate Ammonium persulfate Styrene Hexadecane Potassium persulfate Styrene Polystyrene Styrene /methyl Hexadecane Sodium lauryl sulfate Sodium lauryl sulfate Sodium lauryl sulfate methacrylate methacrylate Potassium persulfate Ammonium persulfate /sodium hydrogen sulfite Miniemulsions have been stabilized with long-chain alkanes, alcohol and polymeric hydrophobic. Hexadecane and cetyl alcohol are the co-stabilizers most often used in preparation. However, these co-stabilizers remain in the polymer particles and may have deleterious effects on the properties of the polymer. Alducin et al. (1994) proposed to minimize these negative effects of the co-stabilizers by incorporating it into the polymer backbone by means of covalent bounds. A series of oil-soluble initiators with different water solubility were used. It was found that only lauroyl peroxide was water-insoluble enough to stabilize styrene miniemulsions 38 droplets. A drawback of this method is that the initiator concentration affects not only miniemulsion stability, but also polymerization rate and molecular weight distribution (MWD). In addition, it was found that rather low-molecular weights were obtained 20,000 – 40,000 g/mol. Anionic, cationic, non-ionic and mixed anionic/non-ionic, non reactive surfactants, as well as reactive surfactants have been used. The surfactants useful for miniemulsion polymerization should meet the same requirements as in conventional emulsion polymerization: (i) it must have a specific structure with polar and non-polar groups (ii) it must be more soluble in the aqueous phase so as to be readily available for adsorption on the oil droplets surface (iii) it must adsorb strongly and not easily displaced when two droplets collide (iv) it must reduce the interfacial tension to 5 x 10-3 N/m or less (v) it must impart a sufficient electro kinetic potential to the emulsion droplets (vi) it must work in small concentrations (vii) it should be relatively inexpensive, non toxic, and safe to handle. 2.5.4.2 Method of Preparation Figure 2.6 shows the experimental scheme most often used for miniemulsion preparation. The surfactant system is dissolved in water, the co-stabilizer is dissolved in the monomers and mixed under stirring. Then, the mixture is subjected to high efficient homogenization. Other methods have also been reported. Thus, when a long chain alcohol is used as co-stabilizer, the long chain alcohol is first mixed with water and surfactant at temperature higher than the melting point of the alcohol, the mixture is cooled to room temperature and sonicated to break up the gel phase. Then, monomers are added under stirring and the resulting mixture homogenized to form the miniemulsion. 39 Dissolve Dissolve emulsifier surfactant in water in water Mixing under stirring High efficient homogenization Dissolve costabilizer in monomer Figure 2.6: Schematic for monomer miniemulsion preparation methods 2.5.4.3 Homogenization Devices A variety of equipment is commercially available for emulsification. The most important are rotor-stator system, sonifiers and high-pressure homogenizers. Rotor-stator system and other shear devices rely on turbulence to produce the emulsification. The minimum droplets size that can be achieve with this type of equipment depends on the size of the smallest turbulent formed, which in turn depends on the geometry of the rotor-stator system and on the rotation speed. Sonication tip Sonication region Sonication time Figure 2.7: Schematic of sonication process 40 The sonifier produces ultrasound waves that cause the molecules to oscillate about their main position as the waves propagate. During the compression cycle, the average distances between the molecules decrease, while during rarefaction the distance increases. The rarefaction result is negative pressure that may cause the formation of voids or cavities (cavitations bubbles) that may grow in size. In the succeeding compression cycle of the wave, the bubbles are forced to contract and may be even disappear totally. The shock waves produced on the total collapse of the bubbles cause the break up of the surrounding monomer droplets. A problem associated with the sonifier is that, as shown in Figure 2.7, only a small region of the fluid around sonifier tip is directly affected by the ultrasound waves. In order to be broken up, the monomer droplets should pass through this region. Therefore, when sonication is used to form the miniemulsion, additional stirring must be used to allow all the fluid to pass through the sonication region. This process makes the miniemulsion characteristics dependent on the sonication time. Thus, there are evidence that droplet size decrease with sonication time. The decrease is initially pronounced and later the droplets size evolves asymptotically towards a value that depends on both the formulation and the energy input. 2.5.4.4 Particle Nucleation Droplet nucleation is unique of miniemulsion polymerization. The characteristics of miniemulsion that is charged into reactor depend on the formulation, homogenization procedure and storage time. In general, the miniemulsion is composed of submicron droplets stabilized with a surfactant against coalescence and with a co-stabilizer to minimize Ostwald ripening. These droplets are not equal in size. Depending on the amount of surfactant used in the formulation and on the homogenization procedure, micelles may be present. Assuming that a water-soluble initiator is used, particle nucleation involves the following seriesparallel processes: a) Radical formation in the aqueous phase by initiator decomposition b) Polymerization of the radicals in the aqueous phase to give oligomers of increasing hydrophobicity. 41 c) Once the oligoradicals are hydrophobic enough, they can enter into monomer droplets (droplet nucleation) and micelles (micellar nucleation). They may also grow to a length that makes them insoluble in water and precipitate (homogeneous nucleation). d) In addition to droplet nucleation, monomer droplets may disappear by coalescence with other droplets and polymer particles, as well as by diffusional degradation if the co-stabilizer is water-soluble enough to diffuse from small to large droplets. For practical reasons, it is important to maximize the fraction of particles generated by droplet nucleation. This fraction depends on the number of droplets and micelles and on the relative values of the rate coefficients for entry of radicals into monomer droplets and micelles, and the propagation rate in the aqueous phase. The presence of micelles depends on the amount of surfactant and the homogenization procedure. In miniemulsion polymerization using hexadecane, Hansen and Ugelstad (1979a, 1979b) observed that as the intensity of the homogenization increased, the number of polymer particles initially decreased and after going through a minimum increased again. At low homogenization intensity, big droplets were formed and free surfactant formed micelles that gave particles by micellar nucleation. As homogenization intensity increased, smaller droplets were formed and more surfactant was required to stabilize these droplets, leaving less free surfactant to form micelles. The number of particles decreased because micellar nucleation decreased. At a higher homogenization intensity, micelles disappeared and formation of particles occurred by droplets nucleation. Under these conditions, increasing homogenization intensity produced more droplets, and hence the number particles increased. Lim and Chen (2000) investigated the miniemulsion polymerization of styrene using a block copolymer as surfactant and hexadecane as co-stabilizer finding that a significant fraction of particles were formed by micellar nucleation at high surfactant concentrations. In the miniemulsion polymerization of vinyl chloride 42 Saethre et al. (1995) found that particles were produced by micellar nucleation when the concentration of free surfactant (taking into account the amount adsorbed on the droplets and particles) was above the CMC. 2.5.5 Applications The applications of polymerization miniemulsion is tabulated in Table 2.6. Table 2.6: Somes application of polymerization miniemulsion (Asua, 2002) Application Production of high solids low viscosity latexes Continuous polymerization reactors Controlled radical polymerization in dispersed media Catalytic polymerization Encapsulation of inorganic solids Incorporation of hydrophobic monomers Hybrid polymer particles, Polymerization in non-aqueous media Anionic polymerization Step polymerization in aqueous dispersed media Production of low molecular weight polymers in dispersed media Latexes with special particle morphology 2.5.5.1 Encapsulation of Inorganic Material Encapsulation of inorganic particles in polymer matrices is of interest in cosmetics, pharmaceuticals, paint and for reinforcing filler particles for polymer. Attempts to encapsulate polymer particles using conventional emulsion 43 polymerization have been reported but it was found difficult to locate the dominant polymerization at the surface of the inorganic material. Figure 2.8 shows the principle of the encapsulation of inorganic materials in polymer particles by the miniemulsion process. inorganic material surfactant 1 surfactant 2 monomer polymer monomer water addition of water and surfactant 2 stirring ultrasound polymerization Figure 2.8: The principle of the encapsulation of inorganic materials in polymer particles by miniemulsion process (Landfester and Ramirez, 2003) Nunes et al. (2006) synthesized poly(ethylmethacrylate-co-methacrylic acid) magnetite particle via miniemulsion polymerization. Firstly, the hydrophobization magnetite was prepared by adding oleic acid into magnetite particles and then the miniemulsion was prepared using EMA-co-MAA, SDS, and water. The SEM analysis result showed that the magnetite was dispersed within the polymer matrix with the linear dimension below 500 nm. Faridi-Majidi et al. (2006) studied encapsulation of magnetite nanoparticles with polystyrene via emulsifier-free miniemulsion polymerization. The miniemulsion was prepared using hexadecane as hydrophobe agent, monomer and water. The TEM result proved the presence of 44 magnetite in polymer particles which appeared to be monodisperse in size approximately 100 – 300 nm. Tong and Deng (2007) reported the synthesis of polystyrene encapsulated nanosaponite composites via miniemulsion polymerization. Firstly, the saponite nanoparticle was modified by adding vinyl benzyl trimethylammonium chloride (VBTAC) in water solution. Secondly, the miniemulsion system was formed using styrene, hexadecane, TX-405 and water. The results showed that the VBTAC modified nanosaponite could be fully exfoliated and encapsulated inside polystyrene latex. When the concentration of hexadecane as co-stabilizer was high, the final composite particles were composed mainly of spherical particles with sizes less than 100 nm and a small number of hemispherical or bowl-structured particles of size 100 – 1000 nm. Erdem et al. (2000a, 2000b) prepared aqueous miniemulsion of the styrene containing TiO2 dispersed particles by both sonication and a high-pressure homogenizer. Sodium lauryl sulfate was used as surfactant and hexadecane plus polystyrene the costabilizer system. It was found that the presence of TiO2 particles within the miniemulsion droplets resulted in a significant increase in the droplets size (measured by the soap titration method). Increasing surfactant concentration did not result in a significant change in the size of the miniemulsion droplets. An increase in the loading of the TiO2 led to larger miniemulsion droplets. It was also found that the use of a more efficient homogenization (high pressure homogenization vs sonication) could not significantly reduce the droplets size. The reasons for these results were not provided. The polymerization of the miniemulsion trying to encapsulate the TiO2 particles was also investigated. It was found that the extent of the encapsulation strongly depended on the quality of dispersion of the TiO2 particles in the monomer. The maximum encapsulation efficiencies (83 % coverage) were obtained with the smallest TiO2 particles (hydrophilic particles stabilized with 1 % of polybutene-succinimide pentamine). Hydrophobic TiO2 or the use of fewer stabilizers led to larger TiO2 particles that resulted in lower encapsulation efficiencies. The failure to achieve 100 % encapsulation was attributed 45 to: (i) formation of new particles by homogenous nucleation; (ii) presence of large TiO2 aggregates that may not be contained inside the monomer droplets; and (iii) less than 100 % nucleation of droplets. A higher carbon black loading was achieved by Tiarks et al. (2001). The author dispersed the carbon black in an aqueous surfactant solution by using sonication. A miniemulsion of styrene and a polyurethane co-stabilizer was prepared using sonication. Then, the miniemulsion was mixed with the aqueous dispersion of carbon black and the resulting dispersion was sonicated. The resulting dispersion was polymerized with AIBN obtaining stable latexes that contained larger fractions of carbon black (40%) with a good coverage. It was found that particle size was not dependent on surfactant concentration, but the process is sensitive to the type and amount of co-stabilizer used. Hexadecane was less efficient than polyurethane and there is an optimum amount of polyurethane, below which coagulation occurs and above which new particles are formed. The reason for this behaviour remained unclear. Also an optimum monomer/carbon black fraction was found. Larger amounts of monomer yielded pure polymer particles and smaller amount resulted in uncovered carbon black particles. 46 CHAPTER 3 EXPERIMENTAL 3.1 Synthesis and Preparation 3.1.1 Purification of Monomer Monomer usually contains stabilizer and it can affect the polymerization process. It is important to purify the styrene and divinyl benzene by removing the stabilizer with 10 % (w/v) NaOH solution before the polymerization process. These monomers were washed with 20 mL of 10 % (w/v) NaOH in a separating funnel. Two layers were formed and the lower layer which contains NaOH solution with stabilizer was removed. The washing process was repeated 3 times to ensure that all the stabilizer in the monomer was removed. The monomer was washed with distilled water until neutral before it was kept in a dry bottle and stored in a refrigerator. Monomer methacrylic acid, monomer ethylene glycol dimethacrylic acid and all the other chemicals were used without further purification (Antolini et al., 2005; FaridMajidi et al., 2006; Kuljanin et al., 2006) 3.1.2 Preparation of CdS Nanoparticles by Reverse Micelles Using n-Decane as Oil-phase CdS nanoparticles was synthesized by reverse micelles in miniemulsion system as follows. Aqueous standard solutions of cadmium nitrate (0.1 M) and 47 sodium sulfide (0.1 M) were first prepared. These solutions were then used to prepare two other solutions : (1) About 45 μL of the Cd(NO3)2 solution was injected into 5 mL of 0.1 M cetyl trimethyl ammonium bromide (CTABr) solution in mixtures of 2.5 mL of 2-propanol and 2.5 mL of n-decane, and (2) About 45 μL of the Na2S solution was injected into 5 mL of 0.1 M CTABr solution in mixtures of 2.5 mL of 2-propanol and 2.5 mL of n-decane. After stirring for 1 h, they were mixed together for 5 minutes resulting in the formation of CdS nanoparticles in miniemulsion. For obtaining CdS nanoparticles at different Wo = [water]/[surfactant], and Po = [2-propanol]/[surfactant] values, different amounts of the aqueous solutions of 0.1 M Cd(NO3)2 and Na2S were added into 5 mL of 0.1 M CTABr solution mixtures of various compositions of 2-propanol and n-decane. Stability of the nanoparticles formed were monitored by UV – Vis spectroscopy and the rate of precipitation of CdS particle (Khiew et al., 2003, 2004; Agostiano et al., 2000). The particle size of CdS nanoparticle was also determined by means of UV – Vis spectroscopy. The absorption onset wavelength (λ), that is calculated from the UV – Vis spectra was converted into a CdS particle size using Brus equation (Brus, 1984). 