. 53:? , .1, 9307 lllllllllllllllllllllllllllllllllllllllllllllllllllllllllll 31293 01823 2094 This is to certify that the dissertation entitled INTERCALATED P( )LYMER-LAYERED INORGANIC NANOCOMPOSITES presented by Lei Wang has been accepted towards fulfillment of the requirements for Ph. D , Chemistry degree In A; Major professor MSW] fl MSU is an Affirmative Action/Equal Opportunity Institution 0- 12771 LEBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE MAW 1/98 c:/ClRC/DateDue.p65—p.14 INTERCALATED POLYMER-LAYERED INORGANIC NAN OCOMPOSITES By Lei Wang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1999 ABSTRACT INTERCALATED POLYMER-LAYERED INORGANIC NAN OCOMPOSITES By Lei Wang The exfoliation-encapsulative precipitation (EEP) method, which was introduced by Kanatzidis and Bissessur in 1993, is both convenient and general. It enables the formation of an intercalated polymer-layered inorganic nanocomposite when a polymer solution and an inorganic monolayer suspension are mixed. Many MoS2/polymer nanocomposites were synthesized. In this dissertation, we expand the investigation of the M082 system, contribute a better understanding and provide a better manipulation of the reactions. Moreover, we further develop the EEP method by extending its application to other layered materials such as TaS2, MoO3, oc-RuClg and W82, and by modifying the method so that the intercalation of intractable polymers such as polypyrrole is possible. In the research, it was found that “LiMoS2” would not be oxidized completely to M082 upon exposure to water or air, and the exfoliation of “LiMoS2” in water is a process of solvation of Lix(H2O)yM082 (x ~ 0.18). Like Ax(H2O)yTiS2 (A = Li, Na, K), Lix(H2O)yMoS2 can be exfoliated by stirring. Additional polymers such as polyacrylamide, polyvinylalcohol, polypropylene, polystyrene, polymethylmethacrylate, polybenzimidazole, polyethyleneterephthalate and polypyrrole were encapsulated in M082. A convenient method to prepare “LiMoS2” was developed, which uses LiBH4 as a reagent and runs at elevated temperatures (300-350 0C). This method can also be used to lithiate other layered transition metal dichalcogenides. The lithium intercalated materials exfoliate well and encapsulate polymers better than the sodium intercalated materials, at least in the systems of 2H-Ta82 and M003. At room temperature, LiBH4 was proven a good lithiation reagent for 2H—TaS2, M003 and oc-RuClg. The amount of LiBH4 used in the reaction is crucial to the exfoliation and encapsulation . properties of the products. Water-soluble polymers such as poly(ethylene oxide) (PEO) and polyvinylpyrrolidone were intercalated in all three layered materials and polypyrrole was intercalated in RuCl3. In addition, oc-RuCl3/polyaniline nanocomposites were prepared by the in situ redox intercalative polymerization method, which up to now was only applicable to FeOCl, V205 and VOPO4. The reaction is driven by the reduction of some of the Ru3+ centers to Ru2+ and completed by the participation of ambient oxygen as an electron acceptor. The large variety of new polymer-nanocomposites thus prepared have many interesting physical properties. For example, TaS2 nanocomposites are superconductors at low temperatures; RuC13 nanocomposites have adjustable magnetic properties. The arrangement of PEO chains in the interlayer galleries was explored and we believe that M082/PE0, TaS2/PE0 and RuCl3/PE0 nanocomposites contain in each gallery two sheets of PEO chains in the type II PEO-HgCl2 complex conformation. The two sheets of PEO chains are believed to have -0- atoms facing each other at the center of the gallery and provide Li+ ions a good two-dimensional migration channel. To My Family, Who Have Believed Faithfully in and Dedicated Heartily to Science for Generations. iv ACKNOWLEDGMENTS I gratefully acknowledge the people who have helped me in all these years to fulfill the requirements for the Ph. D. degree. It would have been impossible to achieve such a high goal without the love and support from many people who are listed and not listed below. First of all, I would like to thank my advisor, Prof. Mercouri Kanatzidis, for his guidance, encouragement and financial support. I have. benefited substantially from his wise choosing of a productive research project for me, his requirement of a serious and aggressive research manner, and his priceless advice in many aspects. I would like to express my great appreciation for Prof. John Allison for my admission to the Ph. D. program in the Department of Chemistry, which provided me the opportunity to pursue this degree. I would like to thank Prof. Thomas Pinnavaia, Prof. Gregory Baker and Prof. Marcos Dantus, who serve as my guidance committee members and lead me successfully to the completion of this degree. I also thank Prof. Pinnavaia and Prof. Baker for their generous permission for letting me use their instruments and facilities. I enjoyed being a member in the Kanatzidis’ group, in which one is always given assistance and offered friendship. I am in gratitude to the day- to-day help from every member of the group, especially from Joy Heising, Rhonda Patschke, Jason Hanko, Jennifer Aitken, Andy Axtell, Chenggang Wang, Rabin Bisserssur, Dr. Xianzhong Chen and Dr. Francois Bonhomme. I also thank other people in the department for their help, especially the people in Dr. Baker’s group, Dr. Pinnavaia’s group and Dr. Smith’s group. I thank Dr. Rui Huang for many training and measurements, Kermit Johnson for help with the solid state NMR measurements, Dr. Thomas Atkinson for the help with computers, Dr. John Heckman and Dr. Stanley Flegler in the Center for Electron Optics for their help in using TEM and SEM, and Dr. Reza Loloee in the Department of Physics and Astronomy for his help in using the SQUID. I also thank Prof. Kannewurf and his group at Northwestern University for the measurements of electrical conductivity and thermopower and Prof. Jin-Ho Choy at Seoul National University in Korea for the measurements of ion conductivity. I thank the Department of Chemistry for guaranteeing assistantships for five years. Thanks also for the research funding from the National Science Foundation, and the Center for Fundamental Materials Research and the Center for Sensor Materials at MSU. Finally, thanks to my wife and family for their deepest love and reserveless support. vi TABLE OF CONTENTS List of Tables ................................................................................ xiii List of Figures ................................................................................. xv List of Abbreviations ........................................................................ xx Introduction ..................................................................................... 1 l. Intercalation Reactions and Layered Host Materials ...................... 1 2. The Development of Polymer Intercalation ................................. 7 3. Exfoliation-Encapsulative Precipitation: An Effective Method for Polymer Intercalation ............................................. 11 4. Applications of Intercalation Compounds and Nanocomposites ..... 14 5. Important Contributions of the Present Dissertation Work ........... 18 References .................................................................................. 22 Chapter 1. Further Exploration of Exfoliated M082 and Synthesis of New M082 Nanocomposites with the Exfoliation-Encapsulative Precipitation Method ........................................................................ 32 Introduction ................................................................................ 32 Experimental Section ................................................................... 35 1. Reagents ........................................................................... 35 2. Lithiation and Exfoliation of M082 ..................................... 35 3. Investigation of LiMoS2 and Freshly Restacked M082 ........... 38 4. Preparation of M082/Polymer Nanocomposites .................... 39 5. Instrumentation ................................................................. 44 vii Results and Discussion .................................................................. 46 1. The LiBH4 Method of Producing LiMoS2 ............................ 46 2. Manipulation and Properties of Exfoliated and Restacked M082 ................................................................ 49 3. (PEO)XMoS2 and (PVP)xMoS2 Nanocomposites .................... 51 4. Investigations in (PA6)XM082 Nanocomposites ..................... 55 5. Synthesis of (PS)XM082 Nanocomposites .............................. 59 6. Synthesis of Other M082/Polymer Nanocomposites ............... 61 7. The Phase Transition in Restacked M082 and M082 Nanocomposites ............................................................... 64 8. Thermal Stability of M082 Nanocomposites ......................... 67 Concluding Remarks .................................................................... 70 Appendix A ................................................................................ 72 References .................................................................................. 75 Chapter 2. Insertion of Polypyrrole and Poly(N-methyl pyrrole) in M082 and W82 by anin situ Polymerization—Encapsulative Precipitation Method ........................................................................ 78 Introduction ................................................................................ 78 Experimental Section ................................................................... 80 1. Reagents ........................................................................... 8O 2. Preparation of M082/PPY Nanocomposites .......................... 80 3. Preparation of WS2/PPY Nanocomposites ............................ 83 4. Instrumentation ................................................................. 84 viii Results and Discussion .................................................................. 85 1. M082/PPY Nanocomposites ................................................ 85 2. W82/PPY Nanocomposites ................................................. 92 3. An Analysis of Polymer Arrangements inside the Interlayer Galleries Based on the Dimensions of the Polymer Molecules ...................................................... 94 4. Thermal Properties ........................................................... 96 5. Charge Transport Properties ............................................ 100 Concluding Remarks .................................................................. 104 Appendix B ............................................................................... 106 References ................................................................................ l 1 1 Chapter 3. Lamellar TaS2/Polymer Nanocomposites through Encapsulative Precipitation of Exfoliated Layers ............................... 116 Introduction .............................................................................. l 16 Experimental Section ................................................................. 1 18 1. Reagents ......................................................................... l 18 2. Synthesis of 2H-Ta82 ....................................................... 119 3. Synthesis and Exfoliation of LixTaS2 ................................. 119 4. Encapsulative Precipitation of Polymers ............................ 120 5. Instrumentation and Measurements ................................... 121 Results and Discussion ................................................................ 125 1. Exfoliation Properties of LixTaS2 ..................................... 125 2. Polymer Encapsulation .................................................... 126 ix 3. Characterization of LixTaS2/Polymer Nanocomposites ........ 127 4. Structural Studies: The Conformation of PEO in Lix(PEO)yTaS2 ............................................................... 132 5. Superconductive State ...................................................... 137 6. Electrical Transport Properties ........................................ 143 7. Solid State NMR Spectroscopy .......................................... 146 Concluding Remarks .................................................................. 153 References ................................................................................ 155 Chapter 4. Lamellar LIxMoO3/Polymer N anocomposites Via Encapsulative Insertion ............................................................. 159 Introduction .............................................................................. 159 Experimental Section ................................................................. 160 1. Reagents ......................................................................... 160 2. Synthesis of LixMoO3 (0.30 < x < 0.40) ............................ 161 3. Preparation of LixMoO3/Polymer Nanocomposites ............. 162 4. Instrumentation ............................................................... 162 Results and Discussion ................................................................ 163 1. Synthesis and Characterization of LixMoO3 ........................ 163 2. Exfoliation and Polymer Encapsulation Chemistry ............. 165 3. Solid State NMR Spectroscopy .......................................... 176 4. Magnetism ...................................................................... 178 5. Electrical Conductivity .................................................... 181 Concluding Remarks .................................................................. 181 References ................................................................................ 183 Chapter 5. oc-RuClg/Polymer Nanocomposites: the First Group of Intercalative Nanocomposites with Transition-Metal-Halides ............... 186 Introduction .............................................................................. 1 86 Experimental Section ................................................................. 189 1 . Reagents ......................................................................... l 89 2. Reactions and Sample Preparations ................................... 189 3. Instrumentation ............................................................... 194 Results and Discussion ................................................................ 194 1. Preparation of (PANDXRuClg, by in situ Redox Intercalative Polymerization ............................................................... 194 2. Nanocomposites of oc-RuClg with Water Soluble Polymers .. 202 3. oc-RuClg/Polypyrrole Nanocomposites ............................... 204 4. Charge Transport Properties ............................................ 206 5. Magnetic Susceptibility Studies ......................................... 209 6. One-Dimensional Electron Density Calculation and Arrangement of Polymer Chains in Lix(PE0)yRuCl3 ......... 217 7. PEO Conformation from IR Spectra ................................. 226 Concluding Remarks .................................................................. 229 References ................................................................................ 23 1 xi LIST OF TABLES Table 1. Elements forming layered transition metal dichalcogenides ....... 2 Table 1.1. Nylon-6 intercalation reactions with PA6/CF3CH20H solution ........................................................................................... 59 Table 1.2. Information about the synthesis of M082/PS nanocomposites ................................................................................ 60 Table 1.3. Characteristics of some new M082 nanocomposites ............... 63 Table 1.4. Evolution rates of restacked M082 and nanocomposites derived from the change of electrical resistivity .................................. 67 Table 1.5. The effect of high temperature annealing to some nanocomposites ................................................................................ 68 Table 2.1. Parameters of (PPY)xMoS2 prepared with different pyrrole ratios .................................................................................. 88 Table 2.2. Structural data of PPY and PMPY nanocomposites .............. 94 Table 2.3. Room temperature electrical conductivities of PPY nanocomposites and restacked M082 and WS2 .................................... 100 Table 3.1. The annealing procedure to produce 2H-TaS2 .................... 119 Table 3.2. Polymer Intercalation in LixTaS2 ..................................... 127 Table 3.3. Properties of Li02TaS2 and polymer nanocomposites ......... 129 Table 3.4. Properties of the superconductive state for TaS2 intercalates .................................................................................... 138 Table 3.5. Effect of film orientation on the 7Li-NMR spectrum 0f Li0,2(PEO)yTaS2 ........................................................................ 153 Table 4.1. Chemical and structural characteristics of LixM003/polymer nanocomposites ................................................... 166 xii Table 4.2. Composition and physicochemical properties of the LixM003/polymer nanocomposites .............................................. 175 Table 5.1. Magnetic properties of a-RuCl3 and nanocomposites .......... 213 Table 5.2. A comparison of IR absorptions (cm-1) of PEO ................. 228 xiii LIST OF FIGURES Figure 1.1. Experimental set-up for the high temperature LiBH4 lithiation reaction ............................................................................. 37 Figure 1.2. XRD patterns of a (PEO)XMoS2 nanocomposite used to calculate the one—dimensional electron density map .............................. 53 Figure 1.3. One-dimensional electron density map of a (PEO)xMoS2 nanocomposite ................................................................................. 53 Figure 1.4. Structural model for the (PE0)xMoS2 nanocomposite ........ 54 Figure 1.5 . XRD patterns of the new nanocomposite phases ................. 56 Figure 1.6. Evolution of the electrical resistivity of a restacked M082 sample as a function of time .............................................................. 65 Figure 1.7. XRD patterns of the (PE0)xMoS2 and (PP)XM082 nanocomposites before and after the high temperature annealing ........... 68 Figure 1.8. Intensity of the 001 peak in the XRD patterns of a (PE0)xMoS2 sample at increasing temperatures ................................ 69 Figure 1.9. Curve fitting for the time dependence of electrical conductivity of (PE0)xMoS2 ............................................................. 74 Figure 2.1. XRD patterns of M082 nanocomposites .............................. 86 Figure 2.2. IR spectra of PPY and a M082/PPY nanocomposite prepared by procedure 1 ................................................................... 87 Figure 2.3. IR spectrum of a M082/PPY nanocomposite prepared by procedure 1 and heating treated twice up to 300 0C under N2 atmosphere ...................................................................................... 87 Figure 2.4. IR spectra of a nanocomposite (PPY)0,51(H20)0,13M082 prepared by procedure 2 ................................................................... 90 Figure 2.5. XRD patterns of WS2/PPY nanocomposites ........................ 93 xiv Figure 2.6. IR spectrum of (PPY)0,25(H2O)0_53WS2 heating treated twice up to 320 0C under N2 atmosphere ............................................ 93 Figure 2.7. DSC measurements at different heating rates for (PPY)0_51(H20)0,18MOSZ .................................................................... 97 Figure 2.8. Plot of the DSC heating rate versus the temperature of the M082 structural transition for (PPY)0,51(H2O)0,18M082 ............... 97 Figure 2.9. DSC measurement for (PPY)0,25(H20)0_53WS2 .................... 98 Figure 2.10. Plot of the DSC heating rate versus the W82 structural transition temperature in (PPY)0,25(H20)0,53WS2 ................................. 98 Figure 2.11. Variable temperature electrical conductivity measurements on pressed pellets for pristine M082 and a M082/PPY nanocomposite prepared with Procedure 1 .............................................................. 102 Figure 2.12. Thermoelectric power measurement on a pressed pellet for (PPY)0_25(H2O)0_16M082 prepared with Procedure 1 ..................... 102 Figure 2.13. Variable temperature electrical conductivity measurements on pressed pellets for restacked W82 and (PPY)0,25(H20)0,53WS2 ........ 103 Figure 2.14. Thermoelectric power measurement on a pressed (PPY)0.25(H20)0.53WSZ pellet ........................................................... 