3.1.3 Preparation of CdS Nanoparticles by Reverse Micelles Using Monomer as Oil-phase CdS nanoparticles were synthesized by reverse micelles in miniemulsion with the following steps. Aqueous standard solutions of cadmium nitrate (0.1 M) and sodium sulfide (0.1 M) were foremost prepared. These solutions were then used to prepare two other solutions: (1) about 45 μL cadmium ion solution was injected into 5 mL of 0.1 M CTABr solution in mixture of 2.5 mL of 2-propanol, 2.0 mL of styrene, and 0.5 mL divinylbenzene and (2) about 45 μL sulfide ion solution was injected into the same solution as above. After stirring for 1 h, they were mixed together for 5 minutes resulting in the formation of CdS nanoparticles. To obtain CdS nanoparticles by reverse micelles in miniemulsion at different Wo, different amount of the aqueous solutions of Cd(NO3)2 and Na2S were added into the same solution as mentioned above. The stabilization and particle size of the CdS nanoparticles were determined by the same procedure at 3.1.2. The CdS/P(S-DVB) 48 nanocomposites were synthesized via in-situ polymerization in miniemulsion at 70 – 80 oC for 6 h using 2,2’-Azobisisobutyronitrile (AIBN) as the initiator. The nanocomposites were filtered and then washed with water, ethanol and methanol, to successively remove excess of monomer, cadmium and sulfide ion. 3.1.4 Preparation of P(MAA-EGDMA) The P(MAA-EGDMA) polymer was synthesized by in-situ polymerization in miniemulsion system. The miniemulsion system was prepared using cetyl triammonium bromide (CTABr) as surfactant, 2-propanol as co-surfactant, water, methacrylic acid-ethylene glycol dimethacrylic acid (MAA-EGDMA) monomer and n-decane acted as the oil-phase. The preparation involved stirring and polymerization of a mixture of 0.3645 g of CTABr, 2.5 mL of 2-propanol, 1.0 g of MAA monomer, 0.5 g of EGDMA monomer, 2.5 mL n-decane and water. After stirring for 1 h, it was followed by in-situ polymerization at 70 – 80 o C for 6 h using 2,2’- Azobisisobutyronitrile (AIBN) as the initiator. The polymer was filtered and then washed with water, ethanol and methanol, to remove the surfactant and monomer residue from the external surface of polymer and finally dried for 4 h at 100 oC. 3.1.5 Preparation of CdS/P(MAA-EGDMA) Nanocomposites The preparation of occluded cadmium inside P(MAA-EGDMA) was carried out by ion exchange method. This preparation involved stirring a mixture of 0.5 gram P(MAA-EGDMA) with various mmol of cadmium ion in methanol solution for 24 h. The solution was then filtered and washed with water, ethanol and methanol, to successively remove excess cadmium ion from the external surface and then dried for 4 h at 100o C. The final step was synthesis of CdS/P(MAA-EGDMA) using precipitation reaction between Cd2+/P(MAA-EGDMA) with sulfide ion solution. About 0.5 gram Cd2+/P(MAA-EGDMA) was soaked in a solution of various mmol 49 Na2S. 9 H2O in 50:50 H2O: methanol. The mixture was stirred for 24 h before filtering the CdS/P(MAA-EGDMA) and washed with water, ethanol and methanol, and dried at 100o C for 4 h. 3.1.6 Preparation of P(S-DVB), SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) Pure poly(styrene-divinylbenzene) [P(S-DVB)] was synthesized by in-situ polymerization in miniemulsion system. Meanwhile the miniemulsion system was prepared using sodium dodecyl sulfate (SDS) as surfactant, 2-propanol as costabilizer, water, and styrene-divinylbenzene as oil-phase. The preparation involved stirring a mixture of 2.5 mL of 2-propanol, 1.80 gram of styrene, 0.60 gram of divinylbenzene and 2,2’-azobisisobutyronitrile (AIBN) for 1 h. After stirring for 1 h, the mixture was added to 20 mL 0.1 M of SDS in water and then ultrasonication was carried out for 15 min followed by more stirring for 1 h. Then in-situ polymerization was done at 70 – 80 o C for 6 h. The polymer was filtered and then washed with water, ethanol and methanol, to remove surfactant and monomer residues, before drying for 4 h at 100o C. The next step is the sulfonation reaction of the polymer with H2SO4/SO3. In the preparation of SO3H-P(S-DVB), 10.0 mL H2 SO4/SO3 was added slowly to 0.5 gram of P(S-DVB). The mixture was slowly stirred and maintained at 50 oC for 1 h, followed by careful dilution of the mixture with water. The SO3H-P(S-DVB) was filtered and then washed with water until neutral pH was obtained and finally dried at 100 oC for 4 h. The preparation of occluded cadmium ion on the surface of SO3HP(S-DVB) was achieved by ion exchange method. This preparation involved stirring a mixture of 0.5 gram SO3H-P(S-DVB) with various mmol of cadmium ion in methanol solution for 24 h. The solution was filtered and then washed with water, ethanol, and methanol, to effectively remove excess cadmium ion and then dried for 4 h at 100o C. The final step was the synthesis of CdS/SO3-P(S-DVB) nanocomposites by precipitation 50 reaction of Cd2+/SO3H-P(S-DVB) with sulfide anion solution. About 0.5 gram of Cd2+/SO3H-P(S-DVB) was soaked in a solution of various mmol Na2S. 9 H2O in 50:50 H2O: methanol and then was stirred for 24 h. The nanocomposites was filtered and then was washed with water, ethanol and methanol, successively and dried at 100 oC for 4 hours. The sulfonate content (degree of sulfonation) of SO3H-P(S-DVB) was determined by the following ways. To perform titration, 0.1 gram of SO3H-P(S-DVB) was dissolved in mixture of 15 mL of toluene and 2 mL ethanol. The solution was titrated with standardized methanol solution of potassium hydroxide (KOH), using phenolphthalein as indicator. The ion exchange capacity of SO3H-P(S-DVB) was determined (with unit of meq/gram of dry polymer of sulfonated polymer) by measuring concentration of hydrogen (H+) that was exchanged with sodium cation when acid sulfonated polymer samples were equilibrated with sodium (Na+) solution. 0.1 gram sample of acid polymer was placed into 100 mL of 0.2 mL NaCl solution and shaken sporadically for 24 h, the amount of hydrogen ion released by the polymer was determined by titration with 0.01 M NaOH. 3.2 Characterization Techniques The as-synthesized polymer nanocomposites were characterized using X-ray Diffraction (XRD), Fourier Transformation Infrared (FTIR), Field Emission Scanning Electron Microscope (FESEM), Transmission Electron Microscopy (TEM), UV – Visible Diffuse Reflectance (UV- Vis DR), Thermal conductivity and Thermogravimetric analyses (TGA). 3.2.1 UV-Vis Spectroscopy The stabilization and particle size of CdS nanoparticles in miniemulsion system were characterized using UV – Vis spectrocospy. Stability of the 51 nanoparticles formed were monitored by UV – Vis spectroscopy and the rate of precipitation of CdS particle. The particle size of CdS nanoparticle was also determined by means of UV – Vis spectroscopy. The absorption onset wavelength (λ), that is calculated from the UV – Vis spectra was converted into a CdS particle size using Brus equation (Brus, 1984). The UV – Vis absorption spectra of the solutions were recorded on a Perkin Elmer Lambda 900 UV/Vis/NIR spectrometer in the wavelength range of 350 – 700 nm using a 10 mm cuvette (Agostiano et al., 2000; Khiew et al., 2003, 2004). 3.2.2 X-Ray Diffraction The structure of the nanocomposites was characterized by XRD using a Bruker Advance D8 Difractometer with Cu Kα (λ = 1.5405 A) radiation as the diffracted monochromatic beam at 40 kV and 40 mA and was scanned in the 2θ range between 20o – 70 o at ambient temperature. The sample was ground to a fine powder using a mortar and then 1 g of dried sample was mounted on a plastic sample holder. The sample was then pressed and smoothed on the sample holder using a glass slide (10×12 mm2) with 2 mm width. The analysis was carried out in the step interval of 0.05o with counting time of 1 second per step. (Ghosh et al., 2005a, Hirai and Ota, 2006). 3.2.3 Fourier Transform Infrared Spectroscopy The pure polymer, sulfonated polymer and CdS/polymer nanocomposites were characterized by FTIR spectroscopy using a Perkin Elmer Spectrum One FTIR Spectrometer at room temperature and the spectra were recorded in the region of 4,000 – 400 cm-1. Preparations of samples for FTIR characterization were done using KBr pellet technique. A small amount of solid sample was mixed with KBr with the ratio of KBr to sample of 100:1 and was ground well to powder form. Then the sample was compressed under 8 tons for 1 - 2 minutes, producing transparent pellet 52 with 13 mm in diameter. FTIR spectra were recorded at room temperature with 4 cm-1 resolution (Chen et al., 2006, El-Tantawy et al, 2004). 3.2.4 UV-Vis Diffuse Reflectance Spectroscopy The structure and particle size of CdS nanoparticles at CdS/polymer nanocomposites were determined by UV – Vis Diffuse Reflectance (UV – Vis DR) using Perkin Elmer Lambda 900 UV/Vis/NIR spectrometer. In preparation of the samples for UV - Vis DR, the samples were finely ground and placed on a sample holder to be analyzed. The scanned wavelength ranges between 350 – 700 nm (Zhang et al,. 2001) 3.2.5 Field Emission Scanning Electron Microscope Field Emission Scanning Electron Microscope (FESEM), Zeiss Supra 35VP VP Series model was used to determine the morphology of nanocomposite sample. Samples were mounted on stubs using double-sided tape and then coated with gold using instrument model BIO-RAD Polaron Divison SEM Coating System machine at 10-1 mbar with 30 mA for 75 minutes. Gold coating is needed to prevent charge build-up on the sample surface, besides increasing secondary electron emission. Then, the stubs were put in field emission scanning electron microscope model Philip XL40 and with pressure of 5 bars. Electron source was attained from tungsten filament. The FESEM micrograph was recorded with resolution of 10 kV for certain times of magnification. 3.2.6 Transmission Electron Microscopy A JEOL JEM-2100 Transmission Electron Microscope (TEM) was used for analyzing the morphology and particle size of CdS polymer nanocomposites. In 53 preparation of the samples for TEM analysis, the nanocomposite was dispersed in ethanol by ultrasonic stirring. Then a drop of the suspension was placed on the grid with Formvar thin film, before carrying out the TEM analysis. 3.2.7 Dielectric Constant and Dissipation Factor The capacitance and dissipation factor of nanocomposite were determined using Impedance Analyzer. Pellets of 2–4 mm thick were prepared by placing sufficient amount of sample (~50 mg) in a steel die measuring 13 mm in diameter, and a pressure of 5 tons were applied and held for 30 seconds. The dielectric constants (εr) of the pellet prepared were calculated from the measured capacitance at various frequencies (0.1 kHz to 1,000 kHz) using AC Impedance Analyzer using the equation: C = εr εo A/d, where εo is the dielectric constant of the free space (8.854 x 10-12 F/m); “A” the area of the electrical conductor; and “d” the thickness of material. 3.2.8 Ionic Conductivity The ionic conductivity was studied using AC Impedance Analyzer with frequency response analyzer (Auto Lab POST AT 30). The resistances of the specimens were measured at frequency of 0.1 kHz – 1,000 kHz using AC Impedance Analyzer. The ionic conductivity, σ, of the polymer nanocomposites were calculated from the resistance data. 3.2.9 Thermal Conductivity The thermal conductivity was measured by Thermal Conductivity Analyzer . Model Therm Test Mathis TC30. Before using, the Thermal Conductivity Analyzer was calibrated using High Density Polyethylene (HDPE) 0.5771 W/mK and Lexan 54 0.25409 W/mK and then, the thermal conductivity of the samples were measured at room temperature. 3.2.10 Thermogravimetric Analysis Thermogravimetric analysis (TGA) was conducted on Mettler Toledo TGA/SDTA instrument. The sample was fine ground and place in a 70 μL pan. The procedures included setting 10-15 mg of sample to be heated from 25 oC up to 1000 o C, with heating rate of 10 oC/min in oxygen atmosphere. 3.2.11 Atomic Absorption Spectroscopy The cadmium (Cd) and sodium (Na) content of the polymer nanocomposite samples were determined using Perkin Elmer AA400 Atomic Absorption Spectrometer equipped with flame atomizer and EDL lamp for Cd and Na. The samples were prepared by decomposing of 50 mg sample in the furnace and then added with 5.0 mL of concentrated HCl solution. Samples solution was diluted by adding in double deionized water to the mark, and homogenized by shaking. Cadmium and sodium in samples were calculated by using standard calibration plot shown in Appendix F and Appendix G. 55 CHAPTER 4 RESULTS AND DISCUSSION This chapter describes the synthesis, characterization, electrical properties and the correlation between the structural and physicochemical properties of three cadmium sulfide (CdS)/polymer composites. This chapter is divided into four sections: • The synthesis, characterization and properties of cadmium sulfide nanoparticles encapsulated in poly(styrene-divinyl benzene) [P(S-DVB)] to produce CdS/P(SDVB) by in-situ polymerization in a miniemulsion system. (Section 4.1) • The synthesis, characterization and properties of cadmium sulfide/poly(methacrylic acid-ethylene glycol dimethacrylic acid) [CdS/P(MAAEGDMA)] nanocomposites. (Section 4.2) • The synthesis, characterization and properties of cadmium sulfide/sulfonatedpoly(styrene-divinyl benzene) [CdS/SO3-P(S-DVB)]. (Section 4.3) • Comparison of physicochemical and electrical properties of those CdS/polymer nanocomposites. (Section 4.4) Firstly, the synthesis of cadmium sulfide (CdS) nanoparticles encapsulated in poly(styrene-divinyl benzene) [P(S-DVB)] by in-situ polymerization in a miniemulsion system is described. In this chapter, the effect of Wo = [water]/[surfactant], Po = [2-propanol]/[surfactant], bulk concentration of cadmium or sulfide ion, and monomer on the stability and particle size of CdS nanoparticle in reverse micelles system using n-decane as oil-phase was studied. The stability of 56 CdS nanoparticle in reverse micelles system was analized semi-quantitatively by observing the initial time when CdS aggregation occurred and by means of UV – Vis spectra of the miniemulsion. The effect of Wo = [water]/[surfactant] at Po = [2propanol]/[surfactant] = 32.5 on the stability and particle size of CdS in reverse micelles system using monomer as oil-phase is also investigated. Following this, synthesis methods and properties of CdS/P(S-DVB) nanocomposites i.e. by in-situ polymerization in two miniemulsion system using n-decane and monomer as oilphase in the presence of CdS nanoparticles dispersed in reverse micelles system are compared and discussed. Secondly, the synthesis, characterization and properties of cadmium sulfide/poly(methacrylic acid-ethylene glycol dimethacrylic acid) [CdS/P(MAAEGDMA)] nanocomposites are detailed. This section is devoted to the synthesis of P(MAA-EGDMA) by in-situ polymerization in miniemulsion system and CdS/P(MAA-EGDMA) by post synthesize route (by ion exchange and precipitation with various concentrations of cadmium and sulfide ions). Subsequent sections discuss the synthesis, characterization and properties of cadmium sulfide/sulfonated-poly(styrene-divynilbenzene) [CdS/SO3-P(S-DVB)] nanocomposites including the synthesis P(S-DVB) as polymer host by in-situ polymerization, the synthesis SO3H-P(S-DVB) by sulfonation process with H2S2O7 and the preparation of CdS/SO3-P(S-DVB) nanocomposites by post synthesize with ion exchange and precipitation processes with various concentration of cadmium and sulfide ion. The schematic representation of results of the studies is outlined in Figure 4.1. To understand the structural and physicochemical-electrical properties, appropriate techniques are used in the characterization, including UV – Vis spectroscopy for determination of the optical properties, stability and particle size of CdS nanoparticles in the miniemulsion. Meanwhile, the UV – Vis DR spectra was to estimate the particles size of CdS in the polymer nanocomposites, along with FTIR spectroscopy for the polymer characterization and verification of the interaction between polymer and CdS nanoparticles. The CdS nanoparticles and the polymer 57 nanocomposites prepared in this study was subsequently characterized by XRD to determine structure, SEM and TEM to determine the morphology, particles size of polymer and average crystallite size of CdS, while AAS analysis was performed to determine concentration of cadmium ion. In addition, results of studies on dielectric properties (dielectric constant and dissipation factor), ionic conductivity, thermal conductivity, and thermal properties were also detailed and discussed in this chapter. CdS P(S-DVB) In-situ polymerization COO P(MAA-EGDMA) – CdS Ion exchange, COO – precipitation SO3H-P(S-DVB) CdS Ion exchange, precipitation SO3Na HSO3 HSO3 SO3Na HSO3 NaSO3 HSO3 SO3Na NaSO3 HSO3 HSO3 SO3Na SO3Na HSO3 NaSO3 HSO3 SO3Na HSO3 SO3Na HSO3 Figure 4.1: Schematic representation of mechanism of CdS nanoparticles attachment and encapsulation on polymer surface 58 4.1 Optimization of Synthesis of CdS Nanoparticles and CdS/P(S-DVB) Nanocomposites A prerequisite for the in-situ polymerization in miniemulsion system is the stability of the initial miniemulsion and stability during the polymerization process. The stability of the miniemulsion has been shown to be strongly affected by parameters such as Wo = [water]/[surfactant], Po = [2-propanol]/[surfactant], bulk concentration of ion, temperature, the number and properties of all the components in the reaction mixture (Agostiano, et al. 2000). In this research, we attempted to synthesize CdS nanoparticles by reverse micelles in a miniemulsion system using cetyltrimethylammonium bromide (CTABr) as surfactant, 2-propanol as co-stabilizer, n-decane as oil phase, and water. A typical CTABr surfactant has low solubility in oil-alcohol system and the solubility usually increases with addition of water. In a miniemulsion system, the stability of CdS nanoparticles in reverse micelles was found to be dependent on Wo = [water]/[surfactant], Po = [2-propanol]/[surfactant] and bulk concentration of cadmium ion. For investigation of the influence of Po value, Wo, concentration of Cd2+ and S2- solutions, and effect of monomer on the stabilization time of CdS nanoparticles, various miniemulsion systems were examined under different conditions. Table 4.1 summarizes the optimum parameters that were obtained in this research including the absorption onset wavelength, and mean diameters of CdS nanoparticles. Under the optimum conditions, stabilization of CdS nanoparticles in the miniemulsion occurs at Wo = 5, [Cd2+] = [S2-] = 3 x 10-4 M and Po = 65. In the present study the stabilization time increased when the Po value increased from 16 to 65, then decreased beginning at Po = 78. Stability of the nanoparticles formed were monitored by UV – Vis spectroscopy and the rate of precipitation of CdS particle (see Appendix A and B). The particle size of CdS nanoparticle was also determined by means of UV – Vis spectroscopy. The absorption onset wavelength (λ), that is calculated from the UV – Vis spectra was converted into a CdS particle size using Brus equation (Brus, 1984). The absorption onset wavelength and calculated particle 59 size of CdS nanoparticles at the optimum condition were 480 nm and 6.3 nm, respectively (see Appendix C). Table 4.1: The optimum conditions for synthesis of CdS nanoparticle using n-decane as oil-phase a Parameter Values Remarks Po b 65 d Onset λ = 480 nm Particle size = 6.3 nm Wo c Concentration of Cd2+ and S2- a 5e f 3 x 10-4 M Amount of styrene 2.0 mL Increase stability of Amount of divinyl benzene 0.5 mL miniemulsion, see appendix E The stability of CdS nanoparticles in miniemulsion was monitored by UV – Vis spectroscopy (see appendices A and B) b c Wo = [water]/[surfactant] d e Po = [2-propanol]/[surfactant] The experiment was carried out in the range 16 – 105 at Wo = 5, C = 3 x 10-4 M (see appendix B) The experiment was carried out in the range 5 and 10 at Po = 65 (see appendix C) f The experiment was carried out in the range 3 x 10-4 M and 6 x 10-4 M (see appendix D) For investigation of the influence of Wo the miniemulsion system at Wo = 5 and 10 with concentration of cadmium and sulfide ions of 3.0 x 10-4 M was observed at different time interval after formation (see Appendix C). The red-shift suggested that at lower Wo CdS nanoparticles in miniemulsion are more stable compared to that at a higher Wo value. The UV – Vis spectra indicate that the larger the values of Wo, the larger the micelle core and the farther red-shifted was the CdS nanoparticle samples relative to those of the bulk samples. For investigation of the influence of concentration of cadmium and sulfide ions on nanoparticles size, the miniemulsion systems at Wo = 5 and Po = 65 with Cd2+ and S2- concentrations ions of 3.0 x 10-4 M and 6.0 x 10-4 M, respectively, were investigated (see Appendix D). From the UV – Vis