103 Figure 2.15. Schematic structures and dimensions of PPY and PMPY . 109 Figure 2.16. Views of PMPY ......................................................... 110 Figure 3.1. XRD patterns of LixTaS2/polymer nanocomposites .......... 128 Figure 3.2. DSC of a Li0,2(PEO)yTaS2 nanocomposite ....................... 130 Figure 3.3. SEM photograph of a Li0.2(PEO)yTaS2 film ..................... 131 Figure 3.4. Powder XRD patterns of a folded Li0,2(PEO)yTaS2 film 133 Figure 3.5. One-dimensional electron density maps for the Lix(PEO)yTaS2 nanocomposite ................................................... 134 XV Figure 3.6. Structural model for the Li0,2(PEO)yTaS2 nanocomposite .. 135 Figure 3.7. Variable temperature magnetic susceptibility for Li02TaS2 and Li0,2(PE0)yTaS2 powders ........................................... 140 Figure 3.8. Variable temperature magnetic susceptibility for Lio,2(PE0)yTaS2 films .................................................................... 141 Figure 3.9. Variable temperature electrical conductivity measurements for pressed pellets of Li0,2(PE0)yTaS2 ............................................. 144 Figure 3.10. Variable temperature thermopower data of pressed pellets of Li0,2(PE0)yTaS2 ........................................................................ 144 Figure 3.11. Electron diffraction pattern and superstructure of Li0,2TaS2 ....................................................................................... 145 Figure 3.12. Static solid state 7Li NMR spectra of LiOI2TaS2 and 1410203130) yTaS 2 ............................................................................ 148 Figure 3.13. Temperature dependence of the linewidth of 7Li NMR resonance peaks for Li0,2TaS2 and Li0,2(PEO)yTaS2 ........................... 149 Figure 3.14. Room temperature 7Li NMR spectra for a Li0,2(PE0)yTaS2 film ................................................................... 152 Figure 4.1. DSC diagram of LixMoO3 .............................................. 164 Figure 4.2. IR spectra of Lix(H2O)y(PVP)zM003 ............................... 167 Figure 4.3. Solid-state optical absorption spectra of the LixM003/polymer nanocomposites ................................................... 168 Figure 4.4. Typical XRD patterns of nanocomposites with poly(ethylene glycol) and poly(ethylene oxide) of different molecular mass ............. 169 Figure 4.5. Typical XRD patterns of the various LiXM003/polymer nanocomposites .............................................................................. 170 Figure 4.6. XRD patterns of Lix(H2O)y(PEO-2000)ZM003 showing the effect of annealing on the stacking regularity of the layered structure of a nanocomposite ........................................................... 174 xvi Figure 4.7. Static 7Li NMR spectra of LixM003 and Lix(H2O)y(PE0)zM003 at -80 0C ..................................................... 177 Figure 4.8. Temperature dependence of the linewidth of the resonance peak in solid-state 7Li MNR spectra of LixMoO3 and Lix(H20)y(PE0)zM003 ................................................................... 177 Figure 4.9. Temperature dependence of magnetic susceptibility of LixM003 and Lix(H20)y(PE0)zM003 ............................................... 179 Figure 4.10. Variable-temperature electrical conductivity measurements for LixM003 and Lix(H2O)y(PE0)ZM003 .......................................... 180 Figure 5.1. The structure of oc-RuClg ............................................... 188 Figure 5.2 Experimental set-up for the synthesis of OL-RuC13 .............. 190 Figure 5.3. X—ray diffraction patterns of oc-RuCl3 and nanocomposites 196 Figure 5.4. Transmission X-ray diffraction patterns of OL—RuC13 and (PANI)xRuCl3 with indexing ............................................................ 197 Figure 5.5(a). Infrared spectra of (PANDXRuClg, PANI and aniline 198 Figure 5.5(b). Infrared spectra of (PPY)xRuCl3 and PPY .................... 199 Figure 5.5(c). Infrared spectra of Lix(PEO)yRuCl3 and PEO .............. 200 Figure 5.5(d). Infrared spectra of Lix(PVP)yRuC13 and PVP .............. 201 Figure 5.6. Electron diffraction patterns with the electron-beam perpendicular to the planess for oc-RuC13 and LixRuCl3 ...................... 203 Figure 5.7. Variable temperature electrical conductivity mesurements for pressed pellets of oc-RuClg and nanocomposites ............................ 207 Figure 5.8. Thermopower measurements for pressed pellets of LixRuCl3 (x ~ 0.2) and nanocomposites ............................................ 208 Figure 5.9(a). Magnetic susceptibility measurements for a sample of 0L-RuCl3 ........................................................................................ 21 1 xvii Figure 5 .9(b). Magnetic susceptibility measurements for LixRuC13 ...... 211 Figure 5.9(c). Magnetic susceptibility measurements for Lix(PE0)yRuC13 ............................................................................. 212 Figure 5.9(d). Magnetic susceptibility measurements for (PANDXRuClg ................................................................................ 212 Figure 5.10(a). Magnetic susceptibility measurements for “Brz oxidized LixRuClg” (Le. RuCl3) .................................................................... 214 Figure 5.10(b). Magnetic susceptibility measurements for “Brz oxidized Lix(PEO)yRuCl3” (Le. “(PEO)xRuCl3”) ............................................ 214 Figure 5.10(c). Magnetic hysteresis measurements for “Brz oxidized LixRuCl3” (Le. RuC13) .................................................................... 215 Figure 5.10(d). Magnetic hysteresis measurements for “Brz oxidized Lix(PEO)yRuC13” (Le. “(PEO)XRuCl3”) ............................................ 215 Figure 5.11. Possible PEO conformations ......................................... 219 Figure 5.12. X—ray powder diffraction pattern of oriented Lix(PEO)yRuCl3 film used for the calculation of one—dimensional electron density maps ..................................................................... 220 Figure 5.13. One-dimensional electron density maps for Lix(PEO)yTaS2 and Agx(PE0)yTaS2 .......................................... 220-222 Figure 5.14. Structural model for Lix(PEO)yRuCl3 ........................... 223 xviii LIST OF ABBREVIATIONS l-D ED ................................................. one-dimensional electron density 1T-MoS2 ............................................................. octahedral form M082 2H— M082 .............................................................. hexagonal form M082 AF ........................................................................... antiferromagnetism xdia .................................................................. diamagnetic susceptibility cm ........................................................................................ centimeter xmolar ...................................................................... molar susceptibility xpam ............................................................... paramagnetic susceptibility xTIP .................... susceptibility of temperature-independent paramagnetism DSC ...................................................... differential scanning calorimeter Ea ............................................................................... activation energy EDS .............................................. energy dispersive X-ray microanalysis EEP ............................................... exfoliation-encapsulative precipitation FM ...................................................................................... formamide FTIR ......................................... Fourier transform infrared spectroscopy g .................................................................................................. gram GR ................ Guaranteed Reagent (suitable for uses in analytic chemistry) h .................................................................................................. hour i-PrOH ................................................................................ isopropynol ICP ................................................................ inductively coupled plasma IR ......................................................................... infrared spectroscopy kJ ............................................................................................ kilojoule M ............................................................................................ molarity 1.13 ................................................................................. Bohr magneton Mcel ............................................................................. methyl cellulose ueff ................................................................ effective magnetic moment min ............................................................................................ minute m1 .......................................................................................... milliliter mmol ..................................................................................... millimole mol ............................................................................................... mole MW ............................................................................ molecular weight NMF ....................................................................... N-methylformamide NMR ............................................................ nuclear magnetic resonance PA6 .......................................................................................... nylon-6 PAM .............................................................................. polyacrylamide PANI ................................................................................... polyaniline PBI ............................................................................ polybenzimidazole PE ..................................................................................... polyethylene PEG ...................................................................... poly(ethylene glycol) xix PEI .............................................................................. polyethylenimine PEO ....................................................................... poly(ethylene oxide) PETP ........................................................... poly(ethylene terephthalate) PLD-CDW ......................... periodic lattice distortion-charge density wave PMMA ........................................................... poly(methyl methacrylate) PMPY ................................................................ poly(N-methyl pyrrole) PP ................................................................................... polypropylene PPG .................................................................... poly(propylene glycol) PPY .................................................................................... polypyrrole PS ....................................................................................... polystyrene PVA ......................................................................... poly(vinyl alcohol) PVP ............................................................... poly(N-vinyl pyrrolidone) 0 (in magnetism) .............................................................. Weiss constant 0 (in XRD) .......................................................................... Bragg angle s ................................................................................................. second SEM .......................................................... scanning electron microscopy Tc ....................................... critical temperature of superconducting state TEM ..................................................... transmission electron microscope TGA ........................................................... thermal gravimetric analysis UV-VIS-NIR ..................... ultraviolet-visible-(near infrared) spectroscopy XRD .......................................................................... X-ray diffraction XX INTRODUCTION 1. Intercalation Reactions and Layered Host Materials In chemistry, the term "intercalation" has been used to describe reactions in which guest atoms, molecules or ions are inserted into layered host lattices while the host lattices reserve their essential structural features. Literally, intercalation chemistry dates from 1840 when Schafhautl inserted sulfate ion in graphite [1]. Nevertheless, the subject had not drawn the attention of chemists, physicists and scientists in other fields until 1926 when Fredenhagen and Cadenbach discovered the insertion of potassium atoms from the vapor into graphite [2]. In the 19603, research in this area grew significantly and extended into many scientific disciplines. Most of the work has focused on the intercalation phenomenon and the effect of intercalation on the physical properties of the host materials [3]. An important aspect of this research is the alkali metal intercalation in transition metal dichalcogenides, which modifies both the lattice structures and the electronic properties of the hosts [4]. The discovery in the late 19603 by Gamble et al [5], which showed that the critical temperature (Tc) for the onset of superconductivity in 2H—Ta82 could shift from 0.8 K to about 7 K upon amine intercalation, was one of the most exciting events in intercalation chemistry. An eye-catching research at present is the intercalation of inorganic or organic compounds in the high—Tc superconductors to adjust the properties and explore the mechanism of the superconductivity [6]. In recent years, with the accumulation of a substantial amount of knowledge about intercalation materials and the improvement of the synthetic skills, the aim is set to explore the use of intercalation reactions to synthesize novel and potentially applicable materials [7]. The investigation of inorganic/polymeric intercalation compounds, or ceramic/polymeric nanocomposites, is accelerated by this aim, and bursts into a major branch in the field of nanophase—materials [8’ 9]. The materials which are intensively studied by intercalation are members of the seven most common groups of layered materials, including graphite, layered transition metal dichalcogenides, layered transition metal oxides, layered metal oxyhalides, layered metal phosphorus trichalcogenides, layered acid salts of tetravalent metals, and sheet silicatesllOl. Graphite is the simplest host material and it shows diverse intercalation chemistry due to its stability to both oxidation and reduction. The known intercalating reagents range from strong oxidants to powerful reductants [11]. Table 1. Elements forming layered transition metal dichalcogenides Transition Metal Chalcogen IV V VI VII VIII VI 8 Ti (V) (Cr) Se Zr Nb Mo Pd Te Hf Ta W Re Pt Other layered materials have been explored since the beginning of the sixties. Layered transition metal dichalcogenides form a large group, since most of the transition metals from group IV, V, VI and some from group VII and VIII can form dichalcogenides of layered structures (Table 1). The intercalation chemistry of the layered dichalcogenides is mainly characterized by the reduction of the host lattices. Simple and hydrated cations, and organic and organometallic ions can be included between the layers of the lattices through electron donating processes. The layered dichalcogenides have many unusual physical properties due to their electronic structures (partial d-orbital filling) as well as their low dimensionality, and these properties are adjustable by intercalation [4]. Similar to layered dichalcogenides are the layered oxides, which exhibit quite the same intercalation chemistry. The best known of these are molybdenum oxide, tungsten oxide and vanadium oxide, which form various bronzes with alkali metals [12]. Layered oxyhalides MOCl ( M: Fe, Ti, V and Cr) [13] and InOX ( X: Cl, Br and 1) [10b] exhibit some interesting intercalation chemistry. Metal phosphorus trisulfides of the formula MPS3 form a broad and diverse class of compounds. The metal M includes Mg, Ca, V, Mn, Fe, Co, Ni, Zn, Pd, Cd, Hg, In, Sn, and Pb. Metal phosphorus triselenides of the similar layered structure also exist. Examples are FePSe3, MgPSeg, MnPSe3, CdPSe3, SnPSe3, PbPSe3 and HgPSeg. Like transition metal dichalcogenides, metal phosphorus trisulfides react with electron donors and their intercalation chemistry is mostly confined to 3 major classes of guests: alkali metals, organic amines and organometallic molecules. Ion exchange reactions with intralayer M2+ ions and non-redox intercalation reactions have also been developed recently [14]. Layered silicates and layered acid salts are currently of considerable interest, because of technological applications in heterogeneous catalysis, and as sorbents and inorganic ion—exchangers. Their intercalation chemistry is predominantly of reactions with cations by ion-exchange, or with neutral molecules [15]. Layered silicates include many naturally existing and synthetic clay minerals [16]. Layered acid salts of tetravalent metals have the general formula MIV(HXO4)2.nH2O (X=P, As). The most extensively investigated species is oc-zirconium bis(monohydrogen orthophosphate) [oc—Zr(HPO4)2.H20] [17]. The protons of these acid salts can be easily replaced by other cations, so the layered acid salts are considered attractive inorganic ion-exchangers. Also included in layered acid salts are hydrogen uranyl phosphate and arsenate, HUO2PO4.4H2O and HU02ASO4.4H20 [13], while vanadyl and niobyl compounds of the type V0X04 and NbOX04 (X = S, P, As, Mo) are quite similar compounds [19]. Besides the seven large groups of layered materials, some groups of layered inorganic compounds and some individuals that are not so common have also been studied in intercalation. Reduced halides such as ZrCl, ZrBr and ThI2 possess an important property involving reversible hydrogen uptake to form ordered, stoichiometric phases. oc—RuCl3, Ni(CN)2 and B- ZrNCl also undergo intercalation reactions [10b]. HFe(SO4)2.4H2O intercalates monovalent or divalent ions along with the reduction of Fe(III) centers [20]. Lamellar double hydroxides of the general formula [MELx me(OH)2][Xm’x/m.nH2O] (where the X” is a halide anion or an oxo-anion) are designated as "anionic clays", and are of great interest for their capability to anion exchange [21]. [Zn1_xAlx(0H)2][Clx.nH2O] and [Cu1- xCrx(OH)2][Clx.nH20] are examples of them. Basic copper acetate is another anion exchangeable layered compound [7b]. A large series of layered niobates, titano-niobates and titanates with perovskite related structure and with alkali metal cations in interlayer sites have been developed in recent years [15, 22]. Ion-exchange, intercalation of amines, exfoliation and even electrochemical alkali metal insertion can be applied to this group of layered oxides. Recently, intercalation chemistry is also found in the pillaring of buaerite (Na4Mn14026.xH20) [23], the ion-exchange of birnessite (Na0.32Mn02.nH20) [24], and the metal atom or simple molecule insertion in the misfit layered sulfides [25]. The variety of layered inorganic compounds mentioned above provides a broad choice of layered materials. To make these materials constituents of nanocomposites, the chemistry of intercalation reaction is the decisive factor. The common synthetic methods for intercalation include direct insertion, ion-exchange, exfoliation and flocculation, and electro-intercalation [10b]. The intercalation of alkali metals and other metals into graphite, transition metal dichalcogenides, transition metal oxides and oxyhalides, and metal phosphorus trichalcogenides can be realized by direct intercalation. The insertion processes are accompanied by the charge transfer from the intercalants to the host lattices. It is believed that the tendency for this charge transfer is the driving force for the intercalation reaction. Organic molecules can also be intercalated into the same type of hosts through direct intercalation. Only Lewis-base (electron-donating) organic compounds can be intercalated. The nature of the bond between an intercalated molecule and the host layer is inquired. An old model proposed a weak covalent bond involving no charge transfer126l. More recent studies suggest that the molecule-host “bond” is substantially ionic. In the new model, part of the intercalated molecules are ionized by donation of an electron to the layer, whereas the remainder of the molecules are neutral and solvate the ionized species [27]. Ion-exchange can be used to prepare intercalation compounds with hosts of charged layers, such as clays, layered acid salts, and lamellar double hydroxides. When a charged layered compound is immersed in a concentrated solution containing a replacing ion, the interlayer ion can often be replaced by the other ion. The exchange reaction is driven by the great excess of the replacement ion. In the cases of neutral oxidizing layered hosts such as transition metal chalcogenides, once a metal ion has been intercalated it can subsequently be ion-exchanged. In many cases, the approach of pre-intercalating with an alkali metal ion and ion-exchanging the alkali metal cation with the target guest cation provides a useful strategy for the intercalation of large guest cations which do not intercalate directly. Frequently, ion-exchange cannot be achieved in one step because the replacing ion is very large whereas the interlayer gallery and the resident ion are small. Multi-step ion-exchange processes are needed to overcome this difficulty. By replacing the resident ion with suitable ions of intermediate sizes, the gallery of the host can be opened gradually so that finally the target guest ion can be inserted [17a]. The introduction of large target ions by stepwise ion-exchanges clearly demonstrates the fact that the high activation energies associated with the deformation of the layered structures to accommodate the incoming guests can be overcome if the host lattices can be expanded by pre-intercalation of smaller molecules or ions. If this strategy is taken to its extreme and the host layers are completely dispersed or exfoliated, very large molecules or ions can be included. This is the case of intercalation by exfoliation and flocculation. Smectic clays can be swelled in water and then intercalated with large cations such as polyoxometallates [16, 28]. Na0,33TaS2 has been exfoliated in N- methylformamide/H20 solution and intercalated with a large cluster cation [F3688(PEt3)6]2+ [29]- Finally, electro-intercalation provides a convenient way to prepare metal intercalation complexes. Its advantages over conventional techniques are its simplicity, facile control of stoichiometry, and fast rate of reaction at room temperature. In addition, it provides a convenient method to carry out detailed thermodynamic measurements and study the staging phenomenon [10b]. These four well developed intercalation methods are very effective in the simple ion and small molecule intercalation. They also shine light on the strategies for polymer intercalation. However, because the macromolecules are much greater in size and are usually chemically inert, the practice of polymer intercalation is still full of obstacles and advances slowly. The following section will show that the strategies applied to the polymer intercalation are often different from the simple ion and small molecule intercalation. 