spectra, it can be seen that as the 60 quantity of Cd2+ and S2- ion decreases, the absorption onset wavelength becomes more red-shifted suggesting that CdS nanoparticles in miniemulsion are more stable at lower ion concentration than at higher concentration, which leads to aggregation of CdS nanoparticles with larger diameters. This phenomenon was an indication that the smaller droplets of the miniemulsion with concentration of 3.0 x 10-4 M Cd2+ and S2- could have resulted from the higher stability and low flocculation rate of the CdS nanoparticles. The effect of addition of monomer namely, styrene and divinyl benzene on the size of CdS nanoparticles in miniemulsion system was investigated (see Appendix E). The results showed that addition of monomer to the miniemulsion system increased the stabilization of reverse micelles. The absorption onset wavelengths after addition of monomer were significantly blue-shifted compared to bulk CdS. It is concluded that the monomer could be used to replace n-decane as oilphase in the present miniemulsion system. CdS nanoparticles has also been synthesized by reverse micelles in miniemulsion using cetyltrimethylammonium bromide (CTABr) as surfactant, 2propanol as co-surfactant, styrene-divinyl benzene as oil-phase and water phase. Schematic representation of the synthesis of CdS nanoparticles and CdS/P(S-DVB) nanocomposites is illustrated in Figure 4.2. While, the schematic reaction of styrene with divinyl benzene and representation of the encapsulated CdS in P(S-DVB) matrices are shown in Figures 4.3 and 4.4 respectively 61 CTABr, monomer (M) 2-propanol, Cd2+/H2O M M M M M M M M M M M + M Cd2+ M M M Divinyl benzene M M M M M M M M M CdS M M M M M M M M M M M M M M M 2- M S2- M M Styrene M M M M M M M Surfactant S M M M M M M M M M M Monomer M M Cd2+ M M CTABr, monomer (M), 2propanol, S2-/H2O M • M M M M M M M CdS nanoparticle in miniemulsion Polymerization Polymer P(S-DVB) CdS Figure 4.2: Schematic representation of the synthesis of CdS nanoparticles by reverse micelles in miniemulsion system and CdS/P(S-DVB) nanocomposites via insitu polymerization 62 CH CH2 CH CH2 CH CH2 AIBN, 70oC + Styrene in miniemulsion system Divinyl benzene Figure 4.3: Reaction of styrene and divinylbenzene by free radical polymerization into P(S-DVB) CdS Polymer chain Figure 4.4: Schematic representation of the encapsulated CdS in P(S-DVB) matrices n 63 The optimum parameters for the synthesis of CdS nanoparticles in miniemulsion are shown in Table 4.2. Table 4.2: The optimum conditions for synthesis of CdS/P(S-DVB) nanocomposite using monomer as oil-phase a a Parameter Values Results Po b 32.5 d Onset λ = 484, particle size = 5.1 Wo c 1.5 – 7.0 e Concentration of Cd2+ and S2- 3 x 10 -4 M Monomer styrene 2 mL Monomer divinyl benzene 0.5 mL The optimum condition miniemulsion for synthesis of CdS nanoparticles was determined by UV – Vis spectroscopy (see Figure 4.4 – 4.5) b c Wo = [water]/[surfactant] d e Po = [2-propanol]/[surfactant] The experiment was carried out at Wo = 5.55 The experiment was carried out at Po = 32.5 The stability and particle size of the CdS nanoparticles in miniemulsion system were monitored by UV –Vis spectroscopy. The stability of CdS nanoparticles could be observed from the absorption onset wavelength of the UV – Vis spectra. Meanwhile, the particle size of the CdS nanoparticles was determined from the absorption onset wavelength and calculated using Brus equation (Equation 4.1) (Brus, 1984). Brus equation : E (R) = 2.43 + 2.446/R2 – 0.3031/R (Equation 4.1) with Ebulk CdS = 2.43 eV, E = hc/λ and R = particle size Figure 4.5 shows the effect of the Wo value of the reverse micelles solution on UV – Vis spectra of CdS nanoparticles. The results show that the absorption onset wavelength and the size of the CdS nanoparticles are successfully controlled by 64 changing the Wo values. The results indicate that the reverse micelles in miniemulsion can produce CdS nanoparticles with increased stabilization and their quantum size effect were also maintained as shown by the blue shifting of the absorption onset wavelength. The absorption onset wavelength and particle size of CdS nanoparticles in miniemulsion at Wo values = 1.5, 2.75, 4.0 and 5.55 determined from the UV – Vis spectra are 445 nm (3.2 nm), 472 nm (4.5 nm), 480 nm (4.9 nm) and 484 nm (5.1 nm), respectively. These results indicated that the particle size of CdS nanoparticles changed as a function of water/surfactant molar ratio. Relative intensity (a.u) (a) (b) (c) (d) 390 420 460 500 Wavelength (nm) 540 580 600 Figure 4.5: UV – Vis spectra of CdS nanoparticles in miniemulsion system using styrene-divinylbenzene as oil-phase at Po = 32.5 after 5 minutes formation and different Wo values (a) 5.50, (b) 4.0, (c) 2.75, and (d) 1.50 The stabilization of CdS nanoparticles in miniemulsion at different Wo values and various time intervals after formation was characterized by UV – Vis spectroscopy as shown in Figure 4.6. 65 (a) Relative intensity (a.u) after 5 min after 1 h a fter 24 h 380 420 460 500 540 580 600 540 580 600 Wavelength (nm) (b) after 5 min Relative intensity (a.u) after 1 h after 24 h 390 420 460 500 Wavelength (nm) Figure 4.6: UV – Vis spectra of CdS nanoparticles in miniemulsion using styrenedivinylbenzene as oil-phase at different Wo values and Po = 32.5 at different time intervals after formation (a) Wo = 1.5 (b) Wo = 2.75 (c) Wo = 4 and (d) Wo = 5.55 66 ( c) Relative intensity (a.u) after 5 min after 1 h after 24 h 390 440 520 480 560 600 Wavelength (nm) Relative intensity (a.u) (d) after 5 min after 24 h after 1 h 390 420 460 500 540 580 600 Wavelength (nm) Figure 4.6 (continue): UV – Vis spectra of CdS nanoparticles in miniemulsion using styrene-divinylbenzene as oil-phase at different Wo values and Po = 32.5 at different time intervals after formation (a) Wo = 1.5 (b) Wo = 2.75 (c) Wo = 4 and (d) Wo = 5.55 67 The UV – Vis spectra showed similar behavior in which the intensity of absorption of CdS nanoparticles in miniemulsion 5 to 60 minutes after formation is almost unchanged. But, 24 hours after formation, the absorption intensity of CdS nanoparticles decreased gradually showing that the stability of CdS nanoparticles in miniemulsion is also decreasing. These results are different from those reported by Agostiano et al. (2000) who synthesized CdS nanoparticles by reverse micelles in quaternary “water-in-oil” miniemulsion comprising CTABr/pentanol/hexane/water and CTABr/2-propanol/n-decane/water, respectively. They observed a significant decrease in optical intensity of the solution following precipitation of CdS nanoparticles in the miniemulsion. The present study demonstrated that miniemulsion using styrene-divinylbenzene as oil-phase could increase stability and favour the formation of CdS nanoparticles. 4.1.1 Structural and Morphological Properties of CdS/P(S-DVB) Nanocomposites Figure 4.7 shows the XRD patterns of pure P(S-DVB), CdS nanoparticles and CdS/P(S-DVB) nanocomposites with various CdS contents. The XRD pattern of the CdS sample exhibit broad peaks centred at 2θ values of 26.4o, 43.7o, 52o and a shoulder at~ 31.5o which could be indexed to scattering from 111, 220, 311 and 200 planes, respectively, of cubic CdS (Wang, et al. 2005). The XRD patterns of the CdS/P(S-DVB) nanocomposites are identical and they exhibit broad peaks centered at 2θ value of 26.4o. Broad XRD peaks are attributed to the absence of long-range order in the samples and reflect the small crystallite size (less than 10 nm as estimated from the Brus equation). This result agrees with the presence of nanocrystalline CdS, for which quantum size effect are expected. 68 26.4o 43.7o 52.0o CdS/P(S-DVB) 0.18 % Relative intensity (a.u) CdS/P(S-DVB) 0.03% CdS/P(S-DVB) 0.02% CdS nanoparticles Pure P(S-DVB) 20 30 40 2 Theta degree 50 60 70 Figure 4.7: XRD patterns of pure P(S-DVB), CdS nanoparticles and CdS/P(S-DVB) nanocomposites with various amounts of CdS 69 Comparing the XRD pattern of CdS nanoparticle with CdS/P(S-DVB) nanocomposites, one can see that the diffraction peak at 2θ = 26.4o is attributed to CdS nanoparticles. The weak reflection peaks around 26.4o should belong to CdS by comparison with the XRD pattern of CdS nanoparticles. to the amount of CdS nanoparticles is very small compared to polymer matrix. The low intensity peaks indicated that the amount CdS nanoparticle is very small. The polymer particles which are amorphous and present in large amount also contribute to the peak broadness. However the size of CdS nanoparticles could not determined by XRD due to the amount of CdS nanoparticles is very small compared to polymer matrix. The morphology of CdS/P(S-DVB) nanocomposites were characterized by TEM and SEM. Figure 4.8 shows the typical SEM images of CdS/P(S-DVB) nanocomposites. SEM showed small spherical particles with diameter 0.2 – 2 μm and some particles appeared to be in agglomerated form. These spherical particles should be ascribed to the aggregation of small CdS nanoparticles during the formation of the nanocomposites. The low tendency of aggregation in the CdS/P(SDVB) nanocomposite as shown in the SEM image suggested that stability of the miniemulsion system was enhanced by using monomer as oil-phase during the polymerization process. In particular, the smooth surface of the spheres which is indicative of polymer particles and the obvious absence of CdS crystals dispersed on the polymer surface, suggested that CdS nanoparticles were encapsulated in P(SDVB). The encapsulation of CdS in P(S-DVB) was also identified from the corresponding change in colour of the products, in which the CdS/P(S-DVB) nanocomposites showed a yellow colour, instead of the white appeareance of the P(S-DVB). The shape and size of the nanocomposite can be seen in the TEM micrograph of Figure 4.9, where the micrograph also shown the dark center and pale edge of the spheres. The dark surface observed in the TEM image indicates that it originated from CdS nanoparticles. Cadmium as heavy element developed a darker color compared to carbon and hydrogen in polymer structure. Of interest, TEM observations showed that cloud-like products were the main morphology, indicating that CdS nanoparticles were very small and homogeneously dispersed in P(S-DVB) 70 matrix. As can be seen, the spheres exhibit a diameter distribution in the range 200 – 2,000 nm and average diameter of ~500 nm. However, it is difficult to observe the particle size and size distribution of CdS nanoparticles by visual inspection of the TEM images, due to the TEM image was obtained by a low accelerating voltage (70 keV), hence the a statistical description of the size distribution of CdS nanoparticles cannot be carried out. (a) (b) Figure 4.8: SEM images of CdS/P(S-DVB) nanocomposites at (a) 2,500 X and (b) 10,000 X 71 (a) (b) Figure 4.9: TEM images of CdS/P(S-DVB) nanocomposites 4.1.2 Fourier Transform Infrared and UV – Vis Diffuse Reflectance Spectroscopy of CdS/P(S-DVB) Nanocomposites The pure P(S-DVB) and CdS/P(S-DVB) nanocomposites were characterized by FTIR spectroscopy. Figures 4.10(a) and 4.10(b) show both of the FTIR spectra were identical with respect to the main peaks of the polystyrene component. Pure polymer and the CdS/P(S-DVB) nanocomposites exhibit peaks assigned to the following groups: OH (at 3,400 – 3,200 cm-1), unsaturated aromatic C-H stretching vibrations (at 3,025 cm-1), CH2 bending vibration (at 2,920 cm-1 – 2,840 cm-1), aromatic ring (at 1,510 cm-1), CH2 (at 1,475 cm-1), and peaks associated with various substitutions of benzene ring between 900 cm-1 and 700 cm-1. Therefore, the chemical composition of the P(S-DVB) matrix was not changed by the presence of the CdS nanoparticles as filler, indicating no chemical interaction was involved but only weak physical interaction between two phases (Chen, M., et al. 2000). (a) 72 - OH (b) Transmittance (%) C-H2 bending C-H streching C-H aromatic - OH C-H2 bending C-H streching C-H aromatic 4000 3200 2400 1800 1400 1000 600 400 Wavenumbers (cm-1) Figure 4.10: FTIR spectra of (a) pure P(S-DVB) and (b) CdS/P(S-DVB) nanocomposites 73 The CdS/P(S-DVB) nanocomposites at different Wo values were characterized by UV – Vis DR spectroscopy and the spectra are shown in Figure 4.11. The absorption onset wavelength of CdS nanoparticles determined from the absorption UV – Vis DR spectra was significantly blue-shifted compared to bulk CdS. The absorption onset wavelength and particle sizes at Wo = 4, 5.55 and 7.5 calculated using Brus equation are 425 nm (3.1 nm), 462 nm (4.1 nm), 483 nm (5.0 nm), respectively. The blue shifting is indicative of quantum confinement effect due to decreasing particle size and dependence of the particle size of CdS/P(S-DVB) nanocomposites on Wo value. Quantum size effects were observed also through an obvious difference in colours of bulk (orange) and nanoparticle samples (yellow). Absorbance (a.u) (c) Wo = 7.5 (5.0 nm) (b) Wo = 5.55 (4.1 nm) (a) Wo = 4 (3.1 nm) P(S - DVB) 380 420 460 500 540 580 600 Wavelength (nm) Figure 4.11: UV-Vis DR spectra of pure P(S-DVB) and CdS/P(S-DVB) nanocomposites prepared by in-situ polymerization in miniemulsion system at different Wo value (a) Wo = 4, (b) Wo = 5.55, and (c) Wo = 7.5 74 A comparison of particles size of CdS/P(S-DVB) nanocomposite at Wo = 4, 5.55 and 7.5 showed a correlation between particle size. Thus, the particle size is expected to increase with increasing water content. This in turn allows us to obtain desired size of nanoparticles by controlling the amount of water in the miniemulsion system. The CdS particles with Wo = 7.5 exhibit larger size than those of particle with Wo = 4 and 5.55. 4.1.3 Dielectric and Electrical Properties of CdS/P(S-DVB) Nanocomposites Figure 4.12 shows the dielectric constant of pure P(S-DVB) and CdS/P(SDVB) nanocomposites over a frequency range of 100 Hz – 1 MHz. Dielectric properties of the nanocomposites are generally enhanced with increasing amount of CdS as compared to the pure P(S-DVB). Dielectric constant 250 200 a. b. c. d. e. 150 P(S-DVB) CdS/P(S-DVB) 0.01 % CdS/P(S-DVB) 0.02 % CdS/P(S-DVB) 0.03 % CdS/P(S-DVB) 0.18 % 100 50 e d c 0 b 2 3 a 4 5 Log frequency (Hz) 6 Figure 4.12: Dielectric constant for P(S-DVB) and CdS/P(S-DVB) nanocomposites with different amount of CdS as a function of various frequencies 75 Meanwhile the dissipation factors of the nanocomposites were higher than those of P(S-DVB) as shown in Figure 4.13. The result suggests the presence of charge carrier with strong mobility and polarizability of the electron attributed largerly to excess polarity of the CdS nanoparticles. According to Ghosh et al. (2005) and El-Tantawy et al. (2004), the possible reasons to explain the enhancement of dielectric properties, especially at low frequencies are, firstly, due to increasing interfacial interaction, including intermolecular and intramolecular, between CdS nanoparticles–CdS nanoparticles and also polymer particle–CdS nanoparticles. Secondly, increasing mobility of charge carrier and polarizability of electron to a larger excess with increasing CdS nanoparticle content. Finally, the increased value of the dielectric constant of the nanocomposites is due to the fact that the CdS nanoparticle and polymer are disconnected from each other, and under the application of voltage, these nanoparticles act as nano-dipoles. As CdS particle size is in the order of nanometers, the number of particles per unit volume increases, hence dipole moment per unit volume also increases, therefore dielectric constant also increases. 1.18 1.6 0.75 0.49 0.2 0.2 0.38 0.09 0.2 0.4 Pure P(S-DVB) 0.18 % CdS/P(S-DVB) 0.2 0.8 0.39 Dissipation factor 1.2 0 2 3 4 5 Log frequency (Hz) 6 Figure 4.13: Dissipation factor of (a) pure P(S-DVB) and (b) 0.18 % CdS/P(S-DVB) nanocomposites 76 The schematic illustration of the mechanism of interfacial interactions in CdS/P(S-DVB) nanocomposites is shown in Figure 4.14. Intermolecular interaction CdS nanoparticle – CdS nanoparticle Intramolecular interaction CdS nanoparticle – CdS nanoparticle Intramolecular interaction CdS nanoparticle – polymer Increase dielectric constant Figure 4.14: Schematic representation of the mechanism of interfacial interaction in CdS/P(S-DVB) nanocomposites The variation of dielectric constant of the CdS/P(S-DVB) nanocomposites with temperature is shown in Figure 4.15. The nanocomposites displayed a decreasing trend in dielectric constant with increasing temperature. This phenomenon is explained to correspond to the occurrence of two kinds of behaviors, which would yield converse effect on dielectric constant of the nanocomposites by changing the temperature. First, increasing temperature would improve the segmental mobility at the polymer matrix, facilitate the polarization of polar fillers and increase dielectric constant consequently. Second, the obvious difference in thermal expansion coefficient of polymer and CdS nanoparticles would disturb the aggregations of polar component, and thus reduce the dielectric constant. Since polystyrene is a of kind polymer with glass transition (Tg) temperature at 95 oC, the second behavior is the main effect (Wikes et al., 2005). It is believed that the disruption of the CdS nanoparticles cluster that have caused the dielectric constant to decrease with increase temperature. 77 120 Dielectric constant 100 100 Hz 80 1000 Hz 60 10 KHz 100 KHz 40 1 MHz 20 0 75 50 25 150 100 Temperature (oC) Figure 4.15: Influence of temperature on dielectric constant of 0.03 % CdS/P(SDVB) nanocomposites at various frequencies. 90 Percentage decrease (%) 80 80 75 70 60 50 38 40 30 20 37 20 10 0 2 3 4 5 6 Log frequency (Hz) Figure 4.16: Percentage decrease of dielectric constant of 0.03 % CdS/P(S-DVB) nanocomposites at various frequencies after heat treatment at 150o C 78 The result was further supported by the lower content of CdS and the lower dielectric constants of nanocomposites measured after heat treatment, where dielectric constant abruptly decreased. Figure 4.16 shows the percentage decrease of dielectric constant of 0.03 % CdS/P(S-DVB) nanocomposites measured after heat treatment relative to that before heat treatment (at room temperature) at various frequencies. Figure 4.17 shows the ionic conductivity of CdS/P(S-DVB) nanocomposite at different amount of CdS The ionic conductivity of the CdS/P(S-DVB) nanocomposite are increase when compared with pure P(S-DVB). This result suggested the CdS nanoparticles play a role in increasing the ionic conductivity due to the mobility of electron between CdS nanoparticles – CdS nanoparticles and also CdS nanoparticles – polymer particles. 3.00E-09 Ionic conductivity (S/ohm x 10-10) 2.50E-09 2.00E-09 1.50E-09 y = 1E-08x R2 = 0.9998 1.00E-09 5.00E-10 0.00E+00 0 0.05 0.1 0.15 0.2 CdS amount (%) Figure 4.17: Ionic conductivity of CdS/P(S-DVB) nanocomposites at different amount of CdS 79 4.1.4 Thermal Properties of CdS/P(S-DVB) Nanocomposites Figure 4.18 shows the thermal conductivity of CdS/P(S-DVB) nanocomposites, CdS nanoparticles and pure P(S-DVB) at room temperature. The thermal conductivity of the CdS/P(S-DVB) nanocomposites is slightly enhanced compared to pure polymer. Two possible reasons are suggested to explain the results. Firstly, the conductive path was enhanced as the consequence of increasing amount of CdS nanoparticles at the CdS/P(S-DVB) nanocomposites, because the CdS nanoparticles are capable to conduction transfer through the specimen sample (polymer matrix). Secondly, the interfacial thermal resistance among CdS nanoparticles between polymer matrix is reduced, leading to the increase in thermal conductivity. From the Figure 4.18 is clear that the addition of CdS improves thermal conductivity to an almost constant value Thermal conductivity (W m-1K-1) 0.2 0.15 0.1 0.05 0 1a 2b 3c d4 e5 Figure 4.18: Thermal conductivity of (a) pure P(S-DVB), (b) 0.01 % CdS/P(SDVB), (c) 0.02 % CdS/P(S-DVB), (d) 0.03 % CdS/P(S-DVB) and (e) 0.18 % CdS/P(S-DVB) nanocomposites at room temperature. 80 Figure 4.19 shows the TGA curve of the pure P(S-DVB) and 0.18 % CdS/P(S-DVB) nanocomposites carried out under an oxygen atmosphere in the temperature range of 40 – 1000 oC. The weight loss of around 310 - 400 oC in the TGA is contributed by the degradation of P(S-DVB). A slight shift in the TGA profile toward a higher temperature indicated a small but significant increase in thermal stability of the CdS/P(S-DVB) nanocomposites. The possible reason for the increased thermal stability of the nanocomposite is the reduced molecular mobility of the polymer chain due to the presence of CdS nanoparticles as filler. This observation was supported by the residue of the CdS/P(S-DVB) nanocomposites that still remained at temperature above 400 oC which was less than 1 %, whereas the pure polymer was completely decomposed at these temperatures. 100 (a) Pure P(S-DVB) Weight loss (%) 80 (b) 0.18 % CdS/P(S-DVB) nanocomposite 60 40 20 0 100 200 300 400 500 600 700 800 900 1000 Temperature (oC) Figure 4.19: Thermogravimetric analyses curves of (a) pure P(S-DVB), and (b) 0.18 % CdS/P(S-DVB) nanocomposites 81 4.2 Synthesis and Physicochemical Sulfide/Poly(methacrylic acid-ethylene Properties glycol of Cadmium dimethacrylic acid) [CdS/P(MAA-EGDMA)] Nanocomposites 4.2.1 Synthesis of P(MAA-EGDMA) and CdS/P(MAA-EGDMA) Nanocomposites The P(MAA-EGDMA) particles has been synthesized by in-situ polymerization in a miniemulsion system. The miniemulsion system was composed of CTABr as surfactant, 2-propanol as co-surfactant, MAA-EGDMA monomers and n-decane as oil-phase and water. Schematic representation of the synthesis of P(MAA-EGDMA) is illustrated in Figure 4.20. As previously discussed in section 4.1 the in-situ polymerization in a miniemulsion system has produced particle polymer. The result was supported by the SEM and TEM studies in which those images showed of the polymer with particles size in the range 100 – 300 nm. The CdS/P(MAA-EGDMA) nanocomposites were synthesized from P(MAAEGDMA) by ion exchange with various concentration of cadmium ion and precipitation with various concentration of sulfide ion. The mechanism of ion exchange and precipitation process as is proposed to include the replacing of hydrogen ion at carboxylic acid groups (-COOH) with cadmium ion and then followed by precipitation reaction with sulfide ion to produce CdS nanoparticles on the surface of polymer matrix. The schematic diagram for the synthetic scheme and reaction of polymerization are illustrated in Figure 4.21 and 4.22, respectively. The carboxylic acids as functional groups of the polymer were suggested to control particle size and in-situ formation of CdS nanoparticles on the surface of polymer matrix and to produce the yellow nanocomposites. These results were supported by UV – Vis DR spectra in which the spectra showed significantly blue shifted onset wavelength compared to bulk CdS. 82 Monomer Co-stabilizer n-decane Surfactant Water Stirring, Ultrasonication Surfactant M M M Monomer in miniemulsion Polymerization Polymer particles Figure 4.20: Schematic representation of synthesis of P(MAA-EGDMA) particles 83 O H3C H3C O + H2C O H2C OH CH2 O O In-situ polymerization CH3 O O OH O O O OH O OH OH O Cd2+ O O O OH O OCd2 O+ O OCd2+ O O OH O OH S2- O O O O CdS O O O O Cd OS OH O Figure 4.21: Schematic representation of preparation of CdS/P(MAA-EGDMA) nanocomposites 84 H3C O H3C O + H2C OH O H2C AIBN, 70oC CH2 O O in-situ polymerization CH3 in miniemulsion system Figure 4.22: Reaction of methacrylic acid and ethylene glycol dimethacrylic acid by free radical polymerization into P(MAA-EGDMA). 4.2.2 Fourier Transform Infrared and UV – Vis Diffuse Reflectance Spectroscopy of CdS/P(MAA-EGDMA) Nanocomposites It is known that the CdS semiconductor nanoparticles have unique sizedependent chemical and physical properties. As the size of CdS semiconductor particles decreases to the nanoscale, the band gap of the CdS semiconductor increases, causing a blue shift in the UV – Vis wavelength. The bulk CdS materials exhibit an onset wavelength in the range of 515 – 520 nm. If the onset wavelength of the obtained CdS sample appeared blue-shifted compared with that of bulk CdS, it could indicate the presence of CdS nanosize. 85 The P(MAA-EGDMA), CdS/P(MAA-EGDMA) nanocomposites with different CdS content, CdS nanoparticles and bulk CdS were characterized by UV– Vis DR spectroscopy (see Fig. 4.23). The spectrum of pure polymer did not exhibit absorption onset wavelength as it did not absorb in the UV – Vis range. The spectra of CdS/P(MAA-EGDMA) nanocomposites were significantly blue-shifted compared to CdS nanoparticle and bulk CdS, in agreement with quantum confinement effect due to decreasing particle size. Bulk CdS Absorbance (a.u ) f CdS nanoparticles e d c b a P(MAA - EGDMA) 380 420 460 500 540 580 Wavelength (nm) Figure 4.23: UV-Vis DR spectra of (a) P(MAA-EGDMA), (b) 0.81 % CdS/P(MAAEGDMA), (c) 0.90 % CdS/P(MAA-EGDMA), (d) 1.08 % CdS/P(MAA-EGDMA), (e) 1.23 % CdS/P(MAA-EGDMA), (f) 1.70 % CdS/P(MAA-EGDMA) and (g) 2.07 % CdS/P(MAA-EGDMA) Considering that the absorption of CdS particles in the composites increased with the CdS content, only a rough estimation of the particle size could be made for CdS particles in nanocomposites by UV-Vis DR spectrometry. Hence, absorption onset wavelength value was utilized to calculate the particle size using the Brus equation. According to the calculation, the sizes of CdS particles is about 7.1 – 8.0 86 nm (see Table 4.3). This result revealed that ion-exchange and precipitation processes with various concentration of cadmium and sulfide ion could control the growth, nucleation and formation of CdS nanoparticles on the surface of P(MAAEGDMA). Table 4.3: The absorption onset wavelength and particle size of CdS nanoparticles in CdS/P(MAA-EGDMA) nanocomposites with various CdS contents Nanocomposites sample Onset wavelength (nm) Particle size (nm) 0.81 % CdS/P(MAA-EGDMA) 504 7.1 0.90 % CdS/P(MAA-EGDMA) 507 7.3 1.08 % CdS/P(MAA-EGDMA) 508 7.3 1.23 % CdS/P(MAA-EGDMA) 512 7.6 1.70 % CdS/P(MAA-EGDMA) 515 7.8 2.07 % CdS/P(MAA-EGDMA) 520 8 The FTIR spectra of the series of P(MAA-EGDMA), CdS/P(MAA-EGDMA) nanocomposites and CdS/P(MAA-EGDMA) physical mixing are shown in Figure 4.24. These spectra show the main peaks due to the P(MAA-EGDMA) component that are at 3,250 cm-1 (C-H alkene), 2,920 cm-1 and 2,850 cm-1 (C-H alkane), 1,730 cm-1 (C=O carboxylic acid), 1,600 cm-1 and 1,580 cm-1 (C=C alkene), and 1,160 cm-1 (C-O carboxylic acid). Two new peaks appeared in the spectra of CdS/P(MAAEGDMA) nanocomposites at 1,550 cm-1 and 1,340 cm-1 which are the consequence of the ionization of carboxylic acid groups by cadmium ion. The ionization leads to the equilibration of the two oxygen atoms attached to carbon, resulting in the appearance of two peaks (in the 1,610 cm-1 – 1,550 cm-1 and 1,420 cm-1 – 1,300 cm-1), which correspond to the symmetrical and asymmetrical vibrations of carboxylic ion. Similar results were obtained in the case of CdS/P(S-Mac) nanocomposites (Nair et al., 2005). Meanwhile, different results were obtained for CdS/P(MAA-EGDMA) prepared by physical mixing as there were no appearance of both of these peaks. The FTIR results revealed that the chemical bonding between CdS nanoparticles with carboxylic group at polymer matrix occurred by ion exchange and precipitation process. 87 P(MAA-EGDMA) C=O carboxylic acid -C-O carboxylic acid -CH alkane Transmittent (a.u) -CH alkene New peak CdS/P(MAA-EGDMA) nanocomposites Physically mixed CdS/P(MAA-EGDMA) 4000 3200 2400 1800 1400 1000 600 400 Wavenumbers (cm-1) Figure 4.24: FTIR spectra of P(MAA-EGDMA), CdS/P(MAA-EGDMA) nanocomposites and physically mixed CdS/P(MAA-EGDMA) 88 4.2.3 Structural and Morphological Properties of CdS/P(MAA-EGDMA) Nanocomposites The morphologies of the CdS/P(MAA-EGDMA) nanocomposites prepared by ion-exchange and precipitation were examined by SEM and TEM (see Figure 4.25 and Figure 4.26). According to SEM micrograph, non-agglomerated (fused) and some agglomerated particles were disperses throughout the nanocomposite. But the fraction of non-agglomerate particles was smaller. From SEM images it is clear that the agglomerated CdS nanoparticles were attached on the surface of P(MAAEGDMA) with sizes below 20 nm. The Energy Dispersive X-ray (EDX) analyses were utilized to confirm the existence of cadmium and sulfur in the CdS/P(MAAEGDMA) nanocomposites. However, the smaller the CdS nanoparticles are the larger is their internal surface and hence their tendency to agglomerate rather than to disperse homogenously in the polymer matrix. The resulting materials may be seen also as filled polymers since there is no or little interaction between two mixed components. The inorganic, relatively small cadmium ion is charged against more polymer (organic) molecules. The exchange reaction has two consequences: First, the polymer swells and the gaps between polymer chains are widened, enabling CdS nanoparticles to more in between them and secondly, the surface property of the polymer is changed from hydrophobic to hydrophilic. The TEM images also show little agglomeration tendency of the CdS/P(MAA-EGDMA) nanocomposites. The dark shadow around the surface of the polymer at TEM images again suggested it came from CdS nanoparticles. This results was supported by the filled surface of P(MAA-EGDMA) that has been filled with CdS nanoparticles. However, it is difficult to observe the particle size and size distribution of CdS nanoparticles by visual inspection of the TEM images, due to the TEM image was obtained by a low accelerating voltage (70 keV), hence the particle size and distribution of CdS nanoparticle cannot be determined. The SEM and TEM images also showed the particles sizes of P(MAA-EGDMA) as host matrix in the nanosize range of 100 - 300 nm. 89 (a) (b) Figure 4.25: SEM images of 2.07 % CdS/P(MAA-EGDMA) nanocomposites at magnification (a) 50,000 X and (b) 100,000 X 90 (a) 200 nm (b) Figure 4.26: TEM images of (a) agglomerated and (b) fused samples 2.07 % CdS/P(MAA-EGDMA) nanocomposites The structure of P(MAA-EGDMA) and CdS/P(MAA-EGDMA) nanocomposites were investigated by XRD. Figure 4.27 show the XRD pattern of P(MAA-EGDMA) and CdS/P(MAA-EGDMA) nanocomposites with different amounts of CdS nanoparticles. 91 2.07 % CdS/P(MAA-EGDMA) Relative intensity (a.u) 1.70 % CdS/P(MAA-EGDMA) 1.23 % CdS/P(MAA-EGDMA) 1.08 % CdS/P(MAA-EGDMA) 0.90 % CdS/P(MAA-EGDMA) 0.81 % CdS/P(MAA-EGDMA) Pure P(MAA-EGDMA) 20 30 40 50 60 70 2-Theta degree Figure 4.27: XRD patterns of pure P(MAA-EGDMA) and CdS/P(MAA-EGDMA) nanocomposites at various CdS amount 92 The XRD pattern of the nanocomposites exhibit broad peaks corresponding to 111, 220, and 311 (plane) indicating that cubic CdS nanoparticles was mixtured in polymer matrix. Broad XRD peaks also are attributed to the absence of long-range order in the samples and reflect the small particles size (less than 10 nm as estimated from the Brus equation). The polymer particles which are amorphous and present in large amount also contribute to the peak broadness This result agrees with the presence of nanocrystalline CdS, for which quantum size effect are expected. The weak reflection peaks should belong to CdS by comparison with the XRD pattern of CdS nanoparticles. The low intensity peaks indicated low contents of the CdS nanoparticle in polymer matrix. 4.2.4 Dielectric and Electrical Properties of CdS/P(MAA-EGDMA) Nanocomposites Figure 4.28 and Table 4.4 shows of the data on dielectric constant of P(MAAEGDMA), CdS nanoparticles, physically mixed CdS/P(MAA-EGDMA) and CdS/P(MAA-EGDMA) nanocomposites at various frequencies. The frequency dependency of the data on dielectric constant of CdS/P(MAA-EGDMA) nanocomposites is presented in Figure 4.28. In all cases of the CdS/P(MAAEGDMA) nanocomposites, the dielectric constant showed a very steep decrease from its initial high values. The decrease in dielectric constant is very obvious for the 1.08 % CdS/P(MAA-EGDMA) sample. At high frequencies above 105 Hz, the dielectric constants remained nearly constant. Interestingly, at 100 Hz, the dielectric constant for the 1.08 % CdS/P(MAA-EGDMA) sample was about 5,200. This value decreased to about 320 for the composite samples having 2.07 wt % of CdS. The result showed a higher dielectric constant as compared to the dielectric constant of CdS/PVA nanocomposites with average dispersibility in the range of 0.3 – 0.8 (dielectric constant about 120 – 250 over the range of frequency 10 – 40 kHz) (Zhou, 2003), CdS/PVA 20 % (dielectric constant about 10 at frequency 1 kHz) (Ghosh et al., 2005) and TiO2/sulfonated styrene-b-(ethylene-ran-butylene)-b-styrene (S-SEBS) block copolymer 6.4 % (dielectric constant about 3 - 3.2 over the frequency range of 1 – 10,000 Hz) (Yang and Kofinas, 2007). 93 6000 5000 Dielectric constant CdS/P(MAA-EGDMA) 0.81 % 4000 CdS/P(MAA-EGDMA) 0.90 % 3000 CdS/P(MAA-EGDMA) 1.08 % CdS/P(MAA-EGDMA) 1.23 % 2000 CdS/P(MAA-EGDMA) 1.70 % 1000 CdS/P(MAA-EGDMA) 2.07 % 0 2 3 4 5 Log frequency (Hz) 6 Figure 4.28: Dielectric constant of CdS/P(MAA-EGDMA) nanocomposites with various amounts of CdS measured at different frequencies Table 4.4: Dielectric constants of P(MAA-EGDMA), CdS nanoparticles, CdS/P(MAA-EGDMA) prepared by physical mixing and CdS/P(MAA-EGDMA) nanocomposites at various frequencies Log frequency (Hertz) 6 5 4 3 2 P(MAA-EGDMA) 1.6 3.8 4.0 4.9 9.2 CdS nanoparticles 4.6 16.2 28.9 49.8 100 1 % CdS/P(MAA-EGDMA) phys. mix. 0.8 5.2 9.5 22.0 60.6 11.9 % Na/P(MAA-EGDMA) 2.3 6.8 7.0 13.0 42.0 1.4 % Cd/P(MAA-EGDMA) 2.1 5.8 5.9 8.0 15.3 0.81 % CdS/P(MAA-EGDMA) 1.7 5.0 7.7 16.6 73 0.90 % CdS/P(MAA-EGDMA) 3.3 8.7 15.6 59.4 213.8 1.08 % CdS/P(MAA-EGDMA) 22.4 159.9 803.7 2,321 5,245 1.23 % CdS/P(MAA-EGDMA) 3.9 10.7 29.4 143.5 732.2 1.70 % CdS/P(MAA-EGDMA) 1.5 7.6 23.9 150.7 455.1 2.07 % CdS/P(MAA-EGDMA) 3.3 10.6 36.0 157.5 320.6 94 Figure 4.29. shows the histogram of the dielectric constants between CdS/P(MAA-EGDMA) nanocomposites with CdS nanoparticles, pure P(MAAEGDMA) and 1.00 % CdS/P(MAA-EGDMA) (physical mixing) measured at 100 Hz. It was found that the maximum value of 5245.6 was achieved for at the 1.08 % concentration of CdS, which is far higher (ca. 500 times) than the value of obtained for pure P(MAA-EGDMA). On the other hand, the value of the dielectric constant of pure P(MAA-EGDMA), CdS nanoparticle, 1 % CdS/P(MAA-EGDMA), 11.9 % Na/P(MAA-EGDMA) and 1.4 % Cd/P(MAA-EGDMA) at this frequency are about 10, 100, 60, 42 and 15, respectively. The enormous enhancement of properties at the nanocomposite containing 1.08 % CdS may be related to the effect of adding CdS nanoparticles. 5245.6 6000 4000 g i j 320.6 f 455.1 213.8 d 73 42 c 15.3 60.6 1000 100 2000 732.2 3000 9.2 Dielectric constant 5000 0 a b e h k Figure 4.29: Dielectric constant of (a) P(MAA-EGDMA), (b) CdS nanoparticles, (c) CdS/P(MAA-EGDMA) physical mixing, (d) 11.9 % Na-P(MAA-EGDMA), (e) 1.4 % Cd-P(MAA-EGDMA), (f) 0.81 % CdS/P(MAA-EGDMA), (g) 0.90 % CdS/P(MAA-EGDMA), (h) 1.08 % CdS/P(MAA-EGDMA), (i) CdS/P(MAA-EGDMA), (j) 1.70 % CdS/P(MAA-EGDMA) and (k) 1.23 % 2.07 % CdS/P(MAA-EGDMA) at 100 Hz. To better understand the above phenomenon, FTIR measurements were carried out on CdS nanoparticles, pure P(MAA-EGDMA), CdS/P(MAA-EGDMA) 1.00 % (physical mixing) and CdS/P(MAA-EGDMA) with various concentrations of 95 CdS. The broad feature between 3,500 cm-1 and 2,500 cm-1 was undoubtedly due to the O-H stretch of the carboxylic acid. Two bands at 2,920 cm-1 and 2,850 cm-1, which were superimposed on the O-H stretch, were attributed to the asymmetric CH2 stretch and the symmetric CH2 stretch, respectively (Smith, 1999). The intense peak at 1,710 cm-1 was derived from the existence of the C=O stretch and the band at 1,282 cm-1 exhibited the presence of the C-O stretch (Smith, 1999). As shown in Fig. 4.30, it is worth noting that the C=O stretch band of the carboxyl group, which is present at 1,710 cm-1 in the FTIR spectrum of pure P(MAA-EGDMA), is lowered in the spectra of the CdS/P(MAA-EGDMA) nanocomposites. Instead, a new band at 1,350 cm-1 appeared. This peak is characteristic of the symmetric (COO-) stretch. This reveals that CdS nanoparticels are chemisorbed onto carboxylate functional groups of P(MAA-EGDMA). To be more comprehensive, Figure 4.31 shows the intensity of the IR peak at 1,350 cm-1 increasing with decreasing intensity of IR peak at 1,710 cm-1, which can be associated to the degree of the interfacial interaction between CdS and polymer. There was no detectable interfacial interaction in 1.00 % CdS/P(MAA-EGDMA) (physical mixing) and CdS/P(MAA-EGDMA) when the concentration of CdS was lower than 1.08 wt % since the peak at 1,350 cm-1 was not observed in the FTIR spectra. The origin of the high dielectric constant of 1.08 % CdS/P(MAA-EGDMA) is still uncertain and not fully understood yet. Because the nanocomposite consists of CdS attached on the surface of polymer, it likely that the dielectric constant may have increased due to interfacial and space charge polarization over a frequency range of 100 Hz to 106 Hz, which is associated with an inhomogeneous dielectric medium containing two layers of materials with different permittivity and conductivity (Kremer and Schonhals, 2002). A very steep decrease in dielectric constant with increasing frequency in the 1.08 % CdS/PMAA-EGDMA sample suggested the occurrence of a strong space charge polarization effect in this nanocomposite. As described above, 1.08 % CdS/P(MAA-EGDMA), having the highest dielectric constant of 5,200, showed a high intensity of IR peak at 1350 cm-1 which is attributed to the interfacial interaction between CdS nanoparticles and P(MAA-EGDMA) (see Figure 4.31). 