2. The Development of Polymer Intercalation The synthesis of inorganic/polymeric intercalative compounds is aimed at the materials for potential applications. Intercalation provides a molecular level combination of two extremely different components, which is expected to produce materials with superior or novel properties. Small- molecule intercalation complexes are not a preferred choice, because they intend to de-intercalate so that their compositions are not stable. Stability is a prerequisite for a material in practical use. The polymer intercalants, in contrast, stay in the interlayer galleries of inorganic lattices after the formation of intercalative complexes, which assure steady compositions for the nanocomposites. In addition, the polymers possess mechanical strength. The polymer chains dangling from the layered structure hold the layered materials together, so that the nanocomposites can be produced in various forms, such as films and bulk materials, without additional supporting substances. Both polymers and layered hosts are usually air and moisture stable, so the nanocomposites can be used in ambient conditions without much concern. There exist a large number of layered materials and polymers, so the combinations are numerous. The advance of the intercalative nanocomposites depends on the development of the synthetic methods for these materials, which are evolving from the conventional intercalation methods. In the 1960s and early 19703, various types of ions and small molecules had been intercalated into layered inorganic compounds, while the intercalation of large molecules such as organic cations, large clusters and polymers was still rare, even though the synthesis of inorganic/polymeric intercalative complexes had been carried on since 1965 [30]. This situation existed because the intercalation of large molecules is kinetically unfavorable and the reaction is technically harder. The intercalation of robust cations was realized in later 19703, which led to the intercalated smectite derivatives with pore sizes larger than those of faujasitic zeolites. These intercalated smectite derivatives are known as "pillared clays" [31]. On the other side, the worldwide interest in polymer intercalation increased substantially in 1987 when Okada et a1 developed clay/polyamide nanocompositesl32'33] and Kanatzidis et a1 synthesized inorganic/(conductive polymer) layered complexes [34]. The clay/nylon—6 nanocomposites exhibited extraordinary mechanical and thermal properties, which were remarkably superior to its individual components, and immediately found their use in automobile industry as the materials for timing-belt covers [33» 35]. Other polymers such as rubber [33], poly(e—caprolactone) [36], epoxy resin [37], polyirnide [38] and polystyrene [39] have also been hybridized with clays to make nanocomposites of excellent mechanical and thermal mechanical properties. Encapsulation of conductive polymers inside layered compounds leads to anisotropically conductive materials. In addition, it is expected that the overlapping and interaction of the electronic bands of the guest and the host will bring up unpredicted novel physical properties. The first complex of this type was FeOCl/polypyrrole and, since then, this aspect has always been a hot topic in intercalation. So far, conductive-polymer intercalation compounds with hosts such as FeOCl [34, 40], V205 [41], M003 [47-1, M082 [43], metal phosphates [2013,44], halides (CdX42' and MnX42‘) [45], HFe(SO4)2.4H2O[20bl, as well as clays [33b] have been reported. The conductive polymers include polypyrrole, polyaniline, polythiophene, polyfuran, poly(p-phenylenevinylene) and poly(diacetylene). Recently, it has been shown that some of these electronically active nanocomposites have better electrochemical properties than either the hosts or the polymersl46, 47]. The importance of the outcome of the clay/nylon-6 and FeOCl/polypyrrole nanocomposites to the research of polymer intercalation was not only the demonstration of the applications of the nanocomposites, but also the introduction of the new synthetic methodologies associated with the preparation. The surfactant modified clays used in the preparation of the clay/nylon-6 nanocomposites are good starting materials for a lot of clay/polymer nanocomposites, while the in situ redox intercalative polymerization used to synthesize the FeOCl/polypyrrole nanocomposite suits for a class of layered hosts and monomers. The procedures of the two methods are easy to carry out, producing nanocomposites of well defined morphologies. Before the establishment of these two methods, polymer intercalation was almost exclusively accomplished by monomer intercalation and subsequent intergallery polymerization. This classical method has not contributed many nanocomposites because neither monomer intercalation nor intergallery polymerization could be accomplished readily. The success in preparing many new nanocomposites brought more researchers into this field and gave rise to more efforts to develop new methodologies for polymer intercalation. N azar et a1 made MoO3/poly(phenylene vinylene) nanocomposites through the insertion of the precursor ionomer by ion-exchange and the subsequent conversion to the polymer [423’ b]. Ruiz-Hitzky et al incorporated poly(ethylene oxide) (PEO) in layered silicate [48] and V205 xerogel [49] by treatment of the host with non—aqueous polymer solution. PEO was also intercalated in MPS3 (M=Mn, Cd) by Lagadic et a1 by stirring the hosts with an aqueous solution or a methanol solution of PEO [50]. In our group, PEO and other water- soluble polymers were encapsulated in V205 xerogel through mixing the aqueous solution of the two [51]. All these accomplishments were reported in a short period in early 19903. By 1993, the preparation of M082/polymer nanocomposites by exfoliation-encapsulative precipitation (EEP) was reported by our groupl52], while the preparation of M082/PEO and TiS2/PEO was reported by Ruiz-Hitzky's group [53] simultaneously. The EEP method demonstrated its usefulness by intercalating a large variety of polymers in M082 [52], a host which otherwise is very difficult to intercalate with even small molecules [54]. It has been expected since that the application of the EEP method to other layered materials would bring in many new nanocomposites. This knowledge provided a promising starting point for the present dissertation [4313’ 551, as well as the work of other researchers 156’ 57]. On the other side, Giannelis et a1 developed the melt intercalation method for organoclays in 1995 [58]. This method has a great advantage in large-scale preparations, so the development is very important to the commercialization of the organoclay nanocomposites. 10 3. Exfoliation-Encapsulative Precipitation: An Effective Method for Polymer Intercalation The exfoliation-encapsulative precipitation (EEP) method corresponds to the exfoliation-flocculation method in conventional intercalation. The latter was developed to intercalate robust cations in negatively charged layered materials in late 19703 and early 19803 [28’ 291. In the case of robust cation intercalation, the driving force was well understood. It is the electrostatic attraction between the positively charged cations and negatively charged layers. The net result of the reaction is the ion-exchange of large cluster cations for small interlayer metal cations. In the case of neutral polymer intercalation, however, no driving force had been expected in such a situation. The anionic hosts, which can be exfoliated, are considered extremely polar materials. The neutral polymers, on the other hand, are usually non-polar or slightly polar materials. It is hard to see the affinity between the two classes of materials. Although Morrison et al had encapsulated a list of small organic molecules in M082 by exfoliation-flocculation in 1986 [54], the event seemed not impacting enough to persuade people to generalize this strategy to polymer intercalation. Nazar et al [423, b] chose an ionomer rather than a polymer to intercalate in an exfoliatable host, NaxMoO3. In the work of Ruiz-Hitzky et al on the layered silicates [43] and the work of Lagadic et al on MP83 [50], exfoliation was not applied. On the other hand, in the work of polymer intercalation in V205 xerogel, which was done by our group [51], an exfoliation step was not involved in the synthesis. Although all these efforts enriched the preparation chemistry of inorganic/polymeric nanocomposites, the peculiarities existing in either the polymers, the hosts, or the reactions stopped people from extensively generalizing these methods. 11 The design of the EEP method [52] was based on the expectation that the monolayers of the inorganic compounds would combine with polymer molecules in solution to form nanocomposites. The success of the method relied on the co-precipitation of the monolayers and the polymer from solution. The design of this reaction was somewhat different from that of Ruiz-Hitzky et al for PEO intercalation in M082 and Ti82 [531, in which the exfoliation was utilized to facilitate the polymer insertion. The trial of the EEP method on M082 nanocomposites was a brilliant success. The direct intercalation in M082 with a list of polymers, which included water soluble polymers such as poly(ethylene oxide), poly(propylene oxide), poly(ethylenimine), poly(vinyl pyrrolidone) and methyl cellulose, commonly used polymers such as polyethylene and nylon-6, and conducting polymer polyaniline [52], through a very simple procedure, brought the attention of other researchers in this area. The successful application to a common transition metal dichalcogenide using so many polymers showed the generality of the method. The intercalation of these polymers also indicated the existing affinity between the polymers and exfoliated layered materials, which is a critical factor for the on going research in polymer intercalation. Accordingly, the EEP method was established as a new promising method for polymer intercalation. For the EEP method to apply, the most important issue is the availability of monolayer suspensions of the host materials. Monolayer suspensions have been explored increasingly in the last 20 years as an approach to prepare intercalative compounds with very large guest molecules and high surface areas as catalysts, and to form monolayer coatings and films. They have been reported in clays, layered transition metal disulfides (TiS2, 2H-NbS2, 2H-Ta82, M082 and W82), transition metal oxides (M003), metal oxychlorides (FeOCl 1591), transition metal 12 phosphorus sulfides (MnPS3), metal hydrogen phosphates [Zr(HPO4)2.H2O and (H3O)U02PO4.3H20 (HUP) [601] and niobium oxides (HTiNb05 and HCa2Nb3010). Dispersions of clay minerals have been studied most extensively [61], which led to the synthesis of "pillared clays" and delaminated clay-polymer nanocomposites. Alkali metal or proton intercalated Ti82 [62], 2H—Nb82 [62, 63] and 2H-TaS2 [62, 63v 641 exfoliate when mechanical stirring is applied. Lithiated M082 [54] and M003 [65] form monolayer dispersions with the assistance of ultrasonic vibration. W82 [66’ 67] monolayer suspension has been prepared following M082. Lithium ion exchanged MnPS3 forms a colloidal dispersion in water spontaneously [68]. Acid zirconium phosphates, Zr(HPO4)2.H2O [69], and proton exchanged niobium oxides (HTiNb05 [70] and HCa2Nb3010 [711) exfoliate after the interlayer surface modification with pre-intercalated alkylamines. Further development of the surface modification on zirconium phosphates leads to zirconium organophosphates and phosphonates that can exfoliate spontaneously in water [72]. A review for colloidal dispersion of layered inorganic compounds was written recently by Jacobson [73]. In addition, more reports about exfoliation have come out recently in misfit—layered metal sulfides [74], layered metal phosphates [75] and niobium oxides [76], and even high-Tc superconductors [77]. Our new success in exfoliation of oc- RuC13 is described in Chapter 5. With the increasing interest in colloidal monolayer dispersions, more exfoliated materials will be discovered. Therefore, many host materials exist which might be amenable to the EEP method and this promises a bright future for polymer intercalation and associated nanocomposite materials. The successful application of the EEP method also depends on the affinity between polymers and the dispersed monolayers. This factor is 13 more decided by the nature of the dispersed monolayers rather than the manipulation of the chemistry. Thus, even with the monolayer suspension of a host material, intercalation of polymer is not guaranteed. However, it is possible to change the preparative method or condition to adjust the nature of the monolayer dispersions, so that the monolayers become attractive to the polymers. The new preparation procedures for LixTaS2 and LixMoO3 described in Chapter 3 and 4 are examples of this modification. It is also likely to choose some special conditions of intercalation for some monolayer dispersions so that polymer intercalation is favored. Examples of these reaction designs can be found in Chapter 1. As a result of the factors described above, searching for the best conditions to conduct polymer encapsulation is still inevitable. 4. Applications of Intercalation Compounds and Nanocomposites The study of chemical intercalation is not only interesting but also beneficial from a practical viewpoint. Intercalation offers a mild way to modify the structures and properties of the host lattice through electron transfer, ion-exchange, and guest molecule insertion. It is a unique approach to obtain new materials that cannot be prepared by the conventional techniques of solid state chemistry. Searching for new materials for potential applications has been a hot topic in the field of chemical intercalation for many years, and has become an important driving force for research. The most important application of intercalation materials is as reversible electrodes for rechargeable batteries. The utilization of graphite intercalation compounds as electrode materials in electrochemical generators was initially proposed by Armand in 1973 [78]. In 1977, the idea of intercalative chalcogenide batteries was developed by Whittingham [79]. 14 Several ambient temperature secondary lithium battery systems such as Li— TiS2, Li-MoS2, Li—NbSe3, Li-MoO2, Li—V205 etc. have been investigated at various battery companies and R&D organizations for consumer and defense purposes [80]. At present, the research on positive electrode materials for secondary lithium batteries is mainly focused on LiCoO2 and LiMnO2 [31]. On the other hand, the application or potential application of intercalation materials as catalysts, ion-exchangers and sorbents is well known [82, 33]. Pillared and delaminated clays are studied as shape selective, heterogeneous catalysts for petroleum cracking, and their performance is competitive with zeolite catalysts [16]. Layered double hydroxides are used for the selective conversion of propylene oxide to primary and secondary alcohols, the cross-aldol condensation reactions to form methylvinylketone, the polymerization of lactones and the decomposition of ketones [34]. Pillared MoS2 can be used in hydrodesufurization [851. As sorbents, smectite clays are used for decoloring edible oils, clarifying alcoholic beverages, removing grease from raw wool and treating the radioactive waste solutions [83]. Layered double hydroxides are used to take up acidic impurities such as HCl generated from the photo decomposition of poly(vinyl chloride) [84]. As sorbents, intercalation compounds also have potential applications in the controlled releasing of perfumes, insecticides, fertilizers [36], as well as medicines. With advances in intercalation chemistry and the increase in the number of investigated host and guest species, intercalation as an approach for materials design is more and more clearly claimed, and more effort is directed towards the synthesis of materials with novel structures and properties. Intercalation readily provides materials with highly anisotropic dielectric, conductive and optical properties [720. For example, layered 15 vanadyl phosphonates are used to place naphthalene units into ordered arrays to achieve the control of their photophysical properties [371. Intercalation can stabilize materials of nanometer dimensions, which exhibit novel electrical, optical, magnetic, and chemical properties. Intercalates of various functions have been included [38]. Luminescent compounds such as Ru(bpy)32+ have been encapsulated in transparent hosts such as smectite. The luminescent compounds are effectively isolated by the hosts so that self-quenching is suppressed. A molecularly dispersed system is also a basic prerequisite for composing efficient photochemical hole burning (PHB) materials to avoid line broadening due to energy transfer. The intercalation of PI-IB probes such as 1,4—dihydroxyanthraquinone (DAQ) into transparent hosts is an effective way to achieve this prerequisite [891. Electrochromic and photochromic compounds such as viologen (1,1'-disubstituded-4,4'-bipyridinium salts) have been intercalated into layered hosts to make electrochromic, photochromic and optical memory materials [891. Dyes have also been intercalated in order to obtain pigments with special spectral properties [89]. As a major branch of the class of intercalation compounds and probably the most applicable group, nanocomposites currently receive worldwide interest. As mentioned earlier, the organoclay/nylon nanocomposites have already been used in timing-belt covers in automobiles [35] due to their high strength, high yielding temperature and good processibility. In addition, organoclay/polymer nanocomposites have been produced as high-barrier plastic packaging materials [90], which have gas barrier improvements as high as 800%. Nanocomposites can also have better flame retardant properties [91, 9d] and lower dielectric permittivity than the component polymers [921. The regular void structures, which exist in specially designed nanocomposites, provide low dielectric constants and 16 may produce good electronic packaging materials. Materials of extremely low dielectric permittivity can also be generated from the restricted motions of ions and electrons in the network structures of nanocomposites. Those materials have potential applications for optoelectronic packagingl7al. In the field of electrode materials, the preliminary research on V205 nanocomposites shows that V205/polyaniline has a better reversibility and an increased Li capacity in the electrochemical redox cycles, implying that it is a better cathode material for Li rechargeable batteries [47]. The V205/polyaniline also has a Li chemical diffusion coefficient one order of magnitude higher than that of the host itself, which would provide better performance at high current densities [47]. The electroactive polymer intercalated nanocomposites usually have cation diffusion coefficients higher than those of the unexpanded hosts, which makes it worthwhile to explore the possibility of using the nanocomposite analog of a layered electrode material as a substitute. Polyether intercalation compounds, especially poly(ethylene oxide) intercalation compounds, are of interest in solid state ionics. PEO acts as a solid solvent for different salts. The PEO solutions of salts are polyelectrolytes [931 and are used in solid state batteries [94] and other electronic devices. The inclusion of PEO in clays and other charged layered hosts provides fast ion conduction for the incorporated cations, while the host materials act as inert, often insulating counter ions. Therefore, the ionic conductivity is exclusively due to cations of the interlayer region. This makes the transference number of the cation equal to 1 (14:1), which is of interest in the study of ion-transport phenomena in polymer electrolyte systems. Finally, the intercalation of PEO in electronically conductive layered solids opens the way to new polyelectrolyte materials of mixed ionic/electronic conductivity. PEO l7 intercalation compounds have been reported with layered silicates [43], layered metal phosphates [72b], layered metal phosphorus trisulfides [50], V205 1511, M003 [55, 56b], transition metal dichalcogenides [52. 53, 571 and carbon oxides [56d]. Using polymer intercalative nanocomposites, as precursors to produce new ceramics and graphite films, is another direction of research that carries the polymer intercalation into potential applications. Kato and coworkers used magadiite-poly(acrylonitrile) complex as a precursor to prepare silicon carbide and silicon nitride in carbothermal reduction processes [39]. In an ordinary carbothermal reduction process, a mixture of carbon and SiO2 is used as the starting material. A carbonized oxide- polymer nanocomposite, in which the carbon and the oxide are intimately mixed, may produce ceramics of better quality. At least, this is the case with silicon nitride. The same group also prepared aluminum nitride by conversion of a hydrotalcite-polyacrylonitrile intercalative nanocomposite through a carbothermal reduction process. Kyotani and coworkers prepared graphite films from montmorillonite-polyacrylonitrile nanocomposites [95]. The films consist of highly crystallized but small graphite crystallites. They have a much weaker mechanical strength and a much poorer electrical conductivity than the graphite films prepared from polymer films, but they have some flexibility. 5. Important Contributions of the Present Dissertation Work The research described in the present dissertation started at a time when a list of polymers had been intercalated in M082 by the EEP methodl52]. To promote the research of polymer intercalation in M082, which had not been finished by Bissessur [96], the study in the exfoliation of M082, intercalation of polymers, and properties of hydrated M082, 18 restacked M082 and nanocomposites was continued. The study provides a better understanding of this new group of materials, which leads to a better manipulation of the materials and of the encapsulation reactions. During this study, many new M082 nanocomposites, such as (polystyrene)xMoS2, (polyacrylamide)xMoS2, (polyvinylalcohol)xMoS2, (polypropylene)xMoS2, (polymethylmethacrylate)xMoS2, (polybenzimidazole)xMoS2 and (polyethyleneterephthalate)xMoS2 have been synthesized. This part of the research is described in Chapter 1. In order to intercalate insoluble polymers in M082, a modification was made to the EEP method, which brought in the in situ polymerization-encapsulative precipitation method. This method has been successfully applied to the synthesis of M082/polypyrrole nanocomposites, and is discussed in Chapter 2. The polypyrrole intercalation in W82 is also reported in Chapter 2. The most important achievement with respect to M082, in this dissertation, is the invention and development of the high temperature LiBH4 lithiation method. In this procedure, LiBH4 is proved to be a very powerful reducing reagent at high temperatures. Up to now, this method has been used to prepare other lithium-intercalated materials such as LiWS2l67l and LiNbSe2 [571, which can exfoliate in water. It is expected to be applicable to the lithiation of many additional layered materials. The new lithiation method provides a convenient approach to prepare large quantities of LiMoS2. Believing that the EEP method can be generalized, other hosts were also explored. Polymer intercalation was tried on 2H-TaS2, the exfoliation of which had been reported [62» 63’ 64]. However, exfoliated TaS2’" prepared according to the literature did not include polymers except polyimines. In order to achieve the polymer intercalation, different forms of LixTaS2 19 were prepared with different lithiation reactants and under different reaction conditions. Although most of the LixTaS2 phases can be hydrated and dissolve somewhat in water, no polymer other than a polyimine can be included. All this changed after the preparation of consistently uniform LixTaS2 phases through a quantitatively controlled lithiation reaction using a 0.2 equivalent of LiBH4. The exfoliation of and polymer intercalation in 2H-Ta82 is described in Chapter 3. The two LiBH4 lithiation methods, by high temperature solid state reaction and room temperature solution reaction, are complementary. The high temperature lithiation, which has a strong driving force, works with layered materials hard to reduce, such as M082 and W82. Over-lithiation should not be a problem in these materials because they oxidize readily when exposed to water in the exfoliation step. The controlled room- temperature solution lithiation, using a low equivalent of LiBH4, can deal with hosts that are readily reduced, such as TaS2 and M003. Since the reduced forms of these materials are not readily oxidized by water and the high charge density in the layers favors neither exfoliation nor polymer intercalation, the charge carried by the layers should be controlled. The two lithiation methods are expected to cover the exfoliation of a broad range of layered materials. Since the availability of stable monolayer suspensions is the prerequisite to applying the EEP method, the importance of the two lithiation methods is obvious. The exfoliation and polymer intercalation of W82 [67, 97] and NbSe2 [571 is based on the first lithiation method. By using the second lithiation method, M003 and oc-RuCl3 are exfoliated and encapsulated with polymers. The nanocomposites with the latter two hosts are described in Chapter 4 and Chapter 5. 20 The polymer insertion in M082, W82, NbSe2 [57], 2H-TaS2, M003 and oc-RuCl3 using the EEP method has demonstrated its effectiveness. The diversity of the host materials shows the generality of this method. At this stage, the EEP method has been firmly established and become one of the most effective tools in polymer intercalation. The present dissertation work contributes a substantial part to the generalization of this new polymer intercalation method and introduces many new lamellar nanocomposites. 21 References 1 C. Schafhautl, J. Prakt. Chem. 1840, 3, 129. 2 K. Fredenhagen and G. Cadenbach, Z. Anorg. Allg. Chem. 1926, fl, 249. 3 (a) L. Pietronero and E. Tosatti ed., Physics of Intercalation Compounds (Proceedings of an International Conference, Trieste, Italy, July 6-10, 1981), Springer-Verlag, Berlin/New York, 1981. (b) M. S. Dresselhaus ed., Intercalation in layered Materials (NATO ASI series, Series B: Physics, Vol. £18), Plenum Press, New York, 1986. (c) A. P. Legrand and S. Flandrois ed., Chemical Physics of Intercalation (NATO ASI series, Series B; Physics, Vol. 17_2_), Plenum Press, New York, 1987. (d) P. Bemier ed., Chemical Physics of Intercalation II (NATO ASI series, Series B: Physics, Vol. 305), Plenum Press, New York, 1993. (e) G. Slberti and T. Bein ed. Comprehensive Supramolecular Chemistry, Vol. 1 (Solid-State Supramolecular Chemistry: Two— and Three-Dimensional Inorganic Networks), Elsevier Sci. Ltd., 1996. 4 (a) R. H. Friend and A. D. Yoffe, Adv. Phys. 1987, E, 1. (b) R. F. Frindt in A. P. Legrand and S. Flandrois ed., Chemical Physics of Intercalation (NATO ASI series, Series B; Physics, Vol. H2), Plenum Press, New York, 1987, pp 195. (c) W. Y. Liang in M. Balkanski ed., Microionics-Solid State Intergrable Batteries, Elsevier Science Pub. B. V., 1991, pp 237. (d) G. Ouvrard in P. Bemier, J. E. Fischer, 8. Roth and S. A. Solin ed., Chemical Physics of Intercalation II (NATO ASI series, Series B: Physics, Vol. fl), Plenum Press, New York, 1993, pp 1. (e) J. Rouxel in G. Slberti and T. Bein ed. Comprehensive Supramolecular Chemistry, Vol. _7_ (Solid-State Supramolecular Chemistry: Two- and Three- Dimensional Inorganic Networks), Elsevier Sci. Ltd., 1996, Chapter 3, pp 77. 5 F. R. Gamble, F. J. DiSalvo, R. A. Klemm and T. H. Geballe, Science 1970, 1_68_, 568. 6 (a) J.-H. Choy, N.-G. Park, Y.-I. Kim, S.-H. Hwang, J .-S. Lee and H.-I. Yoo, J. Phys. Chem, 1995, Q, 7845. (b) J.-H. Choy, N.-G. Park, S.-J. Hwang and D.-H. Kim, J. Am. Chem. Soc., 1994, 116, 11564. (c) Y. Koike, K. Sasaki, A. Fujiwara, K. Watanabe, M. Kato, 22 10 ll 12 13 14 15 T. Noji and Y. Saito, Physica C 1995, _24_5, 332. (d) Z. J. Huang, J. G. Lin, J. J. Lin, C. Y. Huang, L. Grigoryan and K. Yakushi, Physica C 1995, m, 305. (a) E. P. Giannelis, V. Mehrotra and M. W. Russell, Mat. Res. Soc. Symp. Proc. Vol. m, 1990, 685. 03) S. Yamanaka, New Functional Mterials, Vol. C (Synthetic Process and Control of Functionality Materials), Elsevier Science Pub. B. V., 1993, 677. (a) G. A. Ozin, Adv. Mater. 1992, 4, 612. (b) R. Schollhom, Chem. Mater. 1996, _8_, 1747. (a) E. Ruiz-Hitzky, Adv. Mater. 1993, 5, 334. (b) E. Ruiz-Hitzky and P. Aranda, Anales de Quimica 1997, 23, 197. (c) E. P. Giannelis, Adv. Mater., 1996, _8_, 29. (d) E. P. Giannelis, Appl. Organometal. Chem. 1998, _l_2, 675. (e) M. M. Lerner and C. O. Oriakhi, in A. N. Goldstein ed Handbook of Nanophase Materials, Marcel Dekker, New York, 1997, Chap. 7, pp 199. (a) M. S. Whittingham and A. J. Jacobson ed., Intercalation Chemistry Academic Press, 1982. (b) D. O'Hare. in D. W. Bruce and D. O'Hare ed. Inorganic Materials, John Wiley & Sons Ltd, 1992, Chapter 4, pp 165. M. Armaud and P. Touzain, Mater. Sci. Eng. 1977, 3_1_, 1. (a) C. Schlenker ed., Low-Dimensional Electronic Properties of Molybdenum Bronzes and Oxides, Kluwer Academic Publishers, 1989. (b) A. M. Chippindale, P. G. Dickens and A. V. Powell, Prog. Solid State Chem. 1991, A, 133. (a) A. J. Jacobson in A. K. Cheetham and P. Day ed. Solid State Chemistry Compounds, Clarendon, Oxford, 1992, pp 182. (b) H Schafer—Stahl and R. Abele, Angew. Chem, Int. Ed. Engl. 1980, 12, 477. R. Clément, I. Lagadic, A. Léaustic, J. P. Audiere and L. Lomas, in P. Bemier, J. E. Fischer, 8. Roth and S. A. Solin ed., Chemical Physics of Intercalation II, Plenum Press, New York, 1993, pp 315. G. Lagaly and K. Beneke, Colloid Polym. Sci. 1991, 262, 1198. 23 l6 17 18 19 20 21 22 23 24 25 T. J. Pinnavaia, in A. P. Legrand and S. Flandrois ed., Chemical Physics of Intercalation (NATO ASI Ser., Series B; Physics, Vol. i2), Plenum Press, New York, 1987, pp 233. (a) G. Alberti and U. Costantino, in J. L. Atwood, J. E. D. Davies and D. D. MacNicol ed., Inclusion Compounds, Vol. 5, Oxford Sci. Pub., 1991, pp 136. (b) A. Clearfield and U. Costantino in G. Slberti and T. Bein ed. Comprehensive Supramolecular Chemistry, Vol. 1 (Solid—State Supramolecular Chemistry: Two- and Three- Dimensional Inorganic Networks), Elsevier Sci. Ltd., 1996, Chapter 4, pp 107. (a) B. Morosin, Phys. Lett. 197 8, QA, 53. (b) A. T. Howe and M. G. Shilton, J. Solid State Chem. 1979, _2_8, 345. (c) A. T. Howe and M. G. Shilton, J. Solid State Chem. 1980, 3_1, 393. (d) M. G. Shilton and A. T. Howe, J. Solid State Chem. 1980, _3_4, 137. J. Votinsk’y, J. Kalousova and L. Bene"s, J. Inclusion Phenom. Mol. Recognit. Chem. 1992, E, 19. (a) D. J. Jones and J. Roziere, Solid State Ionics 1989, 35, 115. (b) D. J. Jones, R. El Mejjad and J. Roziere, in T. Bein ed., Supramolecular Architecture (AC8 Symposium series fl), 1992, pp 220. (a) F. Cavani, F. Trifiro and A. Vaccari, Catalysis Today, 1991, 11, 173. (b) A. de Roy, C. Forano, K. El. Malki and J.-P. Besse, in Synthesis of Microporous Materials, Van Nostrand Reinhold, New York, 1992, pp 108. (c) K. El Malki, M. Guenane, C. Forano, A. de Roy and J. P. Besse, Materials Science Forum 1992, 91-93, 171. A. J. Jacobson, in P. Bernier, J. E. Fischer, 8. Roth and S. A. Solin ed., Chemical Physics of Intercalation II, Plenum Press, New York, 1993, PP 117. S.-T. Wong and 8. Cheng, Inorg. Chem. 1992, Q, 1165. (a) X.-M. Shen and A. Clearfield, J. Solid State Chem. 1986, 64, 270. (b) P. Le Goff, N. Baffier, S. Bach and J. P. Pereira-Ramos, Mater. Res. Bull. 1996, Q, 63. P. Lavela, J. Morales and J. L. Tirado, Chem. Mater. 1992, 4, 2. 24 26 27 28 29 30 31 32 33 34 35 36 (a) F. R. Gamble, J. H. Osiecki, M. Cias, R. Pisharody, F. J. DiSalvo and T. H. Geballe, Science 1971, 174, 493. (b) F. R. Gamble, J. H. Osiecki, F. J. DiSalvo, J. Chem. Phys. 1971, 55, 3525. R. Schbllhom and H. D. Zayefka, Angew. Chem, Int. Ed. Engl. 1977 , m, 199. (a) D. E. W. Vaughan, R. J. Lussier and J. S. Magee, U. S. Patent 4,176,090, 1979. (b) T. J. Pinnavaia, M. 8. Tzou, S. D. Landau and R. H. Raythatha, J. Mol. Catal. 1984, 2_7, 195. L. F. Nazar and A. J. Jacobson, J. Chem. Soc, Chem. Commum, 1986, 570. (a) A. Blumstein, J. Polym. Sci., A, 1965, 3, 2665. (b) C.-H. Hsu, M. M. Labes, J. T. Breslin, D. J. Edmiston, J. J. Winter, H. A. Leupold and F. Rothwarf, Nature, Phys. Sci., 1973, @055), 122. (c) G. Lagaly and K. Beneke, Am. Mineralogist 1975, 50, 650. (d) V. M. Chapela and G. S. Parry, Nature 1979, E, 134. T. J. Pinnavaia, Science 1983, m(4595), 365. (a) A. Okada and M. Kawakado, Japanese Patent Application No. Sho 60 [l985]-217396. (b) A. Okada, M. Kawasumi, T. Kurauchi and O. Kamigaito, Polym. Prepr., Am. Chem. Soc., Div. Polym. 1987, E, 447. (c) Y. Fukushima and S. Inagaki, J. Inclusion Phenom. 1987, 5, 473. (d ) Y. Fukushima, A. Okada, M. Kawasumi, T. Kurauchi and O. Kamigaito, Clay Minerals 1988, Q, 27. A. Okada and A Usuki, Mat. Sci. Eng. C—Biomin. 1995, 3, 109. M. G. Kanatzidis, L. M. Tonge, T. J. Marks, H. O. Marcy and C. R Kannewurf, J. Am. Chem. Soc. 1987, m, 3797. T. Kurauchi, A. Okada, T. Nomura, T. Nishio, S. Saegusa, R. Deguchi, SAE Technical Paper Ser. 1991, 910584. P. B. Messersmith and E. P. Giannelis, Chem. Mater. 1993, 5, 1064. 25 37 38 39 40 41 42 43 (a) M. S. Wang and T. J. Pinnavaia, Chem. Mater. 1994, Q, 468. (b) T. Lan and T. J. Pinnavaia, Chem. Mater. 1994, Q, 2216. (c) P. B. Messersmith and E. P. Giannelis, Chem. Mater. 1994, Q, 1719. (a) T. Lan, P. D. Kaviratna and T. J. Pinnavaia, Chem. Mater. 1994, 6, 573. (b) E. P. Giannelis, V. Mehrotra, O. Tse, R. A. Vaia and T.-C. Sung, Mat. Res. Soc. Symp. Proc. Vol. 261, 1992, pp 969. R. A. Vaia, H. Ishii and E. P. Giannelis, Chem. Mater. 1993, 5, 1694. M. G. Kanatzidis, C.—G. Wu, D. C. DeGroot, J. L. Schindler, M. Benz, E. LeGoff and C. R. Kannewurf, in P. Bernier , J. E. Fischer, 8. Roth and S. A. Solin ed., Chemical Physics of Intercalation II Plenum Press, 1993, pp 63. (a) C.-G. Wu, M. G. Kanatzidis, H. O. Marcy, D. C. DeGroot and C. R. Kannewurf, Polym. Mat. Sci. Eng. 1989, _6_l, 969. (b) M. G. Kanatzidis, C.-G. Wu, H. O. Marcy and C. R. Kannewurf, J. Am. Chem. Soc. 1989, 111, 4139. (c) M. G. Kanatzidis, C.-G. Wu, H. O. Marcy, D. C. DeGroot and C. R. Kannewurf, Chem. Mater. 1990, 2, 222. (d) C.-G. Wu, M. G. Kanatzidis, H. O. Marcy, D. C. DeGroot and C. R. Kannewurf, in R. M. Metzger et al ed., Lower- Dimensional Systems and Malecular Electronics, Plenum Press, New York, 1991, pp 427. (e) Y.-J. Liu, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, J. Chem. Soc., Chem. Commun. 1993, 593. (a) L. F. Nazar, X. T. Yin, D. Zinkweg, Z. Zhang and S. Liblong, Mat. Res. Soc. Symp. Proc. 1991, 250, 417. (b) L. F. Nazar, Z. Zhang and D. Zinkweg, J. Am. Chem. Soc. 1992, 154, 6239. (c) R. Bissessur, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, J. Chem. Soc., Chem. Commun. 1993, 687. (a) M.G. Kanatzidis, R. Bissessur, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf, Chem. Mater. 1993, 5, 595. (b) L. Wang, J. L. Schindler, J. A. Thomas, C. R. Kannewurf and M. G. Kanatzidis, Chem. Mater. 1995, l, 1753. (a) Y.-J. Liu and M. G. Kanatzidis, Inorg. Chem. 1993, Q, 2989. (b) G. Cao and T. E. Mallouk, J. Solid State Chem. 1991, _9_4, 59. 26 45 46 47 48 49 50 51 52 53 54 55 (a) P. Day and R. D. Ledsham, Mol. Cryst. Liq. Cryst. 1982, &, 163. (b) B. Tieke, Mol. Cryst. Liq. Cryst. 1983, 9_3_, 119. T. A. Kerr, H. Wu and L. F. Nazar, Chem. Mater. 1996, 3, 2005. (a) F. Leroux, B. E. Koene and L. F. Nazar, J. Electrochem. Soc., 1996, _lfi(9), L181. (b) F. Leroux, G. Goward, W. P. Power and F. Nazar, J. Electrochem. Soc., 1997, M, 3886. (a) E. Ruiz-Hitzky and P. Aranda, Adv. Mater. 1990, 2, 545. (b) P. Aranda and E. Ruiz-Hitzky, Chem. Mater. 1992, 3, 1395. E. Ruiz-Hitzky, P. Aranda and B. Casal, J. Mater Chem. 1992, 2, 581. I. Lagadic, A. Léaustic and R. Clement, J. Chem. Soc., Chem. Commun. 1992, 1396. (a) Y.-J. Liu, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Chem. Mater. 1991, 3, 992. (b) Y.-J. Liu, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Adv. Mater. 1993, 5, 369. (a) R. Bissessur, J. L. Schindler, C. R. Kannewurf and MG. Kanatzidis, Mol. Cryst. Liq. Cryst. 1993, 2:15, 249. (b) R. Bissessur, M. G. Kanatzidis, J. L. Schindler and C. R. Kannewurf, J. Chem. Soc., Chem. Commun. 1993, 1582. E. Ruiz-Hitzky, R. Jimenez, B. Casal, V. Manriquez, A. 8. Ana and G. Gonzalez, Adv. Mater. 1993, 5, 738. The small organic or inorganic molecules have only been reported being intercalated in M082 by using the M082 monolayer suspension: (a) P. Joensen, R. F. Frindt and S. R. Morrison Mater. Res. Bull. 1986, 21, 457. (b) M. A. Gee, R. F. Frindt, P. Joensen and S. R. Morrison Mat. Res. Bull, 1986, 21, 543. (o) W. M. R. Divigalpitiya, R. F. Frindt and S. R. Morrison, Science 1989, 2A6, 369. L. Wang, J. Schindler, C. R. Kannewurf and M. G. Kanatzidis J. Mater. Chem. 1997, Z, 1277. 27 56 57 58 59 60 61 62 63 64 65 66 (a) J. P. Lemmon, J. Wu, C. Oriakhi and M. M. Lerner, Electrochim. Acta 1995, Q , 2245. (b) C. O. Oriakhi, R. L. Nafshun and M. M. Lerner, Mater. Res. Bull. 1996, 3_1, 1513. (c) L. F. Nazar, H. Wu and W. P. Power, J. Mater. Cheml995, 5, 1985. ((1) Y. Matsuo, K. Tahara and Y. Sugie, Carbon 1996, 3A, 672. H.-L. Tsai, J. L. Schindler, C. R. Kannewurf, and M. G. Kanatzidis, Chem. Mater. 1997, 2, 875. (a) R. A. Vaia, S. Vasudevan, W. Krawiec, L. G. Scanlon and E. P. Giannelis, Adv. Mater. 1995, _7_, 154. (b) R. A. Vaia, K. D. Jandt, E. J. Kramer and E. P. Giannelis, Macromolecules 1995 , 28, 8080. (c) R. A. Vaia and E. P. Giannelis, Macromolecules 1997, 30, 7990. (d) R. A. Vaia and E. P. Giannelis, Macromolecules 1997, 3Q, 8000. A. Weiss and E. Sick. Z. Naturforsch., Teil B, 1978, 33, 1087. P. Colomban and M. Pham Thi, Rev. Chim. Miner. 