96 C dS nanopa rticles P (MAA- EGDM A) CdS /P( MAA-EGD MA) (physical mixing) 1.00 % Transmittance (%) CdS/P( MA A-EGD MA) 0.81 % C dS/P(M AA-EGDMA) 0.9 % CdS/P( MA A-EGD MA) 1.08 % CdS /P( MAA-EGD M A) 1.23 % CdS /P( MA A-EGD M A) 1.07 % CdS /P( MA A-EGD M A) 2.07 % 2850 2920 1350 1710 4000 3600 3200 2800 2400 20 00 1800 1600 1160 1282 1400 Wa venumber / cm 1200 10 00 800 600 -1 Figure 4.30: FTIR spectra of P(MAA-EGDMA), CdS/P(MAA-EGDMA) physical mixing, CdS nanoparticle and CdS/P(MAA-EGDMA) nanocomposites with various content of CdS 97 Intensity of IR peak at 1710 cm -1 / a.u. Intensity of IR peak at 1350 cm-1 / a.u. P(MAA-ED MA) P(MAA-EGDMA) CO O – CdS CO O– 11.= P(MAA-EGDMA) P(MA A- EDMA) P(MAA-EGDMA) 22.= CdS/ Cd S-P( MAA-ED MA)1.00 1.0% 0% Phys mix l mix ing) (ph ysica P(MAA-EGDMA) 33.= CdS/ Cd S-P( MAA-ED MA)0.81 0.8% 1% 4. CdS/ P(MAA-EGDMA) 0.90 4 = Cd S-P( MAA-ED MA) 0.9% 0% P(MAA-EGDMA) 55.= CdS/ Cd S-P( MAA-ED MA)1.08 1.0% 8% P(MAA-EGDMA) 66.= CdS/ Cd S-P( MAA-ED MA)1.23 1.2% 3% P(MAA-EGDMA) 77.= CdS/ Cd S-P( MAA-ED MA)1.70 1.7% 0% P(MAA-EGDMA) 88.= CdS/ Cd S-P( MAA-ED MA)2.07 2.0% 7% 1 P(MAA-EGDMA) P(MAA-EDM A) 2 3 4 interphase 5 6 7 8 CdS Figure 4.31: Schematic representation of interfacial interaction in the CdS/P(MAAEGDMA) nanocomposites Even though CdS/P(MAA-EGDMA) nanocomposites with CdS nanoparticles content higher than 1.08 % showed a similar degree of the interfacial interaction, their dielectric constant is very much lower than those of 1.08 % CdS/P(MAAEGDMA). The presence of multilayer CdS nanoparticles on the surface of P(MAAEGDMA) at a high concentration could be the reason of the low dielectric constant of the CdS/P(MAA-EGDMA) sample containing CdS nanoparticles higher than 1.08 98 %. On the basis of this consideration, we model a dielectric nanocomposite as a three-phase material, consisting of a polymer, an interfacial phase, and CdS nanoparticles, schematically shown in Figure. 4.31. The interfacial phase is between the polymer and CdS nanoparticles. It is reasonable to assume that the interfacial area has fixed surface area, since the IR peak at 1,350 cm-1, attributed to the interfacial interaction between polymer and CdS nanoparticle was unchanged. The interfacial areas in the nanocomposite could promote interfacial exchange coupling through a dipolar interface layer leading to enhanced polarization and polarizability in polymer phase near the interface (Zhang et al. 2002, Li, 2003). As a result, enhanced permittivity can be expected in the polymer phase near the interfaces. The schematic model was also supported by the SEM images of 0.81 %, 1.08 % and 2.07 % CdS/P(MAA-EGDMA) nanocomposites of CdS with magnification of 100,000 and 150,000 X (see Figure 4.32). The SEM images of 0.81 % and 1.08 % CdS/P(MAA-EGDMA) nanocomposites show smaller particle size as compared to that at 2.07 %. There are two kinds of interfacial interactions in the proposed model shown in Fig. 4.31: (i) intermolecular interaction between CdS nanoparticles – P(MAA-EGDMA) and (ii) intermolecular interaction between CdS nanoparticles – CdS nanoparticles. Only the first layer of CdS nanoparticles contributed to the interfacial interaction of CdS – polymer and hence to the interfacial space charge polarization or polarization Maxwell-Wagner. The high electrical conductivity of CdS/P(MAAEGDMA) at the same amount of CdS supported the occurance of polarization Maxwell-Wagner. The next layer of CdS nanoparticles lowered the space charge polarization caused by interaction between CdS nanoparticles – CdS nanoparticles. One explains that an external electric field, inducing an electric current along the material, produces mobility of charge carrier across the interface between CdS nanoparticles. Therefore, the decrease in the dielectric constant of CdS/P(MAAEGDMA) nanocomposites with CdS content higher than 1.08 % is observeb. 99 (a) (b) (c) Figure 4.32: SEM images of CdS/P(MAA-EGDMA) at various amount of CdS (a) 0.81 % , (b) 1.08 % and (c) 2.07 % The dissipation factors of CdS/P(MAA-EGDMA) nanocomposites at various contents of CdS nanoparticles and frequencies over the range of 0.10 – 1,000 KHz are shown in Figure 4.33 and Table 4.5. 100 Figure 4.33: Dissipation factor of P(MAA-EGDMA) and CdS/P(MAA-EGDMA) nanocomposites at various CdS contents and frequencies (a) P(MAA-EGDMA), (b) 0.81 % CdS/P(MAA-EGDMA), (c) 0.90 % CdS/P(MAA-EGDMA), (d) 1.08 % CdS/P(MAA-EGDMA), (e) 1.23 % CdS/P(MAA-EGDMA), (f) 1.70 % CdS/P(MAA-EGDMA), and (g) 2.07 % CdS/P(MAA-EGDMA) Table 4.5: Dissipation factors of P(MAA-EGDMA), CdS nanoparticles and CdS/P(MAA-EGDMA) nanocomposites Sample Log frequency ( Hz) 6 5 4 3 2 P(MAA-EGDMA) 0.01 0.02 0.01 0.04 0.24 CdS nanoparticles 0.20 0.20 0.20 0.50 1.60 0.81 % CdS/P(MAA-EGDMA) 0.02 0.03 0.10 0.30 1.30 0.90 % CdS/P(MAA-EGDMA) 0.06 0.04 0.27 0.53 0.48 1.08 % CdS/P(MAA-EGDMA) 0.87 2.48 1.11 1.35 0.53 1.23 % CdS/P(MAA-EGDMA) 0.02 0.13 0.67 0.35 0.10 1.70 % CdS/P(MAA-EGDMA) 0.34 0.25 0.83 0.35 0.30 2.07 % CdS/P(MAA-EGDMA) 0.86 0.85 0.31 0.29 0.31 101 Most of the dissipation factors was higher in values in comparison to those of P(MAA-EGDMA). The result suggests the appearance of charge carrier with strong mobility and polarizability of electron to a larger excess polarity for the CdS nanoparticles. Figure 4.34 shows the ionic conductivity of P(MAA-EGDMA), CdS nanoparticles and CdS/P(MAA-EGDMA) nanocomposites. The ionic conductivity values of the nanocomposites at various CdS content seem to be following the trend for dielectric properties of CdS/P(MAA-EGDMA) nanocomposites. The highest ionic was found in 1.08 % CdS/P(MAA-EGDMA). The enormous enhancement in ionic conductivity again might be cause by the same reasons specified for the dielectric properties. 2.9E-06 3.0E-06 2.0E-06 6.4E-08 2.1E-07 2.3E-07 7.3E-08 b 4.1E-08 a 5.0E-09 5.0E-07 4.0E-08 1.0E-06 6.2E-08 1.5E-06 2.9E-09 Ionic conductivity 2.5E-06 0.0E+00 c d e f g h i j Sample Figure 4.34: Ionic conductivity of (a) P(MAA-EGDMA), (b) CdS nanoparticles, (c) 11.9 % Na-P(MAA-EGDMA), (d) 1.4 % Cd-P(MAA-EGDMA), (e) 0.81 % CdS/P(MAA-EGDMA), (f) 0.90 % CdS/P(MAA-EGDMA), CdS/P(MAA-EGDMA), (h) 1.23 % CdS/P(MAA-EGDMA), CdS/P(MAA-EGDMA) and (j) 2.07 % CdS/P(MAA-EGDMA) (g) 1.08 % (i) 1.70 % 102 4.2.5 Influence of Temperature on Dielectric and Electrical Properties of CdS/P(MAA-EGDMA) Nanocomposites The temperature dependence of dielectric constant in nanocomposites is important for any potential use in electronic devices, because one expects that there is the dependence of the orientation polarizability of the nanocomposites on temperature. Figure 4.35 shows the dielectric constant of the CdS/P(MAA-EGDMA) nanocomposites at different temperatures. The dielectric properties of the nanocomposites display an upward trend over the frequency range of 0.1 – 1,000 KHz as the temperature increased from 25 to 75 oC, but abruptly decreased after 75 o C. This phenomenon is explained by the occurrence of two kinds of behaviors, which would yield converse effect on dielectric constant of the nanocomposites by changing of temperature. First, increasing temperature would improve the segmental mobility of polymer and CdS nanoparticles, facilitate the polarization of polar CdS fillers and increase dielectric constant consequently. Second, the obvious differential in thermal expansion coefficient between polymer and CdS nanoparticles would disturb the aggregations of polar components, and thus reduced the dielectric constant. At temperature range of 25 – 75 oC for the CdS/P(MAA-EGDMA) nanocomposites, the first behavior seems to be the main effect. On the other hand, at temperatures above 75 oC, the second behavior is more dominant. It should be the disruption of the CdS nanoparticles cluster that caused the decreasing dielectric constant with increasing temperature. The results were also supported by the dielectric constant of nanocomposites measured after heat treatment, showing that the dielectric constants were slightly decreased. Figure 4.36 shows the percentage decrease of dielectric constant of 1.70 % CdS/P(MAA-EGDMA) nanocomposites measured after heat preparation at 150 oC in comparison to those measured at room temperature before treatment at various frequencies. 103 2500 100 Hz 2000 . Dielectric constant 1500 1 kHz 10 kHz 1000 100 kHz 500 0 1000 kHz 0 50 100 o Temperature C 150 Figure 4.35: Influence of temperature on dielectric constant of 1.70 % CdS/P(MAAEGDMA) nanocomposites at different frequencies 14 12 12 Percentage decrease (%) . 10 8 6 4 2 1 0.5 0.5 0.5 0 2 3 4 5 Log frequency (Hz) 6 Figure 4.36: Percentage decrease of dielectric constant of 1.70 % CdS/P(MAAEGDMA) nanocomposites after heat treatment and at various frequencies 104 The ionic conductivity of the CdS/P(MAA-EGDMA) nanocomposites at different temperatures is shown in Figure 4.37. The nanocomposites display a decreasing trend in dielectric constant with increasing temperature. The reason for this is the destruction of percolation of CdS nanoparticles as result of the thermal expansion of polymer at higher temperatures Electrical conductivity (S/ohm) 1.20E-05 1.00E-05 8.00E-06 6.00E-06 4.00E-06 2.00E-06 0.00E+00 30 40 50 60 70 80 90 100 Temperature (oC) Figure 4.37: Influence of temperature on ionic conductivity of CdS/P(MAAEGDMA) nanocomposites 4.2.6 Thermal Properties of CdS/P(MAA-EGDMA) Nanocomposites Figure 4.38 shows the thermal conductivities for P(MAA-EGDMA), CdS nanoparticle and CdS/P(MAA-EGDMA) nanocomposites. The thermal conductivity of nanocomposites was enhanced with increasing content of CdS nanoparticle compared to the pure polymer employed. Two possible reasons are put forward to explain the results. Firstly, the conductive path at the nanocomposites was enhanced partly as a consequence of the incorporated CdS nanoparticle. This is because the 105 CdS nanoparticles are more effective in transfering heat through the specimen. The aggregates are more effective in forming heat conduction in the polymer matrix as host of the nanocomposites to transfer heat through the specimen and thereby thermal conductivity enhances with increasing filler content. Secondly, the interfacial thermal resistance among CdS nanoparticles/polymer is reduced, leading to the increase in thermal conductivity. The result clearly concluded that the thermal conductivity of nanocomposite was mostly dependent on the content and properties of the CdS nanoparticles. Similar results and reasons were reported in the case of CdS/PVA nanocomposites (Ghosh et al., 2005). 0.22 0.25 Thermal conductivity (W/mK) 0.2 0.15 0.0907 0.105 0.1493 0.1539 e f 0.1063 0.1 0.05 0 a b c d Figure 4.38: Thermal conductivity of (a) P(MAA-EGDMA), (b) CdS nanoparticle, (c) 0.81 % CdS/P(MAA-EGDMA), (d) 0.90 % CdS/P(MAA-EGDMA) , (e) 1.08 % CdS/P(MAA-EGDMA), and (f) 2.07 % CdS/P(MAA-EGDMA) nanocomposites 106 Figure 4.39 shows the thermogravimetric analysis (TGA) curves of P(MAAEGDMA) and CdS/P(MAA-EGDMA) nanocomposites with different CdS nanoparticles content. The TGA curves indicated that the losses of weight occurred around three temperature periods, ranging from 50 – 100 oC, 250 – 400 oC and 450 – 780 oC. The first stage of mass loss was mainly contributed by the elimination of impurities and water. This stage of mass loss of the CdS/P(MAA-EGDMA) nanocomposites showed higher mass loss as compared to pure polymer, indicating that the surfaces of the nanocomposites are more hydrophilic. The second and third stages of mass losses were contributed by the thermal degradation of the backbone structure of polymer. The thermal stability of the nanocomposites improved significantly at higher decomposition temperatures in the range of 700 – 750 oC. The result indicated that the addition of CdS nanoparticles to polymer made the polymer more stable thermally. 100 a Weight loss (%) 80 b 60 40 c d 20 0 100 200 300 400 500 600 700 800 900 1000 Temperature (°C) Figure 4.39: Thermogravimetric analysis curve of (a) P(MAA-EGDMA), (b) 0.90 % CdS/P(MAA-EGDMA), (c) 1.23 % CdS/P(MAA-EGDMA) and (d) 2.07 % CdS/P(MAA-EGDMA) nanocomposites 107 The latter was related to the combination of CdS nanoparticles with the polymer matrix, which as result from the interaction between CdS nanoparticles and the carboxylic acid groups of the polymer. It can also be seen that the remaining residue obtained after the thermal decomposition at 750 oC is about 1 – 2 % whereas the P(MAA-EGDMA) was completely decomposed at these temperatures and left no residue. Similar thermal behaviour was obtained in the cases of CdS/PVA (ElTantawy et al., 2005), CdS/PS-co-Mac (Nair et al., 2005), CdS/PS (Kuljanin et al., 2002) and CdS/PMMA (Kuljanin et al., 2006). 4.3 Synthesis and Physicochemical Properties of CdS/Sulfonated- Poly(styrene-divinylbenzene) 4.3.1 Synthesis of P(S-DVB), SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) Nanocomposites Schematic representation of the reaction of polymerization is illustrated in Figure 4.40. The in-situ polymerization in miniemulsion was produced polymer particles with a dimension in the micrometer range. CH o CH2 CH CH2 AIBN, 70 C + Polymerization Styrene Divinyl benzene CH2 CH Sulfonation CH CH2 CH n CH2 CH2 CH SO3H n Figure 4.40: Reaction of styrene and divinyl benzene by free radical polymerization and sulfonation process to form SO3H/P(S-DVB) 108 The SO3H-P(S-DVB) was synthesized by sulfonation reaction of P(S-DVB) with fuming sulfuric acid (H2SO4/SO3) at 50 oC for 1 hour. The degree of sulfonation of SO3H-P(S-DVB) which was determined by titration method was 39.10 %. Meanwhile, the CdS/SO3-P(S-DVB) nanocomposites were synthesized from SO3HP(S-DVB) by ion exchange and precipitation process with various concentrations of cadmium ion, followed by reaction with sulfide ion. The mechanism of ion exchange and precipitation process occurred with replacement of hydrogen ion at sulfonic acid groups by cadmium ion with various concentrations which was then followed by precipitation by sulfide ion at various concentrations and producing CdS nanoparticles on the surface of polymer matrix. The ion exchange, precipitation process and functional groups of the polymer helped to control particle size and formation of CdS nanoparticles on the surface of polymer matrix and produced the yellow nanocomposites. Schematic representation of the synthesis of CdS/SO3-P(SDVB) nanocomposites is illustrated in Figure 4.41. 4.3.2 Fourier-Transform Infrared and UV – Vis Diffuse Reflectance Spectroscopy of CdS/SO3-P(S-DVB) Nanocomposites The CdS/SO3-P(S-DVB) nanocomposites with various amounts of CdS were characterized by UV – Vis DR spectroscopy (see Figure 4.42). The absorption onset wavelength of CdS nanoparticles in the nanocomposites was determined using the obtained UV – Vis DR spectra and it was found to be significantly blue-shifted compared to bulk CdS. The direction of the blue shifting is in agreement with quantum confinement effect due to decreasing particle size. The results also suggested that the ion exchange-precipitation processes could control the growth, nucleation and formation of CdS nanoparticles on the polymer surface. Based on the absorption onset wavelength value, the particle size of CdS was calculated using the Brus equation (see Table 4.6). 109 SO3H Sulfonation H2SO4/SO3 50o C, 1 h SO3H SO3H SO3H SO3H SO3H HSO3 SO3H HSO3 Ion exchange with various mmol of Cd 2+ SO3H SO3H SO3 CdS SO3 Precipitation with various mmol Na2S SO3 Cd2+ SO3 SO3H SO3H SO3Na SO3H HSO3 SO3Na HSO Residue Residue HSO3 SO3Na CdS SO3 Cd 2+ HSO SO3 Polymer particle Figure 4.41: Schematic representation of the synthesis CdS/SO3-P(S-DVB) nanocomposites 110 Relative intensity (a.u) h g c f b d e a 380 420 460 500 540 580 600 Wavelength (nm) Figure 4.42: UV-Vis DR spectra of (a) Pure P(S-DVB), (b) SO3H-P(S-DVB), (c) 2.56 % CdS/SO3-P(S-DVB), (d) 3.34 % CdS/SO3-P(S-DVB), (e) 7.64 % CdS/SO3P(S-DVB), (f) 14.65 % CdS/SO3-P(S-DVB). (g) 16.56 % CdS/SO3-P(S-DVB) and (h) CdS nanoparticles Table 4.6: The absorption onset wavelength and particle size of CdS/SO3-P(S-DVB) nanocomposites Nanocomposite sample Onset wavelength (nm) Particle size (nm) *) 2.56 % CdS/SO3-P(S-DVB) Not observed Nd 3.34 % CdS/SO3-P(S-DVB) Not observed Nd 7.64 % CdS/SO3-P(S-DVB) 500 7.0 14.65 % CdS/SO3-P(S-DVB) 507 7.3 16.56 % CdS/SO3-P(S-DVB) 550 >8 Nd : not determined, *) determined using Brus equation 111 The FTIR spectra of the series of P(S-DVB), SO3H-P(S-DVB) and CdS/SO3P(S-DVB) nanocomposites are shown in Figure 4.43. These spectra show the main peaks due to the polystyrene component. The P(S-DVB), SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites show peaks corresponding to the following groups: - OH (at 3,400 cm-1 – 3,200 cm-1), unsaturated aromatic C-H stretching vibrations (at 3,025 cm-1), CH2 bending vibration (at 2,920 cm-1 – 2,840 cm-1), aromatic ring (at 1,510 cm-1), CH2 (at 1,475 cm-1), various substitution of benzene ring between 900 cm-1 and 700 cm-1. Martins et al. (2003) have studied the characteristic infrared spectra of SO3H-P(S-DVB) and suggested that the bonding of the sulfonic groups to the aromatic ring of polystyrene (out of plane deformation bands assigned to substituted aromatic ring δ (Car-H) occurs at wavenumbers of approximately 830 cm-1 to 850 cm-1). Several new peaks were found in the spectra of SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) at 560 cm-1, 615 cm-1, 1,000 cm-1, 1,030 cm-1, 1,110 cm-1, 1,180 cm-1 and 1,230 cm-1. The absorption peak at 1,030 cm-1 was attributed to the symmetric stretching vibration of sulfonic acid groups (-SO3H) and the band at 1110 cm-1 was due to a sulfonate anion attached to phenyl ring. The υassym (S-O) vibration at 1,180 cm-1 appears as broad band at approximately 1,200 cm-1 to 1,300 cm-1. An increasing peak at 1,110 cm-1 and a decreasing peak at 1,230 cm-1 were observed in the spectrum of CdS/SO3-P(S-DVB) are the consequence of the ionization of sulfonic acid groups by cadmium ion. The phenomena clearly showed that the chemical bonding between CdS nanoparticles with sulfonated polymer matrix was quite strong. 112 Transsmitent (a.u) CdS/SO3H-P(S-DVB) (SO3)2-Cd/P(S-DVB) SO3Na-P(S-DVB) SO3H-P(S-DVB) CdS nanoparticles 4000 3600 3200 2800 2400 2000 1800 1600 1400 1200 1000 800 600 450 Wavelength (cm-1) Figure 4.43: FTIR spectra of P(S-DVB), SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites 113 4.3.3 Structural and Morphological Properties of CdS/SO3-P(S-DVB) Nanocomposites The structure and morphology of SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites were investigated by XRD, SEM and TEM techniques. Figures 4.44(a) – 4.44(b) and 4.45 show the XRD patterns of CdS nanoparticles, bulk CdS, SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites with various CdS content respectively. Intensity (a.u) (a) 20 30 40 50 60 70 2-Theta Intensity (a.u) (b) 20 30 40 50 60 70 80 2-Theta Figure 4.44: XRD patterns of (a) CdS nanoparticles and (b) bulk CdS 90 114 26.40 43.75 52.00 Relative intensity (a.u) 16.56 % CdS/SO3-P(S-DVB) 14.65 % CdS/SO3-P(S-DVB) 7.64 % CdS/SO3-P(S-DVB) 3.34 % CdS/SO3-P(S-DVB) 2.56 % CdS/SO3-P(S-DVB) SO3H-P(S-DVB) 20 30 40 50 60 70 2-Theta degree Figure 4.45: XRD patterns of SO3H-P(S-DVB) and the various CdS/SO3-P(S-DVB) nanocomposites 115 The XRD patterns of the nanocomposites are identical with the XRD patterns of the CdS nanoparticle and bulk CdS. The XRD patterns of CdS/SO3-P(S-DVB) nanocomposites exhibit broad peaks corresponding to 111, 220, and 311 (plane) indicating that CdS nanoparticles are in the cubic phase. Broad XRD peaks are attributed to the absence of long-range order in the samples and reflect the small particle sizes (less than 10 nm as estimated from the Brus equation). This result agrees with the presence of nanocrystalline CdS, for which quantum size effect are expected. The weak reflection peaks should belong to CdS by comparison with the XRD pattern of CdS nanoparticles. The low intensity peaks indicated that the dimension of CdS nanoparticles was homogeneously dispersed in polymer matrix. The polymer particles which are amorphous and present in large amount also contribute to the peak broadness. However the size of CdS nanoparticle could not be obtained from the XRD pattern using Scherrer formula owing to poor crystallinity. The morphology of CdS/SO3-P(S-DVB) nanocomposites were characterized by SEM and TEM. Figures 4.46(a) and 4.46(b) show the typical SEM images of CdS/SO3-P(S-DVB) nanocomposites with different magnification under the optimum condition. SEM showed small spherical particles with diameter 300 nm – 2 micrometer and some particles appeared to be in agglomerated form. These spherical particles should be ascribed to the aggregation of polymer particles during polymerization process. The low tendency of aggregation in the polymer matrix as shown in the SEM image suggested that the stability of the miniemulsion system was under control by the use of monomer as oil-phase during the polymerization process. SEM images showed the presence of CdS particles on the surface of SO3HP(S-DVB), suggesting that the CdS nanoparticles were small and dispersed on polymer surface. The dispersion of CdS on the surface of polymer was also identified from the corresponding change in colour of products. The Energy dispersive X-ray (EDX) analyses were also done and confirmed the existence of the elements of cadmium and sulfur in CdS/SO3-P(S-DVB) nanocomposite. 116 (a) (b) Figure 4.46: SEM images of 7.64 % CdS/SO3-P(S-DVB) nanocomposites at magnifications (a) 2,000 X and (b) 5,000 X 117 Figure 4.47: SEM images on the surface of CdS/SO3-P(S-DVB) at magnification 100,000 X The shape and size of the nanocomposite can be seen in the TEM micrographs of Figure 4.48, where the micrograph is shown the dark center and pale edge of the spheres. The dark domain observed in the TEM image indicates that it originated from CdS nanoparticles. Cadmium as heavy element developed a darker color compared to carbon and hydrogen in the polymer structure. Of interest, TEM observations showed cloud-like products were the main morphology, indicating that CdS nanoparticles were small and homogeneously dispersed in the polymer matrix. As can be seen, the spheres exhibit a diameter distribution in the range of 200 nm – 2 micrometer and average diameter of ~500 nm. However, it is difficult to observe the particle size and size distribution of CdS nanoparticles by visual inspection of the TEM images, due to the TEM image was obtained by a low accelerating voltage (70 keV), hence the a statistical description of the size distribution of CdS nanoparticles cannot be carried out. 118 (a) (b) Figure 4.48: TEM images of (a) non agglomerated and (b) agglomerated of CdS/SO3-P(S-DVB) nanocomposites 4.3.4 Dielectric and Electrical Properties of CdS/SO3-P(S-DVB) Nanocomposites Table 4.7 shows the dielectric constants of SO3 H-P(S-DVB) and CdS/SO3P(S-DVB) nanocomposites at various contents of CdS nanoparticles and at various frequencies. Over the frequencies range of 0.1 – 1,000 KHz the dielectric constant of CdS/SO3-P(S-DVB) nanocomposites, shows steep decrease from its initial high values. Figure 4.49 shows the dielectric constants measured at frequency of 100 Hz of SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites at various contents of CdS. The dielectric constants of CdS/SO3-P(S-DVB) nanocomposites decreased with increasing CdS nanoparticles content, as compared to SO3H-P(S-DVB). On the other hand, the dielectric constants of the CdS/SO3-P(S-DVB) nanocomposites showed higher values than P(S-DVB) (~6 measured at 100 Hz) and CdS nanoparticles (100 measured at 100 Hz). 119 Table 4.7: Dielectric constants of P(S-DVB), CdS nanoparticles, SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites Sample Log frequency ( Hz) 6 5 4 3 2 P(S-DVB) 1.4 2.9 2.9 3.1 5.9 CdS nanoparticles 4.6 16.2 28.9 49.8 100.0 Cd(SO3)2-P(S-DVB) 4.2 39.3 294.7 2145.1 9097.8 NaSO3-P(S-DVB) 8.6 55 359 1286.8 2016.6 2,338.0 10,947.0 25,345.0 61,634.0 236,190.0 2.56 % CdS/SO3-P(S-DVB) 1,989.0 7,425.0 13,931.0 30,069.0 87,118.0 3.34 % CdS/SO3-P(S-DVB) 1,281.0 7.702.0 14,933.0 26,988.0 74,394.0 7.64 % CdS/SO3-P(S-DVB) 144.0 923.0 2,862.0 6,326.0 22,861.0 14.65 % CdS/SO3-P(S-DVB) 46.0 297.0 1,090.0 3,662.0 15,683.0 16.56 % CdS/SO3-P(S-DVB) 18.4 39.5 145.0 391.0 1,309.0 236,190 SO3H-P(S-DVB) 250000 74,394 87,118 15,683 22,861 2016.6 5.9 50000 9097.8 100000 1,309 150000 100 Dielectric constant 200000 0 a b c d e f g h i j Figure 4.49: Dielectric constant of (a) P(S-DVB), (b) CdS nanoparticles, (c) Cd(SO3)2-P(S-DVB), (d) NaSO3-P(S-DVB), (e) SO3H-P(S-DVB), (f) 2.56 % CdS/SO3-P(S-DVB), (g) 3.34 % CdS/SO3-P(S-DVB), (h) 7.64 % CdS/SO3-P(SDVB), (i) 14.65 % CdS/SO3-P(S-DVB) and (j) 16.56 % CdS/SO3-P(S-DVB) at 100 Hz. 120 The possible explanation for this result is the higher strength of electrical properties of proton and polarity of SO3H-P(S-DVB) as a host matrix. The protons at the sulfonic acid functional groups are very easily moved and when parts of the protons were replaced with CdS, the amount of proton was decreased and the dielectric properties of the nanocomposites decreased correspondly. This phenomenon was supported by the decreasing meq H+/gram of SO3H-P(S-DVB) with increasing amounts of CdS nanoparticles amount in the CdS/SO3-P(S-DVB) (see Table 4.8). The low dielectric constants of Cd(SO3)2-P(S-DVB) and NaSO3-P(SDVB) also supported the above explanation. Nevertheless, this result showed substantially higher dielectric constant as compared to the dielectric constant of 20 % CdS/PVA (dielectric constant ~10) (El-Tantawy et al., 2004), CdS/PVA with average dispersibility in the range of 0.3 – 0.8 (dielectric constant about 120 – 250 at in over the range frequency 10 – 40 khz) (Ghosh et al., 2005) and TiO2/sulfonated styrene-b-(ethylene-ran-butylene)-b-styrene (S-SEBS) block copolymers (Yang and Kofinas, 2007). The dielectric constants of 0.26 % and 6.4 % of TiO2/S-SEBS nanocomposites at 10 kHz were around 2.15 and 3.2, respectively and the values were low when compared to un-crosslinked S-SEBS (dielectric constant of about 5,258 at 10 kHz and 160,000 at 1 Hz). Table 4.8. Meq H+/gram of SO3H-P(S-DVB) with degree of sulfonation = 39.10 % and CdS/SO3-P(S-DVB) nanocomposites with various amount of CdS Sample Meq H+/gram SO3H-P(S-DVB) 3.50 2.56 % CdS/SO3-P(S-DVB) 2.64 3.34 % CdS/SO3-P(S-DVB) 2.35 7.64 % CdS/SO3-P(S-DVB) 1.53 14.65 % CdS/SO3-P(S-DVB) 0.66 16.56 % CdS/SO3-P(S-DVB) 0.24 121 Figure 4.50 shows the graph of decreasing meq H+ in the CdS/SO3-P(S-DVB) nanocomposite versus dielectric constant at 100 Hz. The curve shows a good correlation with the coefficient correlation of 0.9171. 100,000 y = -23033x + 109372 Dielectric constant 80,000 2 R = 0.9171 60,000 40,000 20,000 0 0 1 2.64 2 2.35 3 1.53 4 0.66 5 0.24 6 Meq H+ Figure 4.50: A graph of the relationship between amount of meq H+ and dielectric constant of CdS/SO3-P(S-DVB) nanocomposites Table 4.9 and Figure 4.51 show the dissipation factors of CdS/SO3-P(S-DVB) nanocomposites at various amounts of CdS and frequencies. The dissipation factor values of the nanocomposites decreased as compared to the dissipation factor values of SO3H-P(S-DVB). These results suggested that part of the proton in the sulfonic functional groups were being replaced by sodium ions and CdS as filler and the mobility of proton in the polymer was also hindered by the CdS nanoparticles on the surface. 122 Table 4.9: Dissipation factors of P(S-DVB), CdS nanoparticles, SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites Sample Log frequency ( Hz) 6 5 4 3 2 P(S-DVB) 0.40 0.10 0.20 0.20 0.50 CdS nanoparticles 0.20 0.20 0.20 0.50 1.60 SO3H-P(S-DVB) 4.13 0.37 0.22 0.38 1.47 2.36 % CdS/SO3-P(S-DVB) 1.92 0.17 0.16 0.35 0.67 3.34 % CdS/SO3-P(S-DVB) 0.45 0.49 0.10 0.16 0.52 7.64 % CdS/SO3-P(S-DVB) 4.25 0.84 0.24 0.32 1.00 14.65 % CdS/SO3-P(S-DVB) 1.77 1.13 0.79 0.92 1.98 16.56 % CdS/SO3-P(S-DVB) 0.56 0.52 0.45 0.33 1.48 4.50 4.00 3.00 2.50 2.00 1.50 1.00 0.50 0.00 3 4 a 5 b c d 6 e f Log frequency (Hz) Dielectric constant 3.50 2 Figure 4.51: Dissipation factor of (a) SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) at various CdS content and frequencies. (b) 2.36 % CdS/SO3-P(S-DVB), (c) 3.34 % CdS/SO3-P(S-DVB), (d) 7.64 % CdS/SO3-P(S-DVB), (e) 14.65 % CdS/SO3-P(SDVB), (f) 16.56 % CdS/SO3-P(S-DVB) 123 The variation on dielectric constant of the CdS/SO3-P(S-DVB) nanocomposites with different temperature is shown in Figure 4.52. The nanocomposites display an increasing tendency in dielectric constant at frequencies 0.1 – 1,000 KHz with increasing of temperature in the range of 25 - 75 oC and then abruptly decreased after 75 oC. 140000 120000 100000 Dielectric constant 80000 60000 0.1 KHz Temperature oC 1 KHz 40000 20000 10 KHz 100 KHz 1000 KHz 0 0 25 50 75 100 125 150 200 Temperature oC Figure 4.52: Influence of temperature on the dielectric constant of 7.64 % CdS/SO3P(S-DVB) nanocomposites 124 This phenomenon is explained by the occurrence of two kinds of behaviors, which would yield converse effect on the dielectric constant of the nanocomposites by changing the temperature. First, increasing temperature would improve the segmental mobility of polymer and CdS nanoparticles, facilitate the polarization of polar fillers and consequently increase dielectric constant. Second, the obvious differential in thermal expansion coefficient between the polymer and CdS nanoparticles would disturb the aggregations of polar components (CdS naoparticles), thus reducing the dielectric constant. Over the temperature range of 25 – 75 oC for the CdS/SO3-P(S-DVB) nanocomposites showed that the first behavior was the main effect. On the other hand, at temperatures above 75 oC, the second behavior more dominant. It was likely that disruption of the CdS nanoparticles cluster was the cause dor the dielectric constant to decrease with rising temperature. The result was further supported by the glass transition temperature (Tg) obtained SO3H-P(S-DVB) of around 100 to 160 oC, depended on degree of sulfonated (Wallace, 1971; Yang et al., 2003). The result consistent with the dielectric constant of the nanocomposites measured after heat treatment which showed that the dielectric constant was slightly decreased, compared with that before heat treatment. Table 4.10 and Figure 4.53 show the ionic conductivities of P(S-DVB), SO3H-P(S-DVB), CdS nanoparticles and CdS/SO3-P(S-DVB) nanocomposites. The ionic conductivity values of CdS/SO3-P(S-DVB) nanocomposites decreased with increasing CdS nanoparticle content in the nanocomposites compared to SO3H-P(SDVB). The same reason which explained the dielectric properties can also be applied to explain this ionic conductivity property. When some the protons in sulfonic acid were replaced with sodium and CdS nanoparticles attached on the surface of polymer, the mobility of the proton residue was hindered. The result was also evidenced by the decrease of meq H+ in the nanocomposite with increasing amount of CdS (see Table 4.8) Figure 4.54 shows the relationship between meq H+ in the nanocomposite with ionic conductivity of CdS/SO3-P(S-DVB) nanocomposites. The curve shows a good correlation between meq H+ and decreasing the ionic conductivity of proton with coefficient correlation of 0.9913. 125 Table 4.10: Ionic conductivities of CdS nanoparticles, P(S-DVB), Cd(SO3)2-P(SDVB), NaSO3-P(S-DVB), SO3H-P(S-DVB), and Cd/SO3-P(S-DVB) with various amount of CdS Sample Ionic Conductivity S/ohm 1.6 E-10 CdS nanoparticles 6.2 E-08 Cd(SO3)2-P(S-DVB) 6.0-E-07 NaSO3-P(S-DVB) 2.0 E-07 SO3H-P(S-DVB) 1.1 E-04 2.36 % CdS/SO3-P(S-DVB) 1.7 E-05 3.34 % CdS/SO3-P(S-DVB) 1.7 E-05 7.64 % CdS/SO3-P(S-DVB) 1.2 E-05 14.65 % CdS/SO3-P(S-DVB) 6.4 E-06 16.56 % CdS/SO3-P(S-DVB) 1.8 E-06 1.10E-04 P(S-DVB) 1.20E-04 8.00E-05 d f g 1.80E-06 c 6.40E-06 b 1.20E-05 2.00E-07 a 1.70E-05 6.00E-07 2.00E-05 6.20E-08 4.00E-05 2.00E-05 6.00E-05 1.60E-10 Ionic conductivity (S/ohm) 1.00E-04 0.00E+00 e h i j Figure 4.53: Ionic conductivity of (a) Pure P(S-DVB), (b) CdS nanoparticles, (c) Cd(SO3)2-P(S-DVB), (d) NaSO3-P(S-DVB), (e) SO3H-P(S-DVB), (f) 2.36 % CdS/SO3-P(S-DVB), (g) 3.34 % CdS/SO3-P(S-DVB), (h) 7.64 % CdS/SO3-P(SDVB), (i) 14.65 % CdS/SO3-P(S-DVB) and (j) 16.56 % CdS/SO3-P(S-DVB) 126 Ionic conductivity (S/ohm) 2.50E-05 y = -5E-06x + 3E-05 R2 = 0.9913 2.00E-05 1.50E-05 1.00E-05 5.00E-06 0.00E+00 0 1 2.64 2 2.35 3 1.53 Meq H+ 4 0.66 5 0.24 6 Figure 4.54: A graph of the relationship between amount of meq H+ and ionic conductivity of CdS/SO3-P(S-DVB) nanocomposites The ionic conductivity of the 7.46 % CdS/SO3-P(S-DVB) nanocomposites at various temperatures is shown in Figure 4.55. Similar behavior and reasons explained for dielectric properties result of the nanocomposites were observed in the ionic conductivity of the nanocomposites. Ionic conductivity (S/ohm) 2.50E-04 2.00E-04 1.50E-04 1.00E-04 5.00E-05 0.00E+00 30 40 50 60 70 80 90 100 Temperature oC Figure 4.55: Ionic conductivity of 7.64 % CdS/SO3-P(S-DVB) nanocomposite at different temperatures 127 4.3.5 Thermal Properties of CdS/SO3-P(S-DVB) Nanocomposites Figure 4.56 shows the thermal conductivity of P(S-DVB), SO3H-P(S-DVB), CdS nanoparticles and CdS/SO3-P(S-DVB) nanocomposites. The thermal conductivity of the nanocomposites increased with increasing CdS nanoparticles content. Two possible reasons is suggested to explain the results. Firstly, with increasing CdS nanoparticle content, the conductive path of the nanocomposites was enhanced as consequence of higher amount of CdS nanoparticles. CdS nanoparticles are more effective in heat conduction through the polymer matrix as host of the nanocomposites to transfer heat through the specimen. Secondly, the interfacial thermal resistance among CdS nanoparticles-polymer is reduced, leading the increase thermal conductivity. The result concluded that the thermal conductivity of nanocomposites was mostly dependent on the content and properties of the CdS nanoparticles. Similar results and reasons were reported in the case of CdS/PVA nanocomposites (El-Tantawy, 2004). 