1985, 22, 143 (Fr). (a) R. E. Grim, Clay Mineralogy, 2nd ed., McGraw-Hill, New York, 1968. (b) P. H. Nadeau, Appl. Clay Sci. 1987, 2, 83. (c) H. Van Olphen, in A. C. D. Newman ed., Chemistry of Clays and Clay Minerals, Wiley, New York, 1987, pp 203. (d) L. Lee Nelson Eur. Pat. Appl. EP 312,954, 26 Apr 1989 (US Appl. 110,169, 19 Oct 1987). A. Lerf and R. Schbllhorn, Inorg. Chem. 1977, 16, 2950. C. Liu, 0. Singh, P. Joensen, A. E. Curzon and R. F. Frindt, Thin Solid Films 1984, m, 165. D. W. Murphy and G. W. Hull, Jr., J. Chem. Phys. 1975, _6_2, 973. L. F. Nazar, S. W. Liblong and X. T. Yin, J. Am. Chem. Soc., 1991, 1_13, 5889. (a) B. K. Miremadi and S. R. Morrison J. Appl. Phys. 1988, 63, 4970. (b) D. Yang and R. F. Frindt, J. Phys. Chem. Solid 1996, 5_7_ (6-8, Proceedings of the 8th International Synposium on Intercalation Compounds, 1995), 1113. 28 67 68 69 70 71 72 73 74 75 76 77 H.-L. Tsai, J. Heisign, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis Chem. Mater. 1997, 9, 879. R. Clement, O. Garnier and J. Jegoudez, Inorg. Chem. 1986, Q, 1404. G. Alberti, M. Casciola and U. Costantino, J. Colloid Interface Sci. 1985, 1_0_7, 256. H. Rebbah, M. M. Borel and B. Raveau, Mater. Res. Bull. 1980, 1_5, 317. M. M. J. Treacy, S. B. Rice, Allan J. Jacobson, J. T. Lewandowski Chem. Mater. 1990, 2, 279. (a) C. Y. Ortiz-Avila and A. Clearfield, Inorg. Chem. 1985, 2A, 1733. (b) A. Clearfield and C. Y. Ortiz-Avila, in T. Bein ed., Supramolecular Architecture (AC8 Symposium Series 499), American Chemical Society, Washington, DC, 1992, pp 178. (c) E. W. Stein, Sr., C. Bhardwaj, C. Y. Ortiz-Avila, A. Clearfield and M. A. Subramanian, Materials Science Forum 1994, 152-153 (Soft Chemistry Routes to New Materials), 115. (d) C. Y. Ortiz-Avila, C. Bhardwaj and A. Clearfield, Inorg. Chem. 1994, 33, 2499. A. J. Jacobson, in G. Alberti and T. Bein ed., Comprehensive Supramolecular Chemistry, Vol. 1 (Solid -State Supramolecular Chemistry: Two— and Three-Dimensional Inorganic Networks), Elsevier Science Ltd., 1996, pp 315. P. Bonneau, J. L. Mansot and J. Rouxel, Mater. Res. Bull. 1993, 28 757. J .-S. Xu, Y. Tang, H. Zhang and Z. Gao, Gaodeng Xuexiao Huaxue Xuebao 1997,fi(1), 93(Ch). K. Domen, Y. Ebina, S. Ikeda, A. Tanaka, J. N. Kondo and K. Maruya Catal. Today 1996, 28, 167. J .—H. Choy et al, unpublished results. 29 78 79 80 81 82 83 84 85 86 87 88 M. Armand, in W. Van Gool ed., Fast Ion Transport In Solids, North Holland, New York, 1973, pp 655. M. S. Whittingham, U. 8. Patent 4,009,052, Feb. 22, 1977. (a) B. B. Owens, Survey of Lithium Ambient Temperature Secondary Systems (Fifth International Seminar on Lithium Battery Technology and Applications), Deerfield Beach, Fla, March 1989. (b) K. M. Abraham, in S. Subbarao. V. R. Koch and B. B. Owens ed., Proceedings of the Symposium on Lithium Batteries, Vol. 9&5 (ECS Fall Meeting 1989), 1990, pp 1. (a) A. R. Armstrong and P. G. Bruce, Nature 1996, E, 499. (b) P. Le Goff, N. Baffier, S. Bach and J. P. Pereira-Ramos, Mater. Res. Bull. 1996, 31, 63. (c) A. Manthiram and J. Kim, Chem. Mater. 1998, 1_0_, 2895. (a) A. Clearfield and M. Kuchenmeister, in T. Bein ed., Supramolecular Architecture (ACS Synosium fl), 1992, Chapter 10, pp 128. (b) A. Clearfield, M. E. Kuchenmeister, K. Wade, R. Cahill and P. Sylvester, in M. L. Occelli et al ed., Synthesis of Microporous Materials, 1992, Chapter 12, pp 245. G. Alberti and U. Costantino, in G. Slberti and T. Bein ed., Comprehensive Supramolecular Chemistry, Vol. 1 (Solid—State Supramolecular Chemistry: Two- and Three-Dimensional Inorganic Networks), Elsevier Sci. Ltd., 1996, Chapter 1, pp 1. S. Carlino, Chem. Brit., Sept. 1997, 59. J. Brenner, C. L. Marshall, L. Ellis, N. Tomczyk, J. Heising and M. G. Kanatzidis, Chem. Mater. 1998 10 1244. 9_, A. Sopkova, J. Inclusion Phenom. Mal. Recognit. Chem. 1992, H, 5. M. R. Torgerson, Layered Metal Phosphonates as Templates for the Assembly of Ordered Molecular Systems, Ph. D. Dissertation, Department of Chemistry, Michigan State University, 1996. J.-F. Nicoud, Science 1994, E, 636. 30 89 90 91 92 93 94 95 96 97 (a) C. Kato, in Memoirs of the School of Science & Engineering, Waseda Univ., 1991, No. Q, pp 41. (b) K. Kuroda, M. Ogawa, T. Yanagisawa and C. Kato, Mat. Res. Soc. Symp. Proc. Vol. &, 1993, pp 335. (a) B. Miyares, Packaging Strategies 1995, (8). (b) P. B. Messersmith and E. P. Giannelis, J. Polym. Sci., A: Polym. Chem. 1995, Q, 1047. (c) J. Gu, The Barrier Characteristics of Clay/Polyimide Nanocomposites, M. S. thesis, School of Packaging, Michigan State University, 1997. J. W. Gilman, T. Kashiwagi and J. D. Lichtenhan, Sample Journal 1997, EM), 40. Y. Kim, W. K. Lee, W. J. Cho, C. 8. Ha, M. Ree and T. Chang, Polymer International 1997, Q, 129. (a) P. V. Wright, Br. Polym. J. 1975, 1, 319. (b) M. B. Armand, J. M. Chabagno and M. Duclot, in P. Vashishta, J. N. Mundy and G. K. Shenoy ed., Fast Ion Transport in Solids Electrodes and Electrolytes, Elsevier, New York, 1979, pp 131. (c) M. Armand, Solid State Ionics 1983, m, 745. (d) M. Armand, Adv. Mater. 1990, 2, 278. M. Gauthier, D. Fauteux, G. Vassort, A. Bélanger, M. Duval, P. Ricoux, J. M. Chabagno, D. Muller, P. Rigaud, M. B. Armand and D. Derco, J. Electrochem. Soc: Electrochem. Sci. Technol. 1985, Q, 1333. (b) C. C. Hunter, D. C. Sinclair, R. West and A. Hooper, J. Power Sources 1988, E, 157. T. Kyotani, T. Mori and A. Tomita, Chem. Mater. 1994, Q, 2138. R. Bissessur Synthesis and Characterization of Novel Intercalation Compounds of Molybdenum Trioxide and Molybdenum Disulfide, Ph. D. Dissertation, Department of Chemistry, Michigan State University, 1994. The intercalation of polypyrrole in WS2 is reported in Chapter 2. The intercalation of other polymers, done by H.—L. Tsai and M. G. Kanatzidis, is to be published in the future. 31 Chapter 1 FURTHER EXPLORATION OF EXFOLIATED M082 AND SYNTHESIS OF NEW M082 NANOCOMPOSITES WITH THE EXFOLIATION— ENCAPSULATIVE PRECIPITATION METHOD Introduction The intercalation of polymers in layered inorganic solids is much more difficult than the intercalation of small molecules because of their large macromolecular size. The conventional direct intercalation method, which accounts for the intercalation of most simple organic molecules, is usually not applicable to polymers. The alternative route, which involves monomer intercalation followed by intra-gallery polymerization, seems elegant but is hard to carry out. Prior to our exploration of the M082/polymer system, through which a new convenient synthesis method was developed [1], the majority of the inorganic/polymeric nanocomposites were classified into two categories: (a) nanocomposites of conjugated polymers in oxidative hosts synthesized by the redox intercalative polymerization, and (b) nanocomposites of clays fabricated with organoclay precursors which can be delaminated. Due to the lack of highly oxidizing layered materials, the application of the redox intercalative polymerization is limited to certain hosts [2]. On the other hand, versatile clay nanocomposites had been prepared with clays swelled and exfoliated by surfactants 13]. This route pointed to a promising approach to new classes of nanocomposites. As for other layered materials, however, the methodology had been seldom applied, since the swelling property of clays is mirrored in only a few layered inorganic compounds. The V205 xerogel, 32 which swells spontaneously in water and encapsulates a variety of water- soluble polymers [4], is an example. The new intercalation method, which we name exfoliation- encapsulative precipitation (EEP) method, was first reported with the simple organic compound intercalation in M082 [5]. It was then introduced in the field of polymer intercalation in the development of the M082/polymer nanocomposites [1] and proved to be convenient and effective. In this method, the layered inorganic host is exfoliated and forms a monolayer suspension. The monolayers combine with dissolved polymeric molecules when a polymer solution is added, and the two types of components precipitate out as a lamellar nanocomposite. The successful application of the EEP method to M082, which is chemically inert, and which cannot be directly intercalated by even simple organic molecules, demonstrates the potential of the delamination methodology to the synthesis of nanocomposites with common layered inorganic compounds. Later, the EEP method was generalized to other layered hosts such as M003 161, TaS2[7], W82 18], NbSe2 [9] and so on. Nanocomposites prepared with this method are even superior to the clay nanocomposites in respect of compositional and structural simplicity, since no surfactant is involved. Earlier work in this group [1] succeeded in encapsulating many types of soluble polymers, which included water soluble polymers such as poly(ethylene oxide) (PEO), poly(propylene glycol) (PPG), polyvinylpyrrolidone (PVP), polyethylenimine (PEI) and methyl cellulose (MCel), common plastics such as polyethylene (PE) and nylon-6 (PA6), and conductive polymers such as polyaniline (PANI). The physical properties of the restacked M082 and the synthesized M082/polymer nanocomposites were also examined. It was noticed at that time that the 33 restacked M082 probably had the intralayer structure of 1T-MoS2 [10] in which the Mo atoms were octahedral coordinated. This form was metastable converting exothermically to the normal 2H form of M082 which had Mo atoms in the trigonal—prismatic coordination. Later, it was proved by electron diffraction that the M0 in the exfoliated and freshly restacked M082 has a distorted octahedral coordination [11]. The conversion of the M082 causes many interesting changes in the properties of restacked M082 and nanocomposites. Although extensive, the investigation in exfoliated/restacked M082 and nanocomposites was not completed in Bissessur’s dissertation work. Since M082 is a readily available commercial material and has many important applications such as lubrication, the catalyzation in hydrodesulfurization [12] and the rechargeable battery fabrication [13], a more extensive exploration in M082/polymer nanocomposites was conducted in the present dissertation research. This ‘ research was carried in four directions: (a) searching for more economic and convenient methods to prepare exfoliated M082, (b) further studies to better understand the nature and properties of exfoliated M082, restacked M082, and important M082/polymer nanocomposites such as (PEO)xMoS2, (c) further development of the EEP method, and (d) exploration of new strategies to synthesize M082 nanocomposites with insoluble polymers. The research on the former three parts is reported in this chapter. The last one will be described in Chapter 2. 34 Experimental Section 1. Reagents M082 (99%, lum size aver.) was purchased from Cerac Inc. n-Butyl lithium (2.5 M solution in hexane), LiBH4 (95%), poly(ethylene oxide) (Molecular Weight 100,000), nylon-6 (MW 10,000), polystyrene (MW 280,000), polyacrylamide (MW 5,000,000), poly(vinyl alcohol) (MW 124,000—186,000), polypropylene (MW 250,000), poly(methyl methacrylate) (MW 120,000), polybenzimidazole (melting point > 300 OC), poly(ethylene terephthalate) (MW 18,000) and high density polyethylene (MW 125,000) were purchased from Aldrich Chemical Company. Polyvinylpyrrolidone (MW 10,000) was from Sigma Chemical Co. CaH2 (>95%), hexane (GR) and ethyl acetate (GR) were purchased from EM Science Inc. Formic acid (88%, ACS Grade), decalin (purified), 200 proof ethanol and phenol (reagent grade) were from Columbus Chemical Industries Inc., Fisher Chemical Company, Quantuum Chemical Company and J. T. Baker Chemical Co. respectively. 2,2,2-Trifluoroethanol (99%) was purchased from both Lancaster Synthesis Inc. and PRG Catalog Inc. Water for experiments and processing was deionized water. 2. Lithiation and Exfoliation of M082 a. Preparation of LiMoSz by reaction of M082 with n-butyl lithium In this classical method of preparation of LiMoS2 [14], M082 was reacted with excess n-butyl lithium (n-LiBu) in solution under an inert atmosphere. In a typical reaction, 4 g of M082 (0.025 mol) was reacted in 75 m1 of 1 M n-BuLi/hexane solution (0.075 mol of n-BuLi) for 3 days. The LiMoS2 was collected by filtration, washed with copious hexane and 35 pumped to dryness. The reaction and processing were under nitrogen atmosphere. The hexane was distilled from CaH2 under nitrogen before use. The product was stored in a nitrogen glove box after preparation. b. Preparation of LiMoS2 by reaction of M082 with LiBH4 This convenient new method of LiMoS2 preparation was developed in the present research. In this method, M082 was combined with LiBH4 in a solid state reaction at an elevated temperature to produce LiMoS2. In a typical reaction, 10.00 g of M082 (0.0625 mol) and 2.04 g of LiBH4 (0.0938 mol) were ground and mixed together in a nitrogen glove box. They were put into a ceramic crucible and placed in a quartz tube (2.5 in. diameter, 34.5 in. length) which had a balloon connected to it, see Figure 1.1. The quartz tube was taken out of the glove box and inserted into a programmable furnace so that the crucible was in the middle of the hot zone while the outlet, connected to the balloon, was out of the furnace. The furnace was heat up to 350 0C in 3 h, kept at this temperature for 72 h and then cooled to room temperature in 3 h. The gas produced by the reaction was accumulated in the balloon which expanded, thus avoiding explosion. After the reaction, the quartz tube was separated from the balloon and taken into the nitrogen glove box. The product was collected as a mixture of LiMoS2, LiBH4 and LiH, and stored under a nitrogen atmosphere. c. Exfoliation of LiMoS2 The LiMoS2 prepared by the LiBu procedure was pure. An amount of 0.2 g of this LiMoS2 (1.2 mmol), was typically used in a reaction, and was exfoliated in 20 ml of water by 1 h of stirring. The exfoliated M082 was collected by centrifugation to separate from the LiOH produced 36 Balloon Rubber tubing Furnace Quartz tube Figure 1.1. Experimental set—up for the high temperature LiBH4 lithiation reaction. 37 by the reaction. It was re-dispersed in 20 ml of fresh water by 30 min of stirring for succeeding reactions. The product from the LiBH4 lithiation method was a mixture of LiMoS2, LiBH4 and LiH. An amount 0.213 g of this mixture contained about 0.2 g LiMoS2 (1.2 mmol). This amount of the mixture was stirred in 20 ml of water for 1 h. Centrifugation was used to collect exfoliated M082 from the solution. The restacked M082 was washed again with 20 ml of water and re-dispersed in 20 ml of fresh water by 30 min of stirring for subsequent reactions. 3. Investigation of LiMoS2 and Freshly Restacked M082 a. Thermal stability of LiMoS2 LiM082 prepared by the LiBu procedure was checked by the differential scanning calorimeter (DSC) up to 300 CC. The curve of DSC did not show exothermic or endothermic peaks, which meant no obvious phase change in the above temperature range. A second heating cycle gave almost the same curve as the first one. The sample could exfoliate after this process. b. Elemental analysis and 7Li NMR of freshly restacked M082 An amount of 1.07 g of the LiBH4, LiH and LiMoS2 mixture, which was produced from the LiBH4 lithiation reaction and contained about 1.00 g LiMoS2 (6.0 mmol), was washed 6 times by stirring in 100 m1 portions of water for 30 min and collecting by contrifugation. The restacked M082 so obtained was pumped for a week to dryness and used in 7Li solid state nuclear magnetic resonance (NMR) experiments and for inductively coupled plasma (ICP) spectroscopy analysis. 38 NMR experiments detected strong 7Li signals which assured the existance of Li inside the sample. The ICP analysis showed 0.686 wt% Li, 51.8 wt% Mo and 35.9 wt% 8, which corresponds to Li0,13MoS2. 4. Preparation of M082/Polymer N anocomposites a. Polyl ethylene oxide) (PEO) nanocomposites The ratio of PEO to M082 used typically in this work was ~ 2 to 1. In a typical reaction, 20 ml of an aqueous solution containing 0.106 g (2.4 mmol) PEO was mixed with 20 ml of aqueous M082 monolayer suspension containing 1.2 mmol of M082, and stirred for 2 days in a stoppered 125 m1 Erlenmeyer flask. The product nanocomposite was collected by centrifugation and washed thrice with 20 ml portions of water. The product was then dried in air and pumped to dryness. To cast a (PEO)xMoS2 film, the paste of the product was spread on a glass plate in air before drying. For reactions using other PEO/MoS2 ratios, the amount of polymer varied accordingly. In reactions designed to produce delaminated (PEO)xMoS2 nanocomposites, the ratios of PEO to M082 used were 2, 4, 8 and 16, and the concentrations of the polymer in solution were 3%, 4%, 8% and 13% respectively. The concentration of the M082 monolayer suspension was twice as that of the typical one described above. After the two solutions, 10 ml of M082/H2O and a corresponding amount of PEO/H2O, were mixed and stirred for 2 days, the mixture was pumped to dryness. X-ray diffraction indicated that the mixture was composed of the known (PEO),,M082 phase ((1 spacing 16.2-17.7 A) and bulk PEO. Neither delaminated (PEO)xMoS2 nanocomposites nor nanocomposites with widely spaced layers were produced. 39 b. Polle-vinyl pyrrolidone) (PVP) nanocomposites In a reaction, 20 ml of aqueous M082 monolayer suspension, which contained 0.2 g of M082 (1.2 mmol), was mixed with 20 ml of aqueous PVP solution which contained 0.533 g of PVP (4.8 mmol VP repeat units), and was stirred for 2 days. The (PVP)xMoS2 was collected by centrifugation, washed 3 times with water, dried first in air and then under vacuum. The (PVP)xMoS2 product had a basal spacing ~ 28 A. Washing with water did not have a significant effect on the d spacing of the product. c. N Ion-6 A6 nanocom osites Procedure 1: As the first of the two procedures, (PA6)xMoS2 nanocomposites were prepared using PA6/CF3CH2OH solution. The method was a modification from the one used earlier [1]. In a typical reaction, 10 ml of M082 suspension containing 0.20 g M082 (1.2 mmol) was added dropwise into 20 ml of PA—6/CF3CH2OH solution under stirring. The CF3CH2OH solution contained 0.49 g of PA6 (4.32 mmol C6H11NO repeat units). The solution was stirred for 2 days before it was centrifuged to isolate the nanocomposite from it. The product was washed with 5 ml of CF3CH2OH/H2O (2:1) solution, and dried first in air and then under vacuum. In the reaction above, the ratio of PA6 to M082 was 3.621. Other ratios such as 0.2, 0.5, 1.0, 2.0, 4.0, or 10 to 1 were also used. Procedure 2: (PA6)xMoS2 nanocomposites can also be synthesized by a reaction between an aqueous M082 monolayer suspension and a PA6 formic-acid solution. In a typical reaction, 20 ml of M082 suspension containing 0.20 g M082 (1.2 mmol) was added dropwise into 100 ml of 40 PA6/HCOOH solution which contained 0.68 g of PA6 (6.0 mmol C6H11NO repeat units) under vigorous stirring. A large volume of HCOOH was required to keep PA6 in solution. The solution was stirred for 3 h before it was centrifuged to separate the nanocomposite. The product was washed with water 3 times and methanol once, and dried first in air and then under vacuum. d. Polystyrene (P8) nanocomposites (PS)xMoS2 nanocomposites were successfully prepared by mixing and reacting the aqueous M082 monolayer suspension with very concentrated PS/(ethyl acetate) solutions. In a typical reaction, 20 ml of M082 suspension containing 0.20 g M082 (1.2 mmol) was mixed with 20 ml of ethyl acetate solution which contained 5.0 g of PS. The two solutions were immiscible, so the mixture was stirred vigorously for 2 days. When the mixture was centrifuged, it separated in 3 layers: PS/(ethyl acetate) solution on the top, water solution at the middle and (PS)xMoS2 nanocomposite at the bottom. The (PS)xMoS2 nanocomposite was collected and dried without washing. Washing with ethyl acetate de-intercalates PS from M082. e. Polyacgylamide (PAM) nanocomposites In a typical reaction to prepare a (PAM)xMoS2, 20 ml of M082 suspension containing 0.20 g M082 (1.2 mmol) was mixed with 40 ml of aqueous PAM solution which contained 0.43 g of PAM (containing 6.0 mmol AM units). The mixture was stirred for 2 days before it was centrifuged to collect (PAM)xMoS2, which was washed with water 3 times, and dried first in air and then under vacuum. 