16.56 % CdS/SO3-P(S-DVB) 14.65 % CdS/SO3-P(S-DVB) 7.64 % CdS/SO3-P(S-DVB) 3.34 % CdS/SO3-P(S-DVB) 2.36 % CdS/SO3-P(S-DVB) CdS nanoparticles SO3H-P(S-DVB) Pure P(S-DVB) 0.24 0.23 0.22 0.20 0.19 0.22 0.14 0.09 Thermal conductivity (W/mK) Figure 4.56: Thermal conductivity of P(S-DVB), SO3H-P(S-DVB), CdS nanoparticles, and CdS/SO3-P(S-DVB) nanocomposites 128 Figure 4.57 show the relationship between amount of CdS nanoparticle in the nanocomposite with thermal conductivity of CdS/SO3-P(S-DVB) nanocomposites. The curve shows a good correlation between the increasing amount of CdS with thermal conductivity with coefficient correlation of 0.9910. Thermal conductivity (W/mK) 0.3 y = 0.0137x + 0.1809 R2 = 0.991 0.25 0.2 0.15 0.1 0.05 2.56 3.34 7.64 14.65 16.56 0 0 1 2.56 2 3.34 3 7.64 4 14.65 516.56 6 % CdS Figure 4.57: A graph of the relationship between amount of CdS nanoparticle and thermal conductivity of CdS/SO3-P(S-DVB) nanocomposites Figure 4.58 shows the TGA curves of the CdS nanoparticles, SO3H-P(SDVB) and CdS/SO3-P(S-DVB) nanocomposites with different CdS nanoparticles contents. Thermogravimetric analyses were carried out under oxygen atmosphere in the temperature range of 40 – 1000 oC. The TGA curves for SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites show a two stage weight losses i.e. from 50 – 200 oC and 300 – 600 oC. The mass loss at 50 – 200 oC in SO3H-P(S-DVB) and CdS/SO3-P(S-DVB) nanocomposites is due to the loss of physically and chemically bounded water. The observed mass loss between 310 - 600 oC in both samples may be attributed to the decomposition of the sulfonic acid group, which on cleavage will eliminate SO3. The mass loss in the range 310 – 600 oC therefore correlated with the amount of sulfonic acid groups on the basis of the assumption that it represents the decomposition of the – SO3H groups only. An improved thermal stability of the 129 CdS/SO3-P(S-DVB) nanocomposites was observed as compared to the SO3H-P(SDVB). The possible reason for the increased in thermal stability of the nanocomposites is reduced molecular mobility of polymer chains due to presence of the CdS nanoparticles. This was supported by the TGA curve of CdS nanoparticles in which it showed a high thermal stability. It can also be seen that the residue of the nanocomposites after the temperature of 600 oC is about 2 - 12 %, while the SO3HP(S-DVB) was completely decomposed at these temperature and left no residue. Similar thermal behaviour was obtained in the cases of CdS/PVA (El-Tantawy, 2004), CdS/P(S-co-Mac) (Nair et al., 2005), CdS/PS (Kuljanin et al., 2002) and CdS/PMMA (Kuljanin et al., 2006). a Weight loss (%) 100 80 c 60 d 40 b 20 0 e 100 200 300 400 500 600 700 800 900 1000 Temperature (oC) Figure 4.58: Thermogravimetric analysis curves of (a) CdS nanoparticles, (b) 16.56 % CdS/SO3-P(S-DVB), (c) 7.64 % CdS/SO3-P(S-DVB), (d) 2.36 % CdS/SO3-P(SDVB) and (e) SO3H-P(S-DVB) Figure 4.59 shows the summary in schematic representation to explain the dielectric properties, electrical conductivity and thermal conductivity of CdS/SO3P(S-DVB) nanocomposites 130 SO3Na HSO3 HSO3 HSO3 HSO3 HSO3 HSO3 SO3Na HSO3 HSO3 HSO CdS Polymer particle 3 HSO3 SO3Na HSO3 HSO3 HSO3 HSO3 HSO3HSO HSO3 3 Increase of amount and particles size of CdS SO3Na HSO3 HSO3 HSO3 HSO3 HSO3 HSO3 SO3Na HSO3 HSO3 HSO 3 SO3Na HSO3 HSO3 SO3Na SO3Na HSO3 HSO3 HSO 3 HSO Decrease of meq proton Decrease of dielectric constant Decrease of ionic conductivity 3 Increase of thermal conductivity SO3Na HSO3 HSO3 SO3Na HSO3 NaSO3 HSO3 SO3Na NaSO3 HSO3 HSO 3 SO3Na SO3Na HSO3 HSO3 NaSO3 SO3Na HSO 3 SO3Na HSO3 Figure 4.59: Schematic representation of mechanism replacing proton with sodium and CdS nanoparticles attached on surface of polymer and their properties 131 4.4 Comparison of Physicochemical and Electrical Properties of CdS/polymer Nanocomposites A summary of the results obtained on the physicochemical and electrical properties of CdS/polymer nanocomposites are presented in Tables 4.11 and 4.12. The different procedures of synthesis were applied in order to disperse CdS nanoparticles on the polymer surface or to be encapsulated within the polymer matrices. Since the chemical properties of polymers are different. Three different methods were applied to synthesize the nanocomposites. The first type, CdS/P(SDVB) nanocomposite was prepared by in-situ polymerization in miniemulsion system using styrene-divinyl benzene monomers (entry 1 in Table 4.11). The second type, CdS/P(MAA-EGDMA) nanocomposite was prepared with three steps procedures i.e., in-situ polymerization in miniemulsion system to produce P(MAAEGDMA), ion exchange and precipitation process to produce the nanocomposite (entry 2 in Table 4.11). Lastly, the CdS/SO3-P(S-DVB) nanocomposite was prepared with four steps procedure i.e., in-situ polymerization in miniemulsion system to produce P(S-DVB) particles, sulfonation reaction with fuming sulfuric acid (H2S2O7) to produce SO3H-P(S-DVB), ion exchange and precipitation process to produce the nanocomposites (entry 3 in Table 4.11). The morphology and particles size of the polymers were observed using SEM and TEM techniques. The SEM and TEM studies showed that the particles size of P(S-DVB) and SO3H-P(S-DVB) (entry 1 and 3 in Table 4.11) is in the range of 0.2 – 2 μm, while the particle size of P(MAA-EGDMA) is in the range of 200 – 300 nm (entry 2 in Table 4.11). The particles size of CdS nanoparticles in the miniemulsion and these nanocomposites was examined using UV – Vis and UV - Vis DR spectroscopy and it showed a blue shifted onset wavelength when compared to bulk CdS, indicating a quantum confinement effect associated with the nanometer particles. There is no significant difference in size of CdS nanoparticles in these polymer nanocomposites. CdS/SO3-P(S-DVB) c, j 3 using styrene-divinyl benzene as CdS/P(MAA-EGDMA) b, i 2 h at CdS 0.18 %, at 1.08 %. i Post synthesize d b High Medium 0.3 – 2 μm e using Brus equation, f 7 – (>8) 7.1 – 8.0 3–5 >500 – 600 oC Increased >350 – 400 oC Increased >350 – 400 oC Increased stabilityf,g Thermal g as compared to pure polymer, polymer was prepared by in-situ TGA analysis, c 200 – 300 nm 0.2 – 2 μm of CdS (nm)e of polymerd of polymer Low Particle size Particles size Relative polarity Physicochemical properties polymer was prepared by in-situ polymerization, from SEM and TEM images, oil-phase, ( ion exchange-precipitation) Post synthesize (ion exchange-preciptitation) polymerization and sulfonation reaction, a In-situ polymerization CdS/P(S-DVB) a, h 1 in miniemulsion system Preparation methods Nanocomposite materials Entry Table 4.11: Physicochemical properties of cadmium sulfide/polymer nanocomposites 132 133 The SEM study clearly showed that the CdS nanoparticles were attached on the surface of P(MAA-EGDMA) and SO3H-P(S-DVB). Meanwhile, the CdS nanoparticles were encapsulated in P(S-DVB) and not observed in the SEM image. These results suggested that the in-situ formation by in-situ polymerization in a miniemulsion system and ion exchange-precipitation in these polymer matrices could prevent agglomeration and control the growth of CdS nanoparticles. As tabulated in Table 4.12, the dielectric properties of three types of CdS/polymer nanocomposite displayed significant difference in the dielectric constant values and low dissipation factor. The CdS/SO3-P(S-DVB) (entry 3 in Table 4.12) had higher dielectric constant value as compared to CdS/P(MAA-EGDMA) (entry 2 in Table 4.12) and CdS/P(S-DVB) (entry 1 in Table 4.12), which is in the following order: CdS/SO3-P(S-DVB) > CdS/P(MAA-EGDMA) > CdS/P(S-DVB). One demonstrates that the difference in dielectric properties of these nanocomposites is affected by the intrinsic polymer properties, amount and particles size of CdS nanoparticles and also the interfacial interaction between CdS nanoparticle and polymer. However, when bulk CdS was used in polymer nanocomposite, the dielectric properties is low, suggesting that the dielectric properties of nanocomposites were also dependent on the size of CdS particles. In conclusion, the change of dielectric constant of polymers can be considered by the following factors, i.e., nanosize of CdS and the interfacial interaction between CdS and polymers. As shown in Table 4.12, there is a correlation between the electrical conductivity and dielectric constant. The dielectric constant of P(S-DVB) and P(MAA-EGDMA) polymers (entry 1 and 2 in Table 4.12) increases with the addition of CdS nanoparticles. This phenomenon is also applicable for electrical conductivity. However, in case of SO3H-P(S-DVB), the electrical conductivity and dielectric constant of the polymer decreased with the addition of CdS nanoparticles. The simple explanation for the decrease of the electrical properties (electrical conductivity and dielectric constant) is that, the attachment of CdS nanoparticle on the surface of SO3H-P(S-DVB) caused by the replacement of the proton with the CdS nanoparticles. It is generally accepted that very high dielectric constant and electrical conductivity are due to the proton mobilty. 134 The thermal conductivity (entry 1, 2, and 3 in Table 4.12) and thermal stability (entry 1, 2, and 3 in Table 4.11) of these nanocomposites have also been observed. CdS/SO3-P(S-DVB) nanocomposite had higher thermal conductivity and thermal stability as compared to CdS/P(MAA-EGDMA) and CdS/P(S-DVB). One demonstrates that the higher electrical conductivity and thermal stability of CdS/SO3P(S-DVB) nanocomposite is affected by the intrinsic polymer properties, especially from sulfonated functional groups a e (1.47) (1.6) d 236,190 g (0.24) (1.6) 100 9.2 (0.37) (1.6) 100 5.9 100 (0.67) 87,118 (0.53) 5245.6 (1.2) 230.4 Composite CdS 6.2 x 10-8 6.2 x 10-8 6.2 x 10-8 1.14 x 10-4 g 2.90 x 10-9 1.63 x 10-10 Polymer 1.73 x 10-5 2.94 x 10-6 2.62 x 10-9 Composite at CdS 0.18 %, at CdS 1.08 %, at CdS 2.07 %, at 100 Hz, at room temperature, fat room temperature, gSO3H-P(S-DVB) CdS/SO3-P(S-DVB) c 3 c CdS/P(MAA-EGDMA) b 2 b CdS/P(S-DVB) a Polymer (S ohm-1) e (Dissipation factor) d Materials CdS Ionic conductivity Dielectric constant 0.22 0.22 0.22 CdS 0.1490 g 0.0907 0.0950 Polymer 0.1940 0.1493 0.1400 Composite (W m-1K-1) f Thermal conductivity Electrical properties Nanocomposite 1 Entry Table 4.12: Electrical properties of cadmium sulfide/polymer nanocomposites 135 136 CHAPTER 5 CONCLUDING REMARKS In this thesis, the physicochemical and electrical properties of CdS/polymer nanocomposites have been studied. The results from the study have proven that CdS/polymer nanocomposites possesses high dielectric properties compared to those of silica. The dielectric properties of CdS polymer nanocomposites changed due to the presence of CdS nanoparticles in polymer matrices. Three types of CdS/polymer nanocomposites with high dielectric constant have been successfully designed and synthesized. The first type, CdS/P(S-DVB) nanocomposite was prepared by in-situ polymerization in miniemulsion system using styrene-divinyl benzene as a monomer. The second type, CdS/P(MAA-EGDMA) nanocomposite, was prepared by in-situ polymerization of P(MAA-EGDMA), followed by ion exchange and then precipitation process. Lastly, the CdS/SO3-P(S-DVB) nanocomposite was prepared by in-situ polymerization of P(S-DVB), sulfonation to produce SO3H-P(S-DVB), ion exchange and then precipitation process. The dielectrical property on P(MAA-EDMA) with nanosized CdS particles attached was studied. The attachment of CdS nanoparticles alters the dielectric property of CdS/P(MAA-EDMA) samples. By addition of a certain amount of CdS nanoparticles, the dielectric constant at room temperature and 100 Hz of the CdS/P(MAA-EDMA) is enhanced by ca. 500 times, which is attributed to the interfacial interaction of CdS nanoparticles with polymer surface. The large value of dielectric constant is due to interfacial space charge polarization. The same phenomenon was also exhibited in CdS/P(S-DVB) nanocomposite. 137 The dielectric of constant of CdS/SO3-P(S-DVB) nanocomposite was decrease as compare to pure sulfonated polymer. As SO3H-P(S-DVB) contains mobile proton which is responsible for high dielectric constant, the attachment of CdS reduced the dielectric constant due to the replacement of proton with CdS. The dielectric properties of three types of CdS/polymer nanocomposite displayed significant difference in the dielectric constant value. The CdS/SO3-P(SDVB) had higher dielectric constant value as compared to CdS/P(MAA-EGDMA) and CdS/P(S-DVB), which is in the following order: CdS/SO3-P(S-DVB) > CdS/P(MAA-EGDMA) > CdS/P(S-DVB). Based on the above results one can suggests that the enhancement of dielectric constant of CdS/polymer nanocomposites can be considered by the following unique properties of CdS/polymer nanocomposite, i.e., nanosize of CdS and the interfacial interaction between CdS and polymer. A highly dielectric constant CdS/polymer nanocomposite may be prepared by ion-exchange of cadmium and followed by reaction with sulfide on the surface of polymer and cannot be achieved by mechanical mixing of CdS and polymer. It is demonstrated that the above phenomena can be applied to CdS-P(MAA-EDMA) nanocomposite prepared by insitu polymerization of P(MAA-EGDMA), followed by ion exchange and then precipitation process. Based on the above results, one concludes that the attachment of CdS nanoparticles on the surface of P(S-DVB), P(MAA-EGDMA) and SO3H-P(S-DVB) polymers can control the dielectrical properties of CdS/polymer nanocomposites. 138 REFERENCES Agostiano, A., Catalano, M., Curri, M. L., Monica, N. D., Manna, L. and Vasanelli, L. (2000). Synthesis and Structural Characterization of CdS Nanoparticles Prepared in a Four-Components “Water-in-Oil” Microemulsion. Micron. 31: 253-258. Alducin, J.A., Forcada, J. and Asua, J.M. (1994). Miniemulsion polymerization using oil-soluble initiators. Macromolecules 27 :2256–2261 Alivisatos, A. P., (1996). Perspectives on the Physical Chemistry of Semiconductor. Nanocrystals. Journal of Physical Chemistry. 100(31): 13226-13239. Ameen, S., Ali, V., Zulfequar, M., Haq, M. M. and Husain, M. (2007). Electrical Conductivity and Dielectric Properties of Sulfamic Acid Doped Polyaniline. Current Applied Physics. 7(2): 215 – 219. Antolini, F., Pentimali, M., Di Luccio, T., Terzi, R., Schioppa, M., Re, M., Mirenghi, L. and Tapfer, L. (2005). Structural Characterization of CdS Nanoparticles Grown in Polystyrene Matrix by Thermolytic Synthesis. Materials Letters. 59(24-25): 3181-3187. Asua, J. M. (2002). Miniemulsion Polymerization. Progress in Polymer Science. 27: 1283 – 1346. Banerjee, R., Jayakrishnan, R. and Ayyub, P. (2000). Effect of the Size-Induced Structural Transformation on The Band Gap in CdS Nanoparticles. Journal . of Physics. : Condensed Matter. 12: 10647-10654. Bard, A. J., (1980). Photoelectrochemistry. Science. 207. 4427:139 -144. Brus, L. E. (1986). Electronic Wave Functions in Semiconductor Clusters: Experiment and Theory. Journal of Physical Chemistry. 90(12): 2555-2560. Brus. L. E. (1984). Electron–Electron and Electron-Hole Interactions in Small Semiconductor Crystallites: The Size Dependence of the Lowest Excited Electronic State. Journal of Chemical Physics. 80(9): 4403-4409. 139 Bunker, C. E., Harruff, B. A., Pathak, P., Payzant, A., Allard, L. F. and Sun, Y. P. (2004). Formation of Cadmium Sulfide Nanoparticles in Reverse Micelles: Extreme Sensitivity to Preparation Procedure. Langmuir. 20: 5642-5644. Chen, H. M., Huang, X. F., Xu, L, Xu, J., Chen, K. J. and Feng, D. (2000). Selfassembly and Photolumineccence of CdS-Mercaptoacetic Clusters with Internal Structures. Superlattices and Microstructures. 27(1): 1-5. Chen, H., Xu, F., Hong, S. and Wang, L. (2006). Quantitative Determination of Protein at Nanogram Levels by the Resonance Light-Scattering Technique with Composite Nanoparticles of CdS/PAA. Spectrochimica Acta Part A. 65: 428-432. Chen, M., Xie, Y., Qiao, Z., Zhu, Y. and Qian, Y. (2000). Synthesis of Short CdS Nanofiber/Poly(styrene-alt-maleic anhydride) Composites using Gammairradiation. Journal of Materials Chemistry. 10: 329-332. Chen, X., Wang, X., Liu, L., Yang, D. and Fan, L. (2005). Functionalized Semiconductor Nanocrystals for Ultrasensitive Detection of Peptides. Analytica Chimica Acta. 542(2): 144-150. Chestnoy, N., Harris, T. D., Hull, R. and Brus, L. E. (1986). Luminescence and Photophysics of Cadmium Sulfide Semiconductor Clusters: The Nature of the Emitting Electronic State. Journal of Physical Chemistry. 90(15): 3393-3399. Crooks, R. M., Zhao, M., Sun, L., Chechik, V. and Yeung, L. K. (2001). DendrimerEncapsulated Metal Nanoparticles: Synthesis, Characterization, and Applications to Catalysis. Accounts of Chemical Research. 34(3): 181-190. Dang, Z. M., Yu, Y. F., Xu, H. P. and Bai, J. (2007). Study on Microstructure and Dielectric Property of The BaTiO3/Epoxy Resin Composites. Composites Science and Technology. 68 : 171 – 177. Dellinger, T. M. and Braun, P. V. (2001). BiOCl Nanoparticles Synthesized in Lyotropic Liquid Crystal Nanoreactors. Scripta Materialia. 44 (8-9): 18931897. Devaraju, N. G., Kim, E. S. and Lee, B. I. (2005). The Synthesis and Dielectric Study of BaTiO3/Polyimide Nanocomposite Films. Microelectronic Engineering. 82: 71-83. Dey, A., De, S., De, A., and De, S. K. (2004). Characterization and Dielectric Properties of Polyaniline-TiO2 Nanocomposites. Nanotechnology. 15 : 12771283. 140 Dong, L., Xiong, C., Quan, H. and Zhao, G. (2006). Polyvynil-Butyral/Lead Zirconate Titanates Composites with High Dielectric Constant and Low Dielectric Loss. Scripta Materialia. 55: 835-837. Dutta, P. and Fendler, J. H. (2002). Preparation of Cadmium Sulfide Nanoparticles in Self-Reproducing Reversed Micelles. Journal of Colloid and Interface Science. 247(1): 47-53. El-Tantawy, F, Abdel-Kader, K. M., Kaneko, F. and Sung, Y. K. (2004). Physical Properties of CdS-Poly(vinyl alcohol) Nanoconducting Composite Synthesized by Organosol Techniques and Novel Application Potential. European Polymer Journal. 40: 415-430. Erdem, B., Sudol,. D., Dimonie, V. L. and El-Aasser, M. S. (2000a). Encapsulation of Inorganic Particles via Miniemulsion Polymerization. II. Preparation and Characterization of Styrene Miniemulsion Droplets Containing TiO2 Particles. Journal of Polymer Science Part. A : Polymer Chemistry. 38: 4431 – 4440. Erdem, B., Sudol,. D., Dimonie, V. L. and El-Aasser, M. S. (2000b). Encapsulation of Inorganic Particles via Miniemulsion Polymerization. II. Characterization of Encapsulation. Journal of Polymer Science Part. A : Polymer Chemistry. 38: 4441 – 4450. Faridi-Majidi, R., Sharifi-Sanjani, N. and Agend, F. (2006). Encapsulation of Magnetic Nanoparticles with Polystyrene via Emulsifier-free Miniemulsion Polymerization. Thin Solid Films. 515: 368-374. Forster, S., Konrad, M. (2003). From Self-Organizing Polymers to Nano- and Biomaterials. Journal of Material Chemistry. 13: 2671 – 88. Ghosh, P. K., Maity, R. and Chattopadhyay, K. K., (2005a). CdS Nanoparticles Grown in a Polymer Matrix by Chemical Bath Deposition. Journal of Nanoscience and Nanotechnology. 5: 1-6. Ghosh, P. K., Mitra, M. K. and Chattopadgyay, K. K. (2005b). ZnS Nanobelts Grown in a Polymer Matrix by Chemical Bath Deposition. Nanotechnology. 16: 107-112. Graetzel, C. K. and Graetzel, M. (1979). Hydrogen Evolution from The Photolysis of Alcoholic Benzophenone Solutions via Redox Journal of the American Chemical Society. 101(26): 7741-7743. Catalysis 141 Graff, A., Winterhalter, M. and Meier, W.(2001). Nanoreactors from PolymerStabilized Liposomes. Langmuir. 17(3): 919-923. Hagfeldt, A. and Graetzel, M. (1995). Light-Induced Redox Reactions in Nanocrystalline Systems. Chemical Reviews. 95(1): 49-68. Hansen, F. K. and Ugelstad, J. (1979a). Particle Nucleation in Emulsion Polymerization. II. Nucleation in Emulsifier-Free Systems Investigated by Seed Polymerization. Journal of Polymer Science Part A: Polymer Chemistry. 17:3033–3045. Hansen, F. K. and Ugelstad, J. (1979b). Particle Nucleation in Emulsion Polymerization. III. Nucleation in Systems with Anionic Emulsifier Investigated by Seeded and Unseeded Polymerization. Journal of Polymer Science Part A: Polymer Chemistry. 17, 3047–3067. Haram, S. K., Quinn, B. M. and Bard, A. J. (2001). Electrochemistry of CdS Nanoparticles: A Correlation between Optical and Electrochemical Band Gaps. Journal of American Chemical Society. 123(36): 8860-8861. Harruff, B. A. and Bunker, C. E. (2003). Spectral Properties of AOT-Protected CdS Nanoparticles: Quantum Yield Enhancement by Photolysis. Langmuir. 19: 893-897. Henglein, A. (1989). Small-Particle Research: Physicochemical Properties of Extremely Small Colloidal Metal and Semiconductor Particles. Chemical Reviews. 89(8): 1861-1873. Hines, M. A. and Guyot-Sionnest, P. (1998). Bright UV-Blue Luminescent Colloidal ZnSe Nanocrystals. Journal of Physical Chemistry B. 102(19): 3655-3657. Hirai, T. and Bando, Y. (2005). Immobilization of CdS Nanoparticles Formed in Reverse Micelles onto Aluminosilicate Supports and Their Photocatalytic Properties. Journal of Colloid and Interface Science. 288 (2): 513-516. Hirai, T. and Komasawa, I. (2000a). Preparation of Nano-CdS-Polyurethane Composites via In-situ Polymerization in Reverse Micellar System. Journal. of Material. Communication. 10: 2234-2235. Hirai, T. and Ota, M. (2006). Immobilization of CdS Nanoparticles from Reverse Micellar System onto Mesoporous Organosilicates and Their Photocatalytic Properties. Materials Research Bulletin. 41: 19-28. 142 Hirai, T., Bando, Y. and Komawasa, I. (2002a). Immobilization of CdS Nanoparticles Formed in Reverse Micelles onto Alumina Particles and Their Photocatalytic Properties. Journal of Physical Chemistry B. 106: 8967-8970. Hirai, T., Miyamoto, M. and Komasawa, I. (1999). Composite Nano-CdSPolyurethane Tranparent Films. Journal of Material Chemistry. 9: 12171219. Hirai, T., Namba, M. and Komasawa, I. (2002b). Dithiol-Mediated Immobilization of CdS Nanoparticles from Reverse Micellar System onto Zn-Doped Silica Particles and Their High Photocatalytic Activity. Journal of Colloid and Interface Science. 252(1): 89-92. Hirai, T., Okubo, H. and Komasawa, I. (1999). Size-Selective Incorporation of CdS Nanoparticles into Mesoporous Silica. Journal of Physical Chemistry B. 103(21): 4228-4230. Hirai, T., Okubo, H. and Komasawa, I. (2000b). Incorporation of CdS Nanoparticles Formed in Reverse Micelles into Silica Matrices via Sol-gel Process: Preparation of Nano-CdS-containing Silica Colloids and Silica Glass. Journal of Material Chemistry. 10: 2592-2596. Hirai, T., Saito, T. and Komasawa, I. (2001). Stabilization od CdS Nanoparticles Immobilized on Thiol-Modified Polystyrene Particles by Encapsulation with Polythiourethane. Journal of Physical Chemistry B. 105: 9711-9714. Hirai, T., Watanabe, T. and Komasawa, I. (2000c). Preparation of Semiconductor Nanoparticle-Polymer Composites by Direct Reverse Micelle Polymerization Using Polymerizable Surfactants. Journal of Physical Chemistry B. 104(38): 8962-8966. Hong, J. I., Winberg, P., Schadler, L. S. and Siegel R. W. (2005). Dielectric Properties of Zinc Oxide/Low Density Polyethylene Nanaocomposites. Materials Letters. 59: 473-476. Iwasaki, K., Torimoto, T., Shibayama, T., Takahashi, H. and Othani, B. (2004). Preparation and Characterization of Water-Soluble Jingle-Bell-Shaped SilicaCoated Cadmium Sulfide Nanoparticles. Journal of Physical Chemistry B. 108: 11946-11952. Jiang, D., Ding, L. and Huang, J. (2006). Synthesis and Characterization of a Novel Poly(N-Vynil)-3-[p-Nitrophenylazo]carbazolyl-CdS Nanocomposites through Chemical Hybridization. Materials Letters. 60 (29-30): 3457-3462. 143 Johnson, B. J. S., Wolf, J. H., Zalusky, A. S. and Hillmyer, M. A. (2004). Template Syntheses of Polypyrrole Nanowires and CdS Nanoparticles in Porous Polymer Monoliths. Chemistry of Materials. 16: 2909-2917. Kalakkunnath, S and Kalika, D. D. (2006). Dynamic Mechanical and Dielectric Relaxation Characteristics of Poly(trimethylene terephthalate). Polymer. 47(20): 7085-7094. Khanna, P. K., Gokhale, R. R., Subbarao, V. V. V. S., Singh, N., Jun, K-W. and Das, B. K. (2005). Synthesis and Optical Properties of CdS/PVA Nanocomposites. Materials Chemistry and Physics. 94: 454-459. Khanna, P. K., Lonkar, S. P. Subbarao, V. V. V. S. and Jun, K. W. (2004). Polyaniline-CdS Nanocomposite from Organometallic Cadmium Precursor. Materials Chemistry and Physics. 87: 49-52. Khiew, P. S., Huang, N. M., Radiman, S. and Ahmad M. S. (2004). Synthesis and Characterization of Conducting Polyaniline-Coated Cadmium Sulfide Nanocomposite in Reverse Microemulsion. Materials Letters. 58: 516-521. Khiew, P. S., Radiman, S., Huang, N. M. and Md. Soot Ahmad.(2003). Studies on The Growth and Characterization of CdS and PbS Nanoparticles Using Sugar-Ester Nonionic Water-in-oil Microemulsion. Journal of Crystal Growth. 254(1-2): 235-243. Kremer, F. and Schönhals, A. (2002). Broadband Dielectric Spectroscopy. Springer. Kuljanin, J., Vuckovic, M., Comor, M. I., Bibic, N., Djokovic, V. and Nedeljkovic, J. M. (2002). Influence of CdS-Filler on The Thermal Properties of Polystyrene. European Polymer Journal. 38: 1659-1662. Kuljanin, J., Stojanovic, Z. and Nedeljkovic, J. M. (2006). Influence of CdS-Filler on The Thermal Properties of Poly(methylmethacrylate). Journal of Materials Science. 41: 5014-5016. Kumar, S. and Chandra, R. (2005). Temperature Dependent Studies of CdS nanoparticles in Viscous Matrix. Optical Materials. 27: 1346-1349. Lai, C.-Y., Trewyn, B. G., Jeftinija, D. M., Jeftinija, K., Xu, S., Jeftinija, S. and Lin, V. S.-Y. (2003). A Mesoporous Silica Nanosphere-Based Carrier System with Chemically Removable CdS Nanoparticle Caps for Stimuli-Responsive Controlled Release of Neurotransmitters and Drug Molecules. Journal of American Chemical Society. 125(15): 4451-4459. 144 Landfester, K. and Ramirez, L. (2003). Encapsulated magnetite particles for biomedical application. Journal of Physics: Condensed Matter. 15(15):S1345-1362. Lemon, B. I. And Crooks, R. M. (2000). Preparation and Characterization of Dendrimer-Encapsulated CdS Semiconductor Quantum Dots. Journal of American Chemical Society. 122: 12886-12887. Li, J. Y. (2003). Exchange Coupling in P(DVF-TrFE) Copolymer Based All-Organic Composites with Giant Electrostriction. Physical Review Letters., 90 (2003) 217601-1 – 217601-4. Li, C., Xiaobao, L., Yaobang, Z. and Anbing, T. (2006). Synthesis and Dielectric Property of Polyarylene Ether Nitriles/Titania Hybrid Films. Thin Solid Films. 515(4): 1872-1876. Li, H., Liu, G., Liu, B., Chen, W. and Chen, S. (2007). Dielectric Properties of Polyimide/Al2O3 Hybrids Synthesized by In-situ Polymerization. Materials Letters. 61(7): 1507-1511. Libert, S., Gorshkov, V., Privman, V., Goia, D. and Matijevic, E. (2003). Formation of Monodispersed Cadmium Sulfide Particles by Aggregation of Nanosize Precursors. Advances in Colloidal and Interface Science. 100-102: 169 -183. Lim, M.S. and Chen, H. (2000). Miniemulsion Polymerization of Styrene with a Block Copolymer Surfactant. Journal of Polymer Science. Part A: Polymer Chemistry., 38, 1818-1827 Lu, J., Moon, K. S., Kim, B. K. and Wong, C. P. (2007). High Dielectric Constant Polyaniline/Epoxy Composites via In-situ Polymerization for Embedded Capacitor Application. Polymer. 48: 510-1516. Lu, J., Moon, K.S., Xu, J. and Wong, C. P. (2006). Synthesis and Dielectric Properties of Novel High-K Polymer Composites Containing In-situ Formed Silver Nanoparticles for Embedded Capacitor Applications. Journal of Materiasl Chemistry. 16: 1543-1548. Martins, C.R., Ruggeri, G. De Paoli, M.A. (2003). Synthesis in Pilot Plant Scale and Physical Properties of Sulfonated Polystyrene. Journal of Brazilian Chemical Society. 14: 797 – 802. Mo, T.C., Wang, H. W., Cen, S. Y., and Yeh, Y. C. (2008). Synthesis and Dielectric Properties of Polyaniline/titanium dioxide Nanocomposites. Ceramics International. 34: 1767 – 1771. 145 Mondal, S.P, Muulick, H., Lavanya, T., Dhar, A., and Ray, S. K. (2007). Optical and Dielectric Properties of CdS Nanocomposites Embedded in Polymer Matrix. Journal Applied Physics. 102. Nair, P. S., Radhakrishnan, T., Revaprasadu, N., Kolawole, G. A., Luyt, A. S. and Djokovic, V. (2005). Polystyrene-co-Maleic Acid/CdS Nanocomposites: Preparation and Properties. Journal of Physics and Chemistry of Solids. 66: 1302-1306. Ni, Y., Ge, X. and Zhang, Z. (2006). Fabrication of CdS/Polyacrylonitrile Nanocomposites by γ-irradiation in an Ethanol Solution. Materials Science and Engineering B. 130: 61-65. Nunes, J. S., de Vasconcelos, C. L., Cabral, F. A. O., de Araujo, J. H., Pereira, M. R. and Fonseca, J. L. C. (2006) Synthesis and Characterization of Poly(ethyl methacrylate-co-methacrylic acid) Magnetic Particles via Miniemulsion Polymerization. Polymer. 47: 7646-7652. Olshasky, M. A. and Allcock, H. R. (1997). Synthesis of CdS Nanoparticles in Solution and in a Polyphosphazene Matrix. Chemistry of Materials. 9: 13671376. Pant, H. C., Patra, M. K., Verma, A., Vadera, S. R. and Kumar, N. (2006). Study of The Dielectric Properties of Barium Titanate-Polymer Composites. Acta Materialia. 54: 3163-3169. Podgrabisnki, T., Scorcik, V., Mackova, A., Hnatowiez, V. and Sajdi, P. (2006). Dielectric Properties of Doped Polystyrene and Polymethylmethacrylate. Journal of Materials Science: Materials in Electronic. 17: 871 – 875. Prijamboedi, B., Takashima, H., Wang, R., Shoji, A. and Itoh, M. (2005). Large Dielectric Constant Arising from Space-Charge Polarization in a SrTiO3 Thin Film Grown on a YBa2Cu3O7-simalayer. Physica Status Solidi (a). 202(14): R152-154. Rao, Y., Ogitani, S., Kohl, P. and Wong, C. P. (2002). Novel Polymer-Ceramic Nanocomposite Based on High Dielectric Constant Epoxy Formula for Embedded Capacitor Application. Journal of Applied Polymer Science. 83: 1084-1090. Rao, Y., Wong, C. P. (2004). Material Characterization of a High-DielectricConstant Polymer–Ceramic Composite for Embedded Capacitor for RF Applications. Journal of Applied Polymer Science. 92 (4):2228-2231. 146 Saethre, B., Mork, P. C., Ugelstad, J. (1995). Preparation of Poly(VinylChloride) Latexes by Polymerization of Stabilized Monomer Droplets. Journal of Polymer Science. Part A: Polymer. Chemistry. 33: 2951–2959. Scorcik, V., Podgrabinski, T., Nahlik, J., Rybka, V. and Hnatowicz, V. (2005). Dielectric Properties of Doped Polymethylmethacrylate. Materials Letters. 59: 341-244. Sear, F. W., Zemansky, M. W., Young, H. D. (1982). University Physics. 6th Edition, Addison Wesley. Shankar, S. S., Chatterje, S. and Sastry, M. (2003). Synthesis of CdS Nanoparticles within Thermally Evaporated Aerosol OT Thin Films. PhysChemComm. 6(9): 36-39. Smith, B. (1999). Infrared Spectral Interpretation: A Systematic Approach. CRC Press: Boca Raton Steigerwald, M. L. and Brus L. E. (1990). Semiconductor Crystallites: A Class of Large Molecules. Accounts of Chemical Research. 23(6): 183-188. Tang, H., Yan, M., Zhang, H., Xia, M. and Yang, D. (2005). Preparation and Characterization of Water-Soluble CdS Nanocrystals by Surface Modification of Ethylene Diamine. Materials Letters. 59: 1024-1027. Tanwar, A., Gupta, K. K., Singh, P. J. and Vijay, Y. K. (2006). Dielectric Parameters and A.C. Conductivity of Pure and Doped Poly(methyl methacrylate) Films at Microwave Frequencies. Bulletin of Material Science. 29(4): 397-401. Tiarks, F., Landfester, K. and Antonietti, M. (2001a). Preparation of Polymeric Nanocapsules by Miniemulsion Polymerization. Langmuir. 17: 908-918. Tiarks, K., Landfester, K. and Antonietti. M. (2001b). Encapsulation of Carbon Black by Miniemulsion Polymerization. Journal of Polymer Science : Part A : Polymer Chemistry. 39: 2520 – 2524. Tiwari, S. and Tiwari, S. (2006). Electrical and Optical Properties of CdS Nanocrystalline Semoconductors. Crystal Research and Technology. 41(1): 78-82. Tong, Z. and Deng, Y. (2007). Synthesis of Polystyrene Encapsulated Nanosaponite Composite Latex via Miniemulsion Polymerization. Polymer. 48: 4337-4334. Torimoto, T., Reyes, J. P., Iwasaki, K., Pal, B., Shibayama, T., Sugawara, K., Takahashi, H. and Ohtani, B. (2003). Preparation of Novel Silica – Cadmium Sulfide Composite Nanoparticles Having Adjustable Void Space by Size- 147 Selective Photoetching. Journal of American Chemical Society. 125: 316317. Wallace, R. A. (1971). Glass Transition in Partially Sulfonated Polystyrene. Journal of Polymer Science Part B. Polymer Physics. 9:1325-1332. Wang, C.C., Chen, A.L. and Chen I.H. (2005). Effect of Chelating Functional Polymer on The Size of CdS Nanocluster Formation. Journal of Colloid and Interface Science. 293(2): 421-429. Wang, Y. (1991). Nonlinear Optical Properties of Nanometer-Sized Semiconductor Clusters. Accounts of Chemical Research. 24(5): 133-139. Wang, Y. and Herron, N. (1987). Chemical Effects on the Optical Properties of Semiconductor Particles. Journal of Physical Chemistry. 91(19): 5005-5008. Wang, Y. and Herron, N. (1991). Nanometer-Sized Semiconductor Clusters: Materials Synthesis, Quantum Size Effects, and Photophysical Properties. Journal of Physical Chemistry. 95(2): 525-532. Wellmann, H., Rathousky, J., Wark, M., Zukal, A. and Schulz-Ekloff, G. (2001). Formation of CdS Nanoparticles within Functionalized Siliceous. Microporous and Mesoporous Materials. 44-45: 419-425. Wikes, C. E., Summers, J. W., Daniels, C. A. and Berard, M. T. (2005). PVC Handbook. Hanser Verlag. Xu, J. and Wong, C. P. (2007). Characaterization and Properties of an OrganicInorganic Dielectric Nanocomposite for Embedded Decoupling Capacitor Applications. Composites. Part A. 38: 13-19. Xu, W. and Akins, D. L. (2004). Reverse Micellar Synthesis of CdS Nanoparticles and Self-assembly into a Superlattice. Materials Letters. 58(21): 2623-2626. Xu, W., Liao, Y. and Akins, D. L. (2002). Formation of CdS Nanoparticles within Modified MCM-41 and SBA-15. Journal of Physical Chemistry B. 106: 11127-11131. Yanagida, S., Yoshiya, M., Shiragami, T., Pac, C., Mori, H. and Fujita, H. (1990). Semiconductor Photocatalysis. I. Quantitative Photoreduction of Aliphatic Ketones to Alcohols Using Defect-free Zinc Sulfide Quantum Crystallites. Journal of Physical Chemistry. 94(7): 3104-3111. Yang, H., Huang, C., Li, X., Shi, R. and Zhang, K. (2005). Luminescent and Photocatalytic Properties of Cadmium Sulfide Nanoparticles Synthesized via Microwave Irradiation. Materials Chemistry and Physics. 90: 155-158. 148 Yang, S., Sun, K. and Risen, W. M. (2003). Preparation and Thermal Characterization og Glass Transition Temperature of Sulfonated Polystyrene Metal Ionomers. Journal of Polymer Science Part B. Polymer Physics. 28 (10):1685-1697. Yang, T. I. and Kofinas, P. (2007). Dielectric Properties of Polymer Nanoparticle Composites. Polymer. 48: 791-798. Yang, Y., Chen, H. and Bao, X. (2003). Synthesis and Optical Properties of CdS Semiconductor Nanocrystallites Encapsulated in a Poly(ethylene oxide) Matrix. Journal of Crystal Growth. 252: 251-256. Zhang, M., Gao, G., Li, C. Q. and Liu, F. Q. (2004). Titania-Coated Polystyrene Hybrid Microballs Prepared with Miniemulsion Polymerization. Langmuir. 20: 1420-1424. Zhang, Q. M., Li, H. F., Poh, M., Xia, F., Cheng, Z. Y., Xu, H. S. and Huang, C. (2002). An All-organic Composite Actuator Material with a High Dielectric Constant. Nature. 419 : 284. Zhang, Y. H., Dang, Z. H., Fu, S. Y., Xin, J. H., Deng, J. G., Wu, J., Yang, S., Li, L. F. and Yan, Q. (2005). Dielectric and Dynamic Mechanical Properties of Polyimide-Clay Nanocomposite Films. Chemical Physics Letters. 401: 553 – 557. Zhang, Z., Dai, S., Fan, X., Blom, D. A., Pennycook, S. J. and Wei, Y. (2001). Controlled Synthesis of CdS Nanoparticles inside Ordered Mesoporous Silica Using Ion-Exchange Reaction. Journal of Physical Chemistry B. 105(29): 6755-6758. Zhao, H., Douglas, E. P., Harrison, B. S. and Schanze. K. S. (2001). Preparation of CdS Nanoparticles in Salt-Induced Block Copolymer Micelles. Langmuir. 17: 8428-8433. Zhao, P. Q., Wu, X. L., Fan, J.Y., Chu, P. K., and Siu, G.G. (2006). Enhanced and Tunable Blue Luminescence from CdS Nanocrystal-Polymer Composites. Scripta Materialia. 55(12): 1123-1126. Zhou, S. M. (2003). Dielectric Properties of Phase-Size-Control CdS Nanoparticles and Conventional Powder. Physica Status Solidi (a). 200(2): 423-428. Zhu, L., Zheng, X., Liu, X., Zhang, X. and Xie, Y. (2004). In-situ Vesicle-TemplateInterface Reaction to Self-Encapsulated Microsphere CdS. Journal of Colloids and Interface Science. 273: 155-159. 149 APPENDIX A Comparison of stabilization time of CdS nanoparticles in miniemulsion at Wo = 5 with different Po values Po value Stability of the CdS in miniemulsion 16 Precipitate out after < 1 26 Precipitate out after 2 min 39 Precipitate out after 3 min 53 Precipitate out after 10 min 65 Some portion precipitate after 60 min 78 Some portion precipitate after 10 min 105 Some portion precipitate after 10 min 150 APPENDIX B UV – Vis spectra of CdS nanoparticles in miniemulsion at Wo = 5, Po = 65 and concentration of cadmium and sulfide ion of 3.0 x 10-4 M at different time after Absorbance ( a.u) formation t = 24 h t = 5 min - 60 min 385 450 500 550 Wavelength (nm) 600 650 700 151 APPENDIX C UV - Vis spectra of CdS nanoparticles in miniemulsion system at Po = 65 with Wo values = 5 and 10 at different time after formation t = 5 min, Wo = 10 Absorbance ( a.u) t = 5 min, W o = 5 t = 60 min Wo = 5 t = 60 min Wo = = 10 385 400 500 550 600 Wavelength (nm) 650 700 152 APPENDIX D UV - Vis spectra of CdS nanoparticles in miniemulsion at Wo = 5 and Po = 65 at different concentration of cadmium and sulfide ion [Cd2+] = 6 x 10-4 M, t = 5 min Absorbance ( a.u) (a) [Cd2+] = 6 x 10-4 M, t = 60 min [Cd2+] = 3 x 10-4 M, t = 5 min and 60 min (b) 385 450 500 550 600 Wavelength (nm) 650 700 153 APPENDIX E UV – Vis spectra of CdS nanoparticles in miniemulsion system before and after addition of styrene-divinyl benzene at different times after formation, (a) and (d) without monomer at 5 min and 24 h, respectively, (b) and (c) after mixing with monomer at 5 min and 24 h, respectively Absorbance (a.u) (a) (b) (c) (d) 385 450 500 550 600 Wavelength (nm) 650 700 154 APPENDIX F Quantitative standard calibration plot of cadmium element by using Perkin Elmer AA400 AAS 1.2 1 Absorbance 0.8 0.6 y = 0.9992x 0.4 2 R = 0.9977 0.2 0 0 0.2 0.4 0.6 Concentration of Cd (mg/L) 0.8 1 155 APPENDIX G Quantitative standard calibration plot of sodium element by using Perkin Elmer AA400 AAS 0.6 0.5 Absorbance 0.4 0.3 0.2 y = 1.0056x 2 R = 0.9967 0.1 0 0 0.1 0.2 0.3 Concentration of Na (mg/L) 0.4 0.5 156 APPENDIX H LIST OF PUBLICATIONS Conference and National Publications: Publications : 1. Eriawan Rismana, Salasiah Endud, Hadi Nur. (2005). Synthesis of CdS nanoparticles in HDTMAB/2-Propanol/Water/n-Decane Miniemulsion System. Buletin Kimia, 21 - 55-65. 2. Eriawan Rismana, Salasiah Endud and Hadi Nur. Synthesis of Different Sized Cadmium Sulfide Nanoparticles Inside Polymer and Mesoporous Al-MCM-41 by Direct Miniemulsion Polymerization and Ion Exchange Techniques”, Proceedings Annual Fundamental Science Seminar 2006 (AFSS 2006) Institut Ibnu Sina, 6-7 June 2006. Conferences : 1. Eriawan Rismana, Salasiah Endud, Hadi Nur. (2005). Synthesis of CdS nanoparticles in HDTMAB/2-Propanol/Water/n-Decane Miniemulsion System. Annual Fundamental Science Seminar 2005 (AFSS 2005) Institut Ibnu Sina, June 2005. 2. Hadi Nur, Eriawan Rismana and Salasiah Endud. (2008). Nanosize Effect of Cadmium Sulfide dimethacrylate) on Attached on Dielectrical Poly(methacrylic Property. acid-ethylene International glycol Conference on Nanoscience and Nanotechnology (ICONN 2008). 25 – 29 February 2008. Melbourne Exhibition and Convention Centre, Victoria, Australia.