41 f. Polylvinyl alcohol) (PVA) nanocomposites PVA is only slightly water soluble. In a typical reaction to prepare a (PVA)xMoS2 nanocomposite, 20 ml of M082 suspension containing 0.20 g M082 (1.2 mmol) was mixed with 80 ml of aqueous PVA solution which contained 0.53 g of PVA (12 mmol VA units). The PVA did not totally dissolve. The mixture was stirred for 9 days before it was filtered through a coarse filter paper to remove undissolved PVA. The solution through the filter paper was centrifuged to collect (PVA)xMoS2. The nanocomposite was washed and dried as in the case of (PAM)xMoS2 above. g. Polypropylene (PP) nanocomposites An amount of 0.056 g PP (1.3 mmol C3H6 repeat units) was dissolved in 40 ml of hot decalin (100 - 140 OC). The hot PP solution was added into 20 ml of cold aqueous M082 monolayer suspension, which contained 0.20 g of M082 (1.2 mmol), under vigorous stirring. The mixture was stirred at 90 0C for 4 h. The product was collected hot by centrifugation, and washed first with copious hot decalin (100 0C) and then with ether. h. Poly(methyl methacrylate) (PMMA) nanocomposites Procedure 1: An amount of 0.12 g of PMMA (containing 1.2 mmol MMA units) was dissolved at 50 0C in a mixture of 50 m1 ethanol and 5 ml water. 10 ml of aqueous M082 monolayer suspension, which contained 0.1 g M082 (0.6 mmol), was added dropwise into the PMMA/C2H50H/H2O solution while stirring vigorously at 50 0C. The mixture was stirred at 50 0C for 12 h after the addition of M082. The product was collected by 42 centrifugation, washed 3 times with formic acid and 3 times with water, and dried first in air and then in vacuum. Procedure 2: An amount of 0.60 g of PMMA (containing 6 mmol MMA units) was dissolved in 100 ml of formic acid. 10 ml of aqueous M082 monolayer suspension, which contained 0.1 g M082 (0.6 mmol), was added dropwise into the PMMA/HCOOH solution, which was stirred vigorously. The mixture was stirred for 12 h after the addition of M082. Since PMMA did not precipitate out, the product was collected by centrifugation. The nanocomposite was washed and dried as in Procedure 1 above. 1. Polybenzimidazole (PBI) nanocomposites An amount of 10 ml of aqueous M082 monolayer suspension, which contained 0.1 g of M082 (0.6 mmol), was added dropwise into 75 ml of saturated PBI formic-acid solution (containing about 0.37 g PBI, or 1.2 mmol BI units), which was stirred vigorously during addition. After the mixture was stirred for 12 h, the product was isolated and processed the same as described in Procedure 1 above. j. Polyl ethylene terephthalate (PETP) nanocomposites An amount of 0.56 g of PETP (containing 3.0 mmol ETP units) was dissolved in 25 n11 of phenol at 50 0C. The PETP/phenol solution was added dropwise into 10 ml of aqueous M082 monolayer suspension which contained 0.1 g M082 (0.6 mmol) and which was stirred vigorously at 50 0C. The mixture was stirred at 50 0C for 2 days after the addition. Phenol and water are immiscible, so vigorous stirring was required. The product 43 was collected by centrifugation at an elevated temperature (50 - 75 0C), washed with 75 0C phenol 3 times, and dried as above. k. Elevated temperature annealing and variable temperature X-ray powder diffraction measurements for nanocomposites Samples of nanocomposites, (PEO),,M082, (PA6)xMoS2, (PS)xMoS2, (PE)xMoS2 and (PP)xMoS2, were vacuum sealed in Pyrex tubes and heated in an oven at temperatures between 150 and 175 0C for several days to check the possible thermal decomposition of the nanocomposites. X-ray powder diffraction patterns were collected as a function of rising temperature for a (PEO),,M082 sample (not heat treated) to investigate the decomposition process. The patterns were collected at room temperature, 80 OC, and then every 10 0C from 100 0C to 230 OC. The sample was stabilized at each temperature for 10 min before the pattern was collected. The patterns were collected from 20 5° to 8° and then from 13° to 16°. Each temperature step took 40 min. 5. Instrumentation Powder X-ray diffraction (XRD) patterns were obtained on a Rigaku Ru-200B X-ray diffractometer, at 45 kV and 100 mA with a scintillation counter detector and a graphite monochromator to produce Cu Koc beam (wavelength 1.54184 A). Powder samples and a continuous scanning mode with a scanning speed of 1°/min in 20 and an increment of 001° were chosen for general purpose spectra. Variable temperature XRD measurements were done with a ceramic heating mask under the protection of nitrogen. The scanning speed was 0.2°/min. XRD experiments for one- 44 dimensional electron density calculations were done as described in Chapter 3. The amount of polymers in the nanocomposites was determined by TGA measurements with a Shimadzu TGA-50 under a 46 ml/min oxygen flow, and a heating rate of 10 oC/min. The decomposition temperature of materials was checked with TGA measurements under a 57 ml/min nitrogen flow. DSC was carried out on a Shimadzu DSC-50 under nitrogen flow of a rate of 20 ml/min. The heating and cooling rates were 5 oC/min. Sample cells were made of aluminum, and were annealed at 450 0C in vacuum sealed tubes after they were cleaned. Samples were sealed in cells under a nitrogen atmosphere before measurement. Infrared spectra were collected with a Nicolet IR/42 FTIR spectrometer in 2 cm-1 resolution. Generally 64 scans were collected. Samples were measured as KBr pellets. The amount of Li in restacked M082 was measured with Inductively Coupled Plasma (ICP) spectroscopy at Animal Health Diagnostic Laboratory [15] on the MSU campus. Variable temperature solid state 7Li NMR spectra were taken on a 400 MHz Varian Nuclear Magnetic Resonance Instrument. Room temperature electrical conductivity measurements for Figures 1.6 and 1.9 were done on pressed sample pellets with a four-probe detector connected to a Keithley-236 source measure unit. Room temperature conductivity measurements for data in Table 1.4 were done on pressed sample pellets with a four—probe detector connected to a Keithley-580 micro-Ohmmeter. 45 Results and Discussion 1. The LiBH4 Method of Producing LiMoS2 The exfoliated M082, which is the crucial material in the preparation of M082 nanocomposites using the EEP method, is generated from the reaction of LiMoS2 with water. In earlier studies of M082/polymer nanocomposites, LiMoS2 was prepared by the conventional method [14] which uses LiBu in a room temperature reaction. Despite the strict anaerobic reaction conditions and the large excess of highly reactive LiBu, this method is still the best way to prepare pure LiMoS2, especially in small quantities. Here this route was still used occasionally to prepare LiMoS2 for the reactions in which pure LiMoS2 was preferred as a starting material. (By “pure”, we mean free of LiH and LiBH4, as discussed below.) In the preparation of most of the M082/polymer nanocomposites, however, the use of pure LiMoS2 is not necessary, which suggests the possibility of a more convenient and economical route for preparation. The synthesis of LiMoS2 with LiBH4 at elevated temperature provides such a route. LiBH4, and its analogs NaBH4 and KBH4, are convenient reducing reagents in terms of storage and manipulation. Their stability in dry air and slow hydration in water [16] requires neither sophisticated equipment nor extreme care in processing. In most of its reducing reactions, the byproducts are B2H6 and H2. These gases are separated from the solid or liquid products spontaneously, therefore, the products are usually pure and no additional purification step is needed. In our LiBH4 route to prepare LiMoS2, the reaction also generates these two byproducts: 2M0$2 + 2LIBH4 ------ > 2L1M0$2 + B2H6 + H2 46 The reaction probably generates some LiH because of the decomposition of LiBH4 at high temperature: 2 LIBH4 ------ > 2LIH + B2H6 The LiH and the excess LiBH4 stay in the solid phase with LiMoS2 as a mixture. Fortunately, in order to generate exfoliated M082, it is not necessary to purify LiMoS2 from this mixture of products which contains LiBH4 and probably LiH. When the product is put into water, the LiBI-I4 and LiH dissolve and decompose to form LiOH, and the resulting exfoliated M082 can be readily purified by washing with water [171. 2LiMoS2+2(1-X)H2O ----- >2LixM082+2(1-X)LiOH+(l-x)H2 2LiBH4 + 2H2O ------ > B2H6 + 2H2 + 2LiOH L1H + H20 ------ > H2 + LIOH Therefore, this new route produces pure exfoliated M082 for intercalation reaction, although the LiMoS2 from the lithiation reaction is in a mixture with LiH and LiBH4. A high quality monolayer M082 suspension should be free from the unexfoliated material. This requires all M082 to be converted to LiMoS2 in the lithiation reaction. Since both M082 and LiBH4 can be ground to fine powder and mixed well before the lithiation reaction, complete lithiation of M082 is achieved. Therefore, the high temperature LiBH4 lithiation reaction produces high-quality exfoliated M082. Since the new reaction is carried out in the solid state, a small reaction vessel can deliver a large quantity of product. The reaction set-up 47 is simple and no strict air-exclusion is necessary. LiBH4 is more expensive than LiBu ($118/mol vs. $53/mol according to the price of Aldrich Chemical Co.). However, considering the fact that only 1.5 equivalents of LiBH4 is used and no solvent purification is required, the LiBH4 approach is actually more economic. LiBH4 is generally considered as a mild reducing reagent and has been used mostly in organic reactions. In the field of inorganic chemistry, it has been used to reduce M03, V205, FeOCl, TiS2, TaS2 [131. At room temperature, LiBH4 cannot reduce chemically inert materials such as M082, while its usage at elevated temperatures had not been seen reported. The preparation of LiMoS2 with LiBH4 at an elevated temperature is the most convenient method to obtain LiMoS2 so far. There have existed several other routes to prepare alkali metal dichalcogenide ternary compounds, however, all of them have some shortcomings. The preparation reaction using binary chalcogenides and alkali metals in liquid NH3 at ca. -30 0C [191 is even harder to deal with than the reaction with LiBu. The reaction of binary chalcogenides with alkali halide melts in a H28 gas flow [20] needs solid-gas reaction facilities. In addition, the constant blowing of toxic H28 is undesirable. The reaction of binary chalcogenides with alkali metals at ca 800 0C [21] seems convenient. However, when we used this method to prepare alkali metal intercalated M082 and then exfoliated M082, we found the existence of un-reacted M082 in the product. The possibility of using another less-expensive reducing reagent such as LiAlH4 or LiH to substitute LiBH4 was also explored without success. Although LiAlH4 can produce LiMoS2 at temperatures higher than 300 0C, the by-product A1H3 forms Al(OH)3 in water, which is hard to separate from exfoliated M082 and interferes with the polymer encapsulation 48 process. It causes the production of low quality nanocomposites with short coherence length. LiH did not lithiate M082 under similar conditions and a more severe condition (720 0C) caused decomposition. Similar to LiBH4, NaBH4 also introduces Na” in M082 and the product forms hydrated NaxMoS2 in an aqueous solution. Nevertheless, hydrated NaXMoS2 does not readily include polymers, therefore more investigation is needed to develop it into a starting material for nanocomposites. 2. Manipulation and Properties of Exfoliated and Restacked M082 In the old exfoliation procedure of LiMoS2, the suspension was sonicated for 30 min after the material had been put in water. The purpose of this ultrasonic wave treatment was to maximize the possibility of full exfoliation. The strong perturbation associated with this process broke the M082 up into small sized layers and reduced the quality of the nanocomposites. Later we found that this ultrasonic process was not necessary because completely exfoliated M082 can be obtained when LiMoS2 is just stirred in water. The cause of exfoliation of LiMoS2 in water was once suggested to be the splitting of the M082 layers by the violent reaction of the material with water and the vigorous generation of the hydrogen gas inside the interlayer space [53, 5C]. This mechanism predicts that pre-oxidation of LiMoS2 would severely affect the exfoliation process, because it would reduce the strength of this explosive reaction. This prediction demanded the strict exclusion of oxygen and moisture from LiMoS2 before its exfoliation. The fact that we were able to insert PEO in LiMoS2, which was exposed to air for ten hours, showed the ability of exfoliation of freshly 49 oxidized LiMoS2 and demonstrated a negligible effect of the pre-oxidation to the exfoliation property of LiMoS2. Therefore, the splitting mechanism of the exfoliation seems questionable. Further investigations show that the exfoliation is due to the hydration or solvation of the M082 layers. The above experiments also suggest that storage of the LiMoS2 under inert atmosphere for the purpose of exfoliation does not have to be extremely stringent. It was believed that when LiMoS2 was exposed to water, it was fully oxidized back to M082 and the exfoliated layers were neutral. Researchers in this area were excited by the thought of neutral suspended M082 monolayers in water because this phenomenon would be an exception among the exfoliated layered inorganic compounds. Other layered inorganic compounds either have charged layers or need assistance from surfactants to form monolayer suspensions. Our experience suggests that the concept of neutral M082 layers is incorrect and they actually carry negative charge. One reason for believing this is that exfoliated M082 encapsulates cations more readily than neutral molecules. Another reason is that solid state 7Li nuclear magnetic resonance (NMR) of restacked M082, which was carefully washed with water, proved the existence of Li in this material. In addition, ICP spectroscopic analysis revealed that the composition of the freshly restacked M082 was probably LixMoS2 (x = 0.18). More convincing experiments which prove that the exfoliated and freshly restacked M082 layers possess charges are those done by Heising and Kanatzidis [11], which include the electron diffraction study revealing the intralayer structure of the freshly restacked M082 different from 2H- MoS2, and 1H—NMR experiments which prove that the freshly restacked M082 evolves hydrogen gas when it is heated. Therefore, restacked M082 is 50 analogous to LixTiS2 and LixTaS2, which exfoliate in water and some polar solvents because the layers carry charges. The incomplete oxidation of the M082 layers when LiMoS2 is exposed to water and the charges that the exfoliated M082 layers carry suggests a solvation mechanism for the exfoliation. The negative charges of the M082 layers hold Li+ cations inside the interlayer galleries to balance the charge, while the Li+ cations are solvated by water molecules. The solvation of the Li+ cations makes the M082” layers exfoliate. This mechanism explains why the pre-oxidized LiMoS2, which in fact is not M082 but LixMoS2, can still exfoliate. It also explains why the water molecules are hard to remove out the freshly restacked M082 1221 and why the freshly hydrated M082 can exfoliate again. Dehydration at elevated temperature such as 45 or 60 0C [51’] is more effective, since heating accelerates the conversion to 2H-MoS2 [11 and the additional oxidation of LixMoS2. A similar explanation applies to the fact that the pre-oxidized and high temperature (150 OC) treated LiMoS2 could not exfoliate. 3. (PEO)xMoS2 and (PVP)xMoS2 Nanocomposites In previous work [1], the reaction for (PEO),,M082 nanocomposites was carried out exclusively around the stoichiometrical ratio of 1 mol PEO to 1 mol M082. In the present research, reactions have been run under different PEO to M082 ratios to explore other (PEO)xMoS2 nanocomposite phases. Although the ratio of PEO to M082 has been increased gradually from 1 to 20 in the reactions, the basal spacing of the resulting nanocomposite always falls in the range between 15.3 to 18.5 A, as indicated by XRD patterns. Even in the reactions designed to produce delaminated (PEO),,M082 nanocomposites, in which the aqueous M082/PEO 51 solution was dried under constant agitation, the reaction would not produce nanocomposites of higher d spacing, or delaminated (PEO)xMoS2 nanocomposites. (A constant agitation during drying the mixture is considered the best way to keep the M082 and PEO from phase separation.) Therefore, basically only one type of (PEO)xMoS2 nanocomposite has been prepared from the solution-encapsulation reaction. Contrary to the results obtained with 1 to 1 ratio of PEO to M082, when the ratio is increased to 2:1 or above, the probability of producing a well ordered nanocomposite product is largely increased, and the products are free from the restacked M082 phase. The order in the structure of the nanocomposite can be judged from the sharpness of the peaks in the X-ray diffraction patterns. The d spacings of the products usually range between 16 and 17 A, and the nanocomposites usually contain 1.5 equivalents of PEO instead of 1.0, as revealed by TGA measurements in an oxygen flow. These experiments indicate that some excess of polymer is needed to fully occupy the galleries of M082. It must be mentioned that prolonged excessive washing of the freshly collected nanocomposites with water will lower the quality of the product. After this treatment, the nanocomposites frequently have X-ray patterns of low and broad peaks, which indicates poor stacking orders. From X-ray diffraction patterns, it is calculated that the coherence lengths of the nanocomposites in the stacking direction may decrease from about 120 A to about 43 A [23]. In some cases, excessive washing would also cause the appearance of the deintercalated phase. The one-dimensional electron density map of a (PEO),,M082 nanocomposite has been calculated from high quality reflection XRD patterns carefully collected from an oriented nanocomposite film, see 52 001 002 3‘ a d a C: 1—1 003005 007008 0011 00 008010 0012 d 0013 x5 J _. AAAAA x10 I I I I I I I I I I I I I I I I I I I l I I T] I—I I 0 20 40 6O 80 100 120 140 20 (deg) Figure 1.2. XRD patterns of a (PEO)xMoS2 nanocomposite used to calculate the one-dimensional electron density map. Electron Density c—axis Figure 1.3. One-dimensional electron density map of a (PEO),,M082 nanocomposite. (Solid line, calculated from experimental data; dash line, from model.) 53 I1.51 A 3.27 A Pacesgssee .0... PEG ‘3‘ ’3‘ a a a a ,. A A ' A A A A A A a A ~ A M082 0.0.0.000... 14.46 A Figure 1.4. Structural model for the (PEO),,M082 nanocomposite. 54 Figures 1.2 and 1.3. The features of the map are almost the same as those for the Li0,2(PEO)xTaS2 nanocomposite which will be presented in Chapter 3. The model proposed according to the one-dimensional electron density map is shown in Figure 1.4, which is similar to the one for Li0_2(PEO)xTaS2. In this model, the PEO chains take a conformation that was found in type II PEO-HgCl2 complex. Since the one-dimensional electron density map and the model for the Li0,2(PEO)XTaS2 nanocomposite will be described in detail in Chapter 3, a similar discussion is not given here. Possible existing phases other than the phases reported in the preceding research have also been searched in (PVP)XM082 nanocomposites. In previous work, the PVP to M082 ratio used in the encapsulation reaction was 1 to 1 [11. It was suspected that the amount of PVP might not have been sufficient to produce the nanocomposite with maximum polymer content. Reactions with excess PVP have verified the suspicion and produced a (PVP)XM082 of higher basal spacing, 28.8 A (see Figure 1.5A) instead of 21.1 A. The amount of polymer in the nanocomposite is 0.96 VP units rather than 0.76 VP units obtained from the 1:1 ratio. The temperature of the structural transition from the octahedral to the trigonal-prismatic M082, however, is 174 OC, almost the same as that of the previously reported (PVP)xMoS2, 177 0C. 4. Investigations in (PA6)xMoS2 Nanocomposites The reaction to encapsulate PA6 in M082 used earlier [1] was modified here. The new procedure makes use of the advantage that a solution composed of 1/3 of water and 2/3 of CF3CH2OH does not cause PA6 to precipitate, so that no PA6 will come out of solution when the 55 Intensity (PAM)O~6M082 r k ‘% 16.9 A -- JPVA)35M082 Figure 1.5A. XRD patterns of the new nanocomposite phases (1). 56 35.2 A (PA6)025M082 - 1. ; ; "':;T 44.4A (PMMA)O.90M082 g.» 60.1 A E” .2 (PMMA) MoS '5 044 2 28.6A (PBI)0_65M082 33.4A (PETP)028M082 [IIIIITIIIIIIIIIIIIIIIllllllfl'—I 0 5 10 15 20 25 3O 20 (deg) Figure 1.5B. XRD patterns of the new nanocomposite phases (2). 57 aqueous M082 monolayer suspension is added into the PA6/CF3CH2OH solution. These conditions are considered good for nylon-6 encapsulation, because they are free from competing side reactions, and the product is free from excess nylon-6. The basal spacing of the nanocomposite is 14.8 A, smaller than the previously reported value 17.5 A [1], see Figure 1.5A. The composition determined by TGA is (PA-6)xMoS2 (x=0.26—0.44). The driving force to form (PA6)xMoS2 is not strong in the CF3CH2OH/water solution. Although the amount of nylon-6 in the nanocomposite is only about 0.26-0.44 C6H11NO per M082, more than 2 equivalents of nylon-6 must be used to make the product free from the restacked M082 phase. When the ratio of nylon-6 to M082 is 1 or less, the (PA6)xMoS2 nanocomposite of 15.0 A d spacing forms together with restacked M082. When the ratio decreases from 1 to 0.2, the amount of restacked M082 increases substantially from 1/3 to 2/3 of the total product. When the ratio is 2, no restacked M082 is observed. Instead, a phase that has a basal spacing of 10.6 A appears. This phase will diminish when the ratio increases, however, it will not be eliminated even when the ratio reaches 10. At a ratio of 2, the amount of the 10.6 A phase is about 40%; at a ratio of 3.6, the amount is about 30%; at a ratio of 10, it is about 22%. The reactions are summarized in Table 1.1. The encapsulation using M082 dispersed in CF3CH2OH did not succeed. The product from this reaction in water-free CF3CH2OH solution was found to be mostly restacked M082. The weak driving force to form (PA6)xMoS2 nanocomposite might be the reason for the failure of encapsulation. CF3CH2OH is not an ideal solvent for reactions. It is volatile, highly toxic and extremely expensive. An alternative solvent for nylon-6 is formic 58 Table 1.1. Nylon-6 intercalation reactions with PA6/CF3CH2OH solution PA6:MoS2 phase of phase of phase of (mol:mol) d=15 A d=10.6 A 6.2 A 02:10 33% — 67% 10:10 67% - 33% 2.0: 1.0 60% 40% - 3.6210 70% 30% - 10: 1.0 78% 22% - acid, however, it causes the exfoliated M082 to precipitate, which is an unwanted competing side reaction. Delaminated (PA6)xMoS2 nanocomposites or high d—spacing products have been obtained by mixing the aqueous M082 monolayer suspension and the nylon-6 formic-acid solution but some restacked M082 phase often exists. Because the quick precipitation of exfoliated M082 in formic acid solution, the arrangement of polymer chains is poor and loose. A sample with a 35 A d spacing (see Figure 4B) has only a polymer content of 0.25 C6H11NO units per M082. 5. Synthesis of (PS)xMoS2 N anocomposites Polystyrene (P8) is the most common commercial polymer. It can be dissolved in many organic solvents, including both polar and non—polar solvents. Therefore, it is often a model polymer for basic studies in polymer science. However, a difficulty has been met in the experiments encapsulating polystyrene in exfoliated M082: polystyrene does not have a strong tendency to combine with the M082 monolayers. The solvents for polystyrene, on the other hand, are more or less prone to combine with 59 M082 monolayers and stay in the interlayer galleries of M082. As a result, in many cases the solvents compete with polystyrene to form solvated M082 ' phases. These solvents include benzene, carbon tetrachloride, chloroform, acetone, toluene, decalin and THF. Out from all of the solvents we tried, ethyl acetate turns out to be suitable for polystyrene encapsulation. To maximize polystyrene's ability to compete, concentrated polystyrene/(ethyl acetate) solutions were used. The concentration of the solution was 20 wt% or above, which is very high compared to the concentrations of other polymer solutions mentioned above. (PS)xMoS2 nanocomposites prepared are either delaminated nanocomposites, or nanocomposites with a very large (1 spacing (see Figure 1.5A), depending on the concentration of the polystyrene solution and other encapsulation conditions. Washing (PS)xMoS2 nanocomposites with ethyl acetate causes the deintercalation of polystyrene and produces restacked M082. The information of some (PS)xMoS2 nanocomposites prepared under different conditions are summarized in Table 1.2. Table 1.2. Information about the synthesis of M082/PS nanocomposites PS reaction PS/MoS2 in reaction concentration temperature d spacing composite (V%) (A) (mol/mol) trial 1 20% RT. 38.4~41.4 2.3/1.0 trial 2 23% RT. -* 3.8/1.0 trial 3 20% 65 0C -* 6.8/1.0 * The latter two samples are delaminated. 60 The reaction at an elevated temperature (65 OC) favors the completion of intercalation. This phenomenon is reminiscent of the successful high temperature encapsulation of the other two non-polar polymers, polyethylene and polypropylene. At elevated temperature, the exfoliated M082 converts to the 2H form, which probably facilitates the intercalation of non-polar polymers. 6. Synthesis of Other M082/Polymer N anocomposites Since many water-soluble polymers were already intercalated in M082 [1] with no difficulty, the effort in the present research was mainly on the intercalation of polymers dissolved only in non-aqueous solutions. The only two new nanocomposites which contain water-soluble polymers are (PAM)xMoS2 (PAM = polyacrylamide) and (PVA)xMoS2 (PVA = polyvinylalcohol) (see Figure 1.5A and Table 1.3). Again, these two polymers are intercalated readily. It was noticed by Bissessur that polypropylene (PP) can be intercalated in M082 with the same procedure as that for polyethylene. However, because no good sample was obtained, he did not mention this nanocomposite in his dissertation. Satisfactory (PP)xMoS2 samples have been obtained in the present research. Like (PE)xMoS2 nanocomposites, (PP)xMoS2 nanocomposites have basal spacings around 10.2-10.3 A, see Figure 1.5A. The weight percentage of polymer is also similar in these two nanocomposites. (PE)1,5MoS2 contains 21% polymer while (PP)1.1M082 contains 22% polymer. Poly(methyl methacrylate) (PMMA) is also a very common polymer, which is soluble in many solvents. Encapsulation of PMMA is successfully carried out in formic acid or a 50 OC ethanol/water solution. The 61 encapsulation using PMMA/(formic acid) solution is similar to the encapsulation of nylon-6 in formic acid. M082 precipitates immediately when it is added into PMMA/(formic acid) solution, encapsulating some polymer to form the nanocomposite. The polymer chains are not well arranged in the interlayer galleries. The product has a large basal spacing, 60 A (see Figure 1.5B), but a low polymer content, 0.44 MMA units per M082. If the reaction is not controlled well, restacked M082 phase may appear. In 50 0C ethanol/water solution, neither polymer nor exfoliated M082 precipitates in short time, so the encapsulation can be carried out without strongly competing side reactions. The polymer chains can take time to arrange themselves inside the interlayer galleries and pack more tightly. Therefore, the product should have a better quality (in terms of stacking order) and a higher polymer content. The basal spacing was only 44 A (see Figure 1.5B) but the polymer content is 0.9 MMA units per M082. PMMA can also be dissolved in ethyl acetate, however, encapsulation with ethyl acetate solution was not successful. Polybenzimidazole (PBI) is a high-temperature engineering polymer. The (PBI)xMoS2 nanocomposite synthesized using PBI/(formic acid) solution has a basal spacing of 28.6 A (see Figure 1.5B) and a polymer content of 0.65 BI units per M082. Poly(ethylene terephthalate), an important engineering polymer like nylon, was also encapsulated in M082 using a 50 oC phenol solution. The nanocomposite has a basal spacing of 33 A (see Figure 1.5B) and a polymer content of 0.28 ETP units per M082. Information on these nanocomposites is summarized in Table 1.3. 62 Table 1.3. Characteristics of some new M082 nanocomposites nanocomposite solvent for Treaction d spacing T0h-->2H reaction (A) (DC) (PEO)1,5MoS2 H20 R.T. 17.0 159 (PVP)0,96MoS2 H2O R.T. 28.8 174 (PA6)0,44M082 CF3CH2OH R.T. 15.1 153 (PA6)0,25M082 HCOOH R.T. 35.2 126 (PS)2,3MoS2 CH3COOCH2CH3 R.T. 38.4 undetected (PS)6_8M082 CH3COOCH2CH3 65 OC delaminated undetected (PAM)0,60MoS2 H2O R.T. 15.6 200 (PVA)35MoS2 H20 R.T. 16.9 - (PP)1_1MoS2 decalin 40-90 0C 10.3 184 (PMMA)0,90M082 ethanol 50 0C 44.4 - (PMMA)0,44M082 HCOOH R.T. 60.1 1 17 (PBI)0.65MoS2 HCOOH R.T. 28.6 - (PETP)0_28M082 phenol 50 0C 33.4 - 63 7. The Phase Transition in Restacked M082 and M082 Nanocomposites The exfoliated M082 and freshly restacked M082 have a metastable structure in which Mo atoms have a distorted local octahedral coordination. This structure originates from the lithiation reaction, which converts the 2H form M082 (trigonal-prismatic coordination) to the IT form LiMoS2 (octahedral coordination). The structural conversion takes place to keep a lower coordination energy in the product LiMoS2 [101. When LiMoS2 is converted to exfoliated M082 by reacting with water, Mo changes an oxidation state +3 to an oxidation state close to +4, in which the trigonal prismatic coordination rather than the octahedral coordination provides lower coordination energy. However, exfoliated M082 or freshly restacked M082 does not complete the structural conversion immediately due to a kinetic detainment. Mo adapts a distorted octahedral coordination [10] and slowly converts to the trigonal prismatic coordination over time. The conversion is exothermic and is accelerated at an elevated temperature or under pressure [11. It is important to have knowledge of the rate of the structural conversion, because the two forms of M082 have different physical properties and the structural conversion of M082 brings about a slow property evolution in restacked M082 and nanocomposites. For example, the octahedral M082 is metallic while the trigonal prismatic M082 is a semiconductor. The conversion of M082 makes the conductivity of restacked M082 drop gradually at room temperature from 12 S/cm to 0.07 S/cm in 4 months. Although the activation energy of the structural transition was measured by Bissessur 11], the room-temperature conversion of the materials was not tracked. Here, the progress of the room- 64 (A) 1 month —~>I 1 year——+§ o 8 1" i E E : _cz _ _ g _ - 29 _ . IE 100 :— _, .‘é E E 3:3 3 curve fit: , . . 101— p = 32.531 1 ~39" 0.000194 (t- 1230);] _ _ I I 1 [III I I I I II:II| I I I l I III] I I : 102 103 104 Time (hr) 1.00 (B) / A 9’ E 0.80— / — U / * / / CF) 060— / — v / >3 0/ E 0.40— / / — ”ti; / ° E / m 0.20- / ’6 — )3 9 000 / I I I I I I 20 40 60 80 100 l 20 140 l 60 Time (hr) Figure 1.6. Evolution of the electrical resistivity of a restacked M082 sample as a function of time. (The sample was dehydrated by methanol.) 65 temperature conversion was followed by a time dependent measurement of the materials' electrical conductivity. Figure 1.6A shows the follow-up of the resistivity of a restacked M082 sample over more than 3 years. In the 3 years, the resistivity increased almost 3 orders of magnitude, with the major change happening in the first four months. DSC experiments showed that the sample after the 3 years of measurement did not have any thermal transition from room temperature up to 300 °C, which indicates the completion of the structural conversion. The profile of resistivity can be roughly fitted with an exponential function as follows [241: p = 32.5 .k [1 _ e — 0.000194(t-18.0)] (2*cm (1) Possible relationships between the change of conductivity (or resistivity) and the conversion of the M082 layers are discussed in Appendix A. Figure 1.6B shows that the resistivity increases linearly along with time in the early stages. Actually, the short-time-period resistivity-time relationship is almost linear in all the ranges. The linear relationship of resistivity and time can be expressed in a first order equation: P = 90 + b t (2) This relationship is also observed in other restacked M082 and M082 nanocomposite samples. The coefficient b, which reflects the rate of the structural conversion of M082, is listed in Table 1.4. The coefficient b of a nanocomposite is smaller than the one of restacked M082, which indicates a slower conversion of M082 in nanocomposites. This agrees well with the previous findings of the higher transition temperatures of nanocomposites detected by DSC. A new fact revealed by the comparison of coefficient b is that the additional oxidation in air accelerates the structural conversion of M082. This conclusion is 66 Table 1.4. Evolution rates of restacked M082 and nanocomposites derived from the change of electrical resistivity samples environment b (Q*cm*day'1) restacked M082 air 3.3, 5.0, 6.7 [25] (PEO)xMoS2 air 0.066, 0.40 (PA6)xMoS2 air 1.0, 2.7, 3.4 restacked M082 nitrogen 1.0, 1.5 * The different b values of the same nanocomposite were obtained from different samples. drawn because the coefficients b of the restacked M082 samples in air are larger than those in nitrogen. 8. Thermal Stability of M082 N anocomposites Annealing at high temperatures, for example 160-170 0C, causes a slow decomposition in some of the nanocomposites. Deterioration has been seen in samples of (PEO)XMoS2 and (PA6)xMoS2 heated at 160-170 0C for 2.5 days. X-ray patterns of the nanocomposites after annealing show a significant weakening and broadening of the 001 peaks. The 001 peaks also shift to lower d spacings. After annealing at 160-170 0C for 2.5 days, the (PS)xMoS2 sample that had a broad 001 peak around 38.4 A lost its peak in the X-ray pattern and became amorphous. However, annealing at this temperature seems to have no effect on (PE)xMoS2 and (PP)xMoS2 nanocomposites. The higher stability of these two nanocomposites probably originates from the PE and PP’s higher heat resistance and stronger tendency to align and pack in lattices or templates. In all the five nanocomposites, annealing does not bring up an intense restacked M082 peak in the X—ray patterns. Table 1.5 summarizes the effects of annealing 67 Table 1.5. The effect of high temperature annealing to some nanocomposites sample dbefore dafter comments on XRD pattern (A) (A) after the annealing (PEO)1,5MoS2 17.0 16.4 001 peak intensity decreased (PA6)0,1M082 16.8 15.2 (PA6)0,25MoS2 35 .2 - 001 peak intensity decreased (PS)2,3MoS2 38.4 amorphous a very low peak at 6.4 A like (PE)4_5MoS2 10.2 10.2 no change in the peak (PP)1,1M082 10.3 10.3 no change in the peak * The samples were annealed at 160-170 0C for 60 hours. AM (PEO) M08 ,9 AK (PEO) M08 150 °c 72 hours a L... 3 E. [I (PP)XMOSZ 160-170 °C 60 hours IllIl llllllIl 0 10 20 30 40 50 60 20 (deg) Figure 1.7. XRD patterns of the (PEO)xMoS2 and (PP)xMoS2 nanocomposites before and after the high temperature annealing. 68 250_ . 200 _ j >. - - 9‘5, : _ : g 150 _ : i H .. _. a - _ p—i _ _ eg 100_' j o - - Q-t - _ 50_§ - t - 2 0 . 80 100 120 140 160 180 200 220 Temperature (0C) Figure 1.8. Intensity of the 001 peak in the XRD patterns of a (PEO)xMoS2 sample at increasing temperatures. on the M082 nanocomposites. In Figure 1.7, the X-ray patterns of some annealed (PEO)xMoS2 and (PP)xMoS2 samples are compared with those of the original samples. A variable temperature powder X-ray diffraction experiment was done on a (PEO),,MoS2 sample to investigate the effect of temperature on the rate of the deterioration. Figure 1.8 shows the result of the experiment with a plot which displays the variation of the intensity of the 001 peak at increasing temperatures. The plot shows that the deterioration was slow when the temperature was less than 100 0C. In the range of 100 to 140 0C, the deterioration speeded up. After 140 0C, the intensity change of the 001 peak slowed down, probably due the fact that the order of the lamellar structure had been so poor that additional deterioration did not bring up a more significant disorder. In this high temperature XRD experiment, the 6.2 A peak, which belongs to restacked M082, was not observed. 69 Concluding Remarks This chapter presents extensive results on the exfoliation and polymer intercalation of M082. A significant achievement was the development of the high temperature LiBH4 lithiation method, which provides a convenient way to produce large quantities of exfoliated M082 and which promises to be applicable to many other layered metal chalcogenides. Further investigation was devoted to the exfoliated M082 and restacked M082 in order to identify the material, explain the mechanism of the exfoliation and seek better design of intercalation reactions. We learned that the identity of the exfoliated M082 and restacked M082 is hydrated LixMoS2, with x about 0.18. This material exfoliates in water under weak shearing forces, similar to hydrated LixTiS2 and LixTaS2. Once the hydrated LixMoS2 is dehydrated, it cannot be re- hydrated or re-exfoliated, probably due to the additional oxidation involved in those de—intercalation experiments, and the conversion to 2H- M082. Better synthesis methods were found for known nanocomposites. For example, (PE0)xMoS2 and (PPV)xMoS2 of higher polymer content were produced when sufficient polymer was provided in the encapsulation reaction. In addition, many new M082 nanocomposites were synthesized and added to the list, which include (PS)xMoS2, (PAM)xMoS2, (PVA)xMoS2, (PP)xMoS2, (PMMA)xMoS2, (PBI)xMoS2 and (PETP)xMoS2. One-dimensional electron density maps obtained for (PEO)xMoS2 based on XRD powder patterns suggest a (PEO)xMoS2 model in which the PEO chains take a conformation found in type II PEO-HgCl2 complex, similar to 70 the cases of Li0_2(PEO)xTaS2 and Li0,2(PEO)xRuCl3 which will be presented in Chapters 3 and 5. The evolution of the electrical conductivity of the restacked M082 was monitored. The curve of the increasing resistivity can be roughly fit with an exponential function of time. The electrical conductivity, which is inversely proportional to the resistivity, mainly changes in the first 4 months, by almost 3 orders of magnitude. The electrical conductivity studies also demonstrate that the intercalation of polymers slows down the process of the structural conversion of the M082 layers. In addition, they show that the conversion is slower in nitrogen than in air. Many M082/polymer nanocomposites are not stable at temperatures higher than 100 OC. Heat treatment not only causes the structural conversion of the M082 layers, but also destroys the lamellar structure. 71 Appendix A The decrease of the conductivity of restacked M082 is caused by the gradual conversion of the M082 layers, from a metallic form to a semiconducting form. There must exist some relationship between the conductivity (or resistivity) and the amount of M082 converted. On the other hand, the conversion of M082 is a certain function of time. These two events together decide the evolution of the conductivity (or resistivity). This section provides a discussion about the possible mathematical functions in these two events by deducing the equations which express the conductivity as a time function and which fit the experimental data. In conductor-insulator mixtures, the conductivity often follows a simple power law: 6 = C2(¢-¢p)° (3) where 0 is the conductivity, C2 a constant, 6 the volume fraction of the conductor, and 0p the percolation threshold. Percolation theory [261 predicts that the (pp for a three-dimensional network of conducting globular aggregates in an insulating matrix should be ~0.16. In many conductor- insulator mixtures, however, this threshold is not seen or is extremely small [27]. Equation 4, which generally fits the conductivity profile (see Figure 1.9), is deduced supposing that the simple power law suits the mixture of the restacked M082 and the conversion of M082 precedes as a second order reaction. 6 = 233*[1 + k(t—18.0)]'°‘ + 0035*{1 + [k(t-18.0)]’l }'B (4) k=0.06155, OL=1.517, 13:2.701 72 The coefficients in the equation are derived by curve fitting to the experimental data. The unit of time, the moment of t = 0, and the time adjustment of 18.0 hours, are the same as explained in discussion section 7 [24]. The deduction of equation 4 uses the equations of the second order reaction [maul - 060:0)? = la. (5) ¢1T(t=0) = 1, (6) and the equation of the simple power law 6 = Cushion” + C2H*[¢2H(t)]B. (7) If the assumption above is correct, the time for half of the lT-MoS2 to convert is k-l, which is about 16 hours. In nature, most self conversion reactions are first order reactions, for example, the decay of the radioactive isotopes. The combination of the first order reaction of the restacked form of M082 and the simple power law of conductivity can not produce a function which fits the experimental conductivity data. As described in the discussion section, the profile of resistivity can be roughly fit with an exponential function, equation 1: p = 32.5 * [1 _ e - 0.000194(t-l8.0)] gram (1) If the conversion of M082 layers from the octahedral to the prismatic coordination proceeds as a first order reaction ¢1T(t) = 6"“, (8) to derive an equation of the above type the resistivity need to be proportional to the amount of the converted M082 P = 90*[1'¢1T(t)]. (9) Nevertheless, the assumption that the resistivity is proportional to the amount of converted M082 does not have a solid experimental or theoretical basis, so this second curve fitting is less likely than the first one. 73 100 curve fit: I IITI I 1 111111 0' = 28.3*[1+k(t-18.0)]'° + 0.035*[1+1/[k(t-1 6.0)]1"3 =0.0615, 0t=1.52, 0:270 10 I I IIIIIII I I IlIlIIl I I IIIIIII I I IlIIIII 0.1 Conductivity (S/cm) I I 111111] Tj IIIIIII 0.01 l 11111111 1 11111411 1 11111111 1 1111.“ 101 102 103 104 105 Time (hr) Figure 1.9. Curve fitting for the time dependence of electrical conductivity of (PEO)XMoS2. 74 References 10 (a) R. Bissessur, M. G. Kanatzidis, J. L. Schindler, C. R. Kannewurf, J. Chem. Soc., Chem. Commun. 1993, 1582. (b) R. Bissessur, Synthesis and Characterization of Novel Intercalation Compounds of Molybdenum Trioxide and Molybdenum Disulfide, Ph. D. ‘ Dissertation , Department of Chemistry, Michigan State University, 1994. (a) M. G. Kanatzidis, L. M. Tonge, T. J. Marks, H. O. Marcy and C. R. Kannewurf, J. Am. Chem. Soc. 1987, £12, 3797. (b) M. G. Kanatzidis, C.-G. Wu, H. O. Marcy and C. R. Kannewurf, J. Am. Chem. Soc. 1989, m, 4139. (c) G. Matsubayashi and H. Nakajima, Chem. Lett. 1993, 31. (a) Y. Kojima, A. Usuki, M. Kawasumi, A. Okada, T. Kurauchi and O. Kamigaito, J. Polym. Sci. A: Polym Chem, 1993, 31, 983. (b) H. Shi, T. Lan and T. J. Pinnavaia, Chem. Mater. 1996, 3, 1584. (c) E. P. Giannelis, Adv. Mater., 1996, 3, 29. Y. -J. Liu, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Adv. Mater. 1993, 5, 369. (a) P. Joensen, R. F. Frindt and S. R. Morrison Mater. Res. Bull. 1986, 21_, 457. (b) M. A. Gee, R. F. Frindt, P. Joensen and S. R. Morrison Mat. Res. Bull, 1986, 21_, 543. (c) W. M. R. Divigalpitiya, R. F. Frindt and S. R. Morrison, Science 1989, A6, 369. L. Wang, J. Schindler, C. R. Kannewurf and M. G. Kanatzidis J. Mater. Chem. 1997, Z, 1277. (a) Chapter 3. (b) L. Wang, J. Schindler, C. R. Kannewurf and M. G. Kanatzidis, paper in preparation. H.-L. Tsai, L. Wang, J. Schindler, C. R. Kannewurf and M. G. Kanatzidis, paper in preparation. H.—L. Tsai, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Chem. Mater. 1997, 2, 875. (a) F. Wypych and R. Schbllhorn, J. Chem. Soc., Chem. Commun. 75 11 l2 13 14 15 16 17 18 19 20 21 22 1992, 1386. (b) D. Yang, S. J. Sandoval, W. M. R. Divigalpitiya, J. C. Irwin and R. F. Frindt, Phys. Rev. B. 1991, $3, 12053. (c) M. A. Py and R. R Haering, Can. J. Phys. 1983, Q, 76. (a) J. Heising and M. G. Kanatzidis, J. Am. Chem. Soc. 1999, 1__1, 638. (b) J. Heising, Synthesis and Characterization of Novel Intercalation Compounds of Molybdenum Trioxide and Molybdenum Disulfide, Ph. D. Dissertation, Department of Chemistry, Michigan State University, 1999. O. Weisser and S. Landa, Sulfided Catalysts, Their Properties and Applications, Pergamon, New York, 1973. (a) H. Tributsch, Faraday Discuss. Chem. Soc. 1980, _7_()_, 190. (b) C. Julien, S. I. Saikh and G. A. Nazri, Mater. Sci. Eng. 1992, B_15, 73. (a) M. B. Dines, Mat. Res. Bull. 1975, 10, 287. (b) D. W. Murphy, F. J. Di Salvo, G. W. Hull, Jr., and J. V. Waszczak, Inorg. Chem. 1976, L5, 17. We thank Dr. W. Emmett Brazelton and Kirk J. Stuart for the measurements. F. A. Cotton and G. Wilkinson, Advanced Inorganic Chemistry, 5th Ed., John Wiley & Sons, New York, 1988, pp 190-191. LiMoS2 itself also generates LiOH when it reacts with water. The presence of the LiOH will usually not significantly affect the polymer intercalation reaction. M. G. Kanatzidis and T. J. Marks, Inorg. Chem. 1987, 2_6, 783. (a) W. Riidorff, Chimia, 1965, _1_9_, 489. (b) J. Rouxel, M. Danot and J. Bichon, Bull. Soc. Chim. Fr. 1971, 3930. R. Schollhorn and A. Lerf, J. Less-Common Met. 1975, :12, 89. W. P. Omloo and F. Jellinek, J. Less-Common Met. 1970, 20, 121. A restacked M082 freshly collected from an aqueous monolayer suspension was a hydrated phase with a basal spacing of 12.1 A. Two days of pumping the sample, at room temperature, only dehydrated about 50% of the material giving a basal spacing of 6.2 A. 76 23 24 25 26 27 Washing the freshly restacked M082 with anhydrous methanol altered the basal spacing of M082 from 12.1 A to 9.4 A. The de- intercalation of methanol could have been accomplished by one t9 two days of dynamic pumping, to produce a phase with a 6.2-6.3 A basal spacing. This dehydrated M082 could not be re-exfoliated. For the calculation of the X-ray scattering coherence length of the nanocomposites in the restacking direction, see Reference 18 of Chapter 2. The unit of time in the equation is hour. The time at which the LiMoS2 was exposed to water is chosen as t = 0. Since the structural conversion of M082 probably started at some time after the first exposure to water, a time adjustment is imposed in the equation. The time adjustment of 18.0 hours is decided by the best fitting of the equation to the experimental data. This time adjustment also compensates the fact that the M082 converted at a slower speed when it was stirred in a closed reaction flask or dried in a vacuum chamber than when it was stored and measured in air. The b value calculated from Figure 5B is 0.16 Q*cm*day-1. Since the measurements for Figure 5 were done with pressed pellets and a four-probe detector connected to a Keithley-236 source measure unit rather than the Keithley-580 micro-Ohmmeter, the values should not be compared to those in this table. R. Zallen, The Physics of Amorphous Solids, Wiley, New York, 1983, Chapter 4. A. J. Heeger and P. Smith, in J. L. Brédas and R. Silbey ed., Conjugated Polymers, Kluwer Academic Pub., Dordrecht/Boston/London, 1991, pp 141. 77 Chapter 2 INSERTION OF POLYPYRROLE AND POLY(N—METHYL PYRROLE) IN M082 AND W82 BY AN IN SITU POLYMERIZATION- ENCAPSULATIVE PRECIPITATION METHOD Introduction Great effort has been dedicated to the development of transition— metal-compound/conductive-polymer nanocomposites [1] under the expectation that novel electrochemical properties will be obtained by the molecular level interaction of the two electrochemically active components. These expectations now begin to materialize based on recent research which reveals that certain polymer nanocomposites indeed possess better electrical or electrochemical properties, especially in ion conductivity [2] and ion mobility 13]. Early transition metal dichalcogenides make a large class of layered materials which possesses interesting electronic, photoconductive and other physical properties [4]. It is an important class of host materials and has been subject to intensive studies [5]. Many papers have been published on the intercalation of metal atoms and small organic compounds in these dichalcogenides in the past 25 years [5’ 6]. In addition, there have been a few reports about polymer intercalation in these compounds [71, including the recent reports from our group [1d, 16]. However, reports about conductive polymer intercalated nanocomposites of these compounds are still scarce. Further work to synthesize and characterize more nanocomposites with conductive polymers and this group of layered chalcogenides would be useful. Polypyrrole (PPY) is an important conductive polymer which has attracted considerable attention due to its high electrical conductivity. It has 78 been investigated as an electrode material for solid-state batteries [8a] and capacitors [8b], as an anode material for polymer light-emitting diodes [8C], an electromagnetic interference shielding material [3d] and sensor materiall8e]. Therefore, PPY intercalative nanocomposites could be promising materials for applications. PPY is an intractable polymer, however, which does not dissolve in any solvent. The exfoliation-encapsulative precipitation (EEP) method, which was described in Chapter 1, cannot be applied to PPY. By 1994, when we started the exploration of encapsulating PPY in M082, the material had already been intercalated in a number of layered inorganic compounds such as FeOCl [la] V205 [1b] and VOPO4 [1‘]. The intercalation was achieved by the method of in situ redox intercalative polymerization, which requires oxidative hosts and is suitable for only a few layered inorganic compounds. Some non-oxidative hosts can be rendered oxidative (suitable to the conductive polymer intercalation) by incorporating in them oxidizing ions such as Cu2+19l. However, this may not be a good method for restacked M082 and W82, because these materials themselves could reduce Cu2+. Therefore, we tried a new methodology to intercalate PPY in this type of hosts. Besides mixing with polymers in solution, exfoliated layered compounds can be mixed with monomers or monomer solutions. When polymerization is initiated in such mixtures, nanocomposites are frequently formed. The important clay/polyamide nanocomposites were synthesized with this strategy [10]. This approach also leads to the new polymer- intercalation method described here, namely the in situ oxidative polymerization-encapsulative precipitation. This new procedure has given several new PPY nanocomposites such as M082/PPY, WS2/PPY and 79 RuCl3/PPY nanocomposites. The former two nanocomposites will be presented in this chapter. The last one will be reported in Chapter 5. Since this procedure of PPY-nanocomposite preparation involves the reaction between pyrrole, FeC13 and exfoliated host materials, and unavoidably includes some side reactions such as the restacking of the host materials and the formation of the bulk polymer, the formation of nanocomposite products and their morphologies depends not only on the thermodynamics but also on the kinetics of these reactions. Therefore, reaction parameters such as the ratio of the reactants, the order of addition of reactants, reactant concentrations and temperature are all important. It will be seen in the following sections that the change of the pyrrole/FeCl3 ratio alone causes changes in the structure of M082/PPY nanocomposites, and affects the formation of WS2/PPY nanocomposites. Experimental Section 1. Reagents Pyrrole (98%), N-methyl pyrrole (99%), anhydrous FeCl3 (purified grade) and W82 (99.8%) were purchased from Mallinckrodt Chemical Inc., Aldrich Chemical Co., Fisher Scientific and Alfa Aesar respectively. Acetone was industry grade purchased from Quantum Chemical Company. Other chemicals were the same as in Chapter 1. Deionized water was used. 2. Preparation of M082/PPY N anocomposites a. Procedure 1: preparation with a controlled amount of pyrrole The M082/PPY nanocomposite was prepared by oxidizing pyrrole in an aqueous monolayer M082 suspension with FeCl3. In this particular procedure described here, pyrrole and exfoliated M082 were chosen as the 80 limiting reactants in the reaction. In a typical reaction, an amount of 0.20 g LiMoS2 (1.20 mmol) was exfoliated in 20 ml of H20. To this solution 0.016 g of pyrrole (0.24 mmol) dissolved in 5 ml of H20 was added. The mixture was stirred in an ice bath for 15 min and to it an aqueous solution of 0.156 g FeCl3 (0.96 mmol) was added dropwise. The reaction mixture was stirred in an ice bath for ~24 h. The product was collected by centrifugation as a black solid. It was washed with acetone and dried in vacuum. Elemental analysis, which was done by Oneida Research Services, Inc., Whitesboro, New York, showed a composition C6.59%, H 0.59%, N 1.31% and Cl 0.00%. The corresponding formula is (C4H3N)0.25(H20)0,16M082. (Calculated: C 6.70%, H 0.60%, N 1.95%.) 1111 Thermal gravimetric analysis (TGA) experiments in oxygen indicated a 1033 around 20.0 wt% at 650 0C, while the above formula suggests a loss of 19.6 wt% supposing that the residue is M003 1121. Energy dispersive X-ray microanalysis (EDS) indicated that there was some Fe compound existing in the product. The ratio of Fe to M0 was about 0.19:1. After having been washed with 1 M HCl, the ratio changed to 0.02:1. No chlorine was found by EDS. The ratio of pyrrole to M082 in the reaction above was 0.2210. In other reactions, the pyrrole to M082 ratio varied from 0.1 to 0.6 ( 0.1, 0.2, 0.25, 0.3, 0.4 and 0.6, etc.), while the ratio of FeCl3 to pyrrole was always held at 4:1. A pyrrole/M082 ratio of ~0.2—0.3 provided well ordered product. Similarly, corresponding M082/poly(N-methyl pyrrole) (M082/PMPY) nanocomposites were also prepared. Elemental analysis showed: C 7.63%, H 0.77%, N 1.38%; C 7.65%, H 0.77%, N 1.38% [131. This corresponds to a formula (C5H5N)0,23(H2O)0,12MoS2. (Calculated: C 7.65%, H 0.77%, N 1.78%.) A nanocomposite with this formula should 81 lose 20.1% of the weight assuming the residue is M003, which agrees with the TGA experiments of 21.6 wt% loss at 650 0C in air. b. Procedure 2: preparation with an excess amount of pyrrole In this procedure, FeCl3 was kept to be equal to or less than 0.5 equivalent to MoSz. In a typical reaction, an amount of 0.20 g LiMoSz (1.20 mmol) was exfoliated in 20 ml of H20. To this solution 0.32 g of pyrrole (4.8 mmol) dissolved in 40 ml of H20 was added. The mixture was stirred in an ice bath for 15 min and to it an aqueous solution of 0.098 g FeC13 (0.60 mmol) was added dropwise. The reaction mixture was stirred in an ice bath for ~24 h. The product was collected by centrifugation as a black solid, washed with water, dried in air for 1 day, and then dried in vacuum. Elemental analysis of a sample showed: C 12.64%, H 0.99%, N 3.19%; C 12.40%, H 0.94%, N 3.15%. The corresponding formula is (C4H3N)0,51(H20)0,13M0S2. (Calculated: C 12.46%, H 0.96%, N 3.64%.) A nanocomposite with the above formula should lose 26.7% of its weight and convert to M003. This agrees well with the TGA experiments which detected a loss around 26.4 wt% at 650 0C in air. MoSZ/PMPY nanocomposites were also prepared this way. Elemental analysis showed a composition: C 14.30%, H 1.37%, N 3.01%; C 14.22%, H 1.26%, N 2.95%. The corresponding formula is (C5H5N)0,475(H20)0,13MoS2. (Calculated: C 14.26%, H 1.32%, N 3.33%.) A nanocomposite with the above formula should lose 28.0 wt% on the basis that the residue is M003. This agrees with the TGA experiments, which showed a loss of 27.5 wt% at 650 0C in air. 82 3. Preparation of WS2/PPY N anocomposites LiWSz was prepared by reacting W82 with 3.0 equiv. of LiBH4 at 350 0C for 3 days, similar to LiMoS2 in Chapter 1. The PPY was encapsulated in WS2 with a procedure similar to the Procedure 2 used to prepare MoSZ/PPY. In a typical reaction, 0.40 g mixture of LiWS2/LiBH4/LiH (~1:2: 1) (about 1.34 mmol) was stirred in 20 ml of H20 for 2 h. Hydrated WS2 was collected by centrifugation, washed with 20 ml of fresh water, and suspended in 20 ml water by 30 min stirring. The resulting suspension was mixed with 10 ml water solution which contained 0.225 g pyrrole (3.35 mmol) and stirred in an ice bath. To this mixture, 10 ml of cold (~ 0—1 0C) aqueous solution of 0.109 g FeC13 (0.67 mmol) was added dropwise. The mixture was stirred for 24 h. The product was collected by centrifugation as a black solid, washed with water, dried in air for 1 day, and then dried under vacuum. Elemental analysis gave: C 4.71%, H 0.88%, N 1.17%; C 4.67%, H 0.87%, N 1.16%, corresponding to (C4H3N)0,27(H20)0,32WSZ [14], (Calculated: C 4.62%; H 0.87%; N 1.35%.) The elemental analysis of restacked WS2 which was washed by hydrochloric acid and had a d spacing of 6.3 A showed a result: C 0.30%, H 0.23%, N 0.07%; C 0.28%, H 0.21%, N 0.07. If the contents of C, H and N from the control experiment is deducted from the results of the elemental analysis of (PPY)XWS2, the suggested formula becomes (C4H3N)0,25(H20)0,53WS2. The formula above should lose 15.3 wt% if the nanocomposite is converted to W03, which does not match the TGA experiments [15] well. An EDS measurement detected only a trace of Fe existing in the product. The ratio of Fe to Mo was about 0.07:1. 83 4. Instrumentation The instrumentation in the measurements such as powder X-ray diffraction (XRD), thermal gravimetric analysis (TGA), differential scanning calorimetry (DSC), infrared (IR) spectroscopy and inductively coupled plasma (ICP) spectroscopy were the same as described in Chapter 1. Room temperature conductivity measurements were done on pressed sample pellets with a four-probe detector connected to a Keithley-236 source measure unit. Variable temperature direct-current electrical conductivity measurements and thermopower measurements were made on pressed pellets, by Dr. Kannewurf’s group at Northwestern University, with 60 and 25 um gold wires used for the current and voltage electrodes respectively. Measurements of the pellet cross-sectional area and voltage probe separation were made with a calibrated binocular microscope. Electrical conductivity data were obtained with a computer-automated system described elsewhere [16]. Magnetic susceptibility measurements were done with a Quantum Design MPMS2 SQUID magnetometer at the Department of Physics and Astronomy at MSU. Samples were sealed in low density polyethylene (LDPE) bags under a nitrogen atmosphere. The magnetic moments of the bags were acquired and deducted from the measurements. To obtain the paramagnetic susceptibility, the diamagnetic susceptibility and XTIP were subtracted from the total susceptibility. The former was derived by adding up the diamagnetic susceptibility of each component, which was obtained from the literature [17]. The latter was determined by adjusting its value by a try-and-error process so that the paramagnetism optimally fit Curie-Weiss law. 84 Results and Discussion 1. MoSz/PPY N anocomposites The synthesis of MoSz/PPY nanocomposites with a controlled amount of pyrrole and an excess amount of FeCl3 (Procedure 1) produces a single phase nanocomposite product with a basal spacing around 10.5 A, based on X-ray diffraction, see Figure 2.1(b). The interlayer expansion of ~4.3 A indicates the successful inclusion in the interlayer gallery of a guest species, which is obviously PPY, as supported by infrared spectra. The IR spectrum of the nanocomposite displayed in Figure 2.2 is of a sample from a reaction with relatively high pyrrole ratio, ~0.4. The nanocomposites produced from reactions using S 0.25 equivalent of pyrrole have IR spectra showing almost no peaks of organic species, due to the strong reflectivity of the samples originated from the conductive MoS2 layers. Heat treatment of the samples causes the structural transition of MOS; from the octahedral to the trigonal prismatic coordination (2H-MoS2), which alters the MoSz from a metallic to a semiconducting state. This reduces the infrared reflectivity of the MOS; layers, so that the absorbance of the intercalated species becomes visible. Figure 2.3 shows the IR spectrum of a heating treated (PPY)0,25(H20)0,16M0S2 sample, which had been produced with 0.2 equivalent pyrrole and was heated up to 300 0C. In both spectra in Figures 2.2 and 2.3, the observed peaks are related to PPY. To explore the possible nanocomposite phases and the optimum conditions to produce nanocomposites, the ratio of pyrrole to exfoliated MoS2 in the reaction was altered from 0.1 to 0.6. As summarized in Table 2.1, the products have (I spacings varying from 10.2 to 12.3 A. It seems that higher pyrrole ratios result in slightly higher expansions. Since the corresponding interlayer gallery expansion is between 4.0 and 6.1 A, 85 Intensity 6-2 A (a) restacked M082 -2 A. m A _J 10.5A (b) (PPY)O.25(H20)0.16MoS2 (by Procedure 1) 11.9A (C) (PPY)O.51(H20)0.18M 0 82 (by Procedure 2) 10-9 A (d) (PMPY)023(H20)012M082 (by Procedure 1) 12.7A (e) ........M082 (by Procedure 2) ""l""l"‘FT""I'T"l"" 10 20 30 40 50 60 20 (deg) Figure 2.1. XRD patterns of MoSz nanocomposites. 86 (PPY)n0 4"M032 /,.._..,... Irnpurity in KBr Transmittance 1550 \O O\ N '—¢ 777 <1- v—I N 00 v—I v—I \D .4 O\ C m C v—r I I I I I I I I I I I I I I r I I I I I I I I I I T I 2000 1800 1600 1400 1200 1000 800 600 Wavenumber (cm' 1 ) Figure 2.2. IR spectra of PPY and a MoS2/PPY nanocomposite prepared by procedure 1. Impurity in KIBr Intensity 8% ‘22 I I I I I I I 1 I I I I I I I I I I l I I T I I I I I I I I I 2000 1800 1600 1400 1200 1000 800 600 400 Wavenumber (cm'l) Figure 2.3. IR spectrum of a M082/PPY nanocomposite prepared by procedure 1 and heating treated twice up to 300 0C under N2 atmosphere at a heating rate of 5 0C/min. 87 Table 2.1. Parameters of (PPY)xMoS2 prepared with different pyrrole ratios pyrrole/MoS2 d spacing of coherence length *[18] ratio nanocomposites (A) (A) 0.1 10.2 40 0.2 10.9, 10.6, 10.5 102, 102, 124 0.25 10.7, 10.5 149, 131 0.3 11.2 89 0.4 12.3 52 0.6 11.8, 11.5 40, 47 * The X-ray scattering coherence length in a hkl direction is calculated from the peak broadening of the hkl reflection and is an index of the structural order in that direction. the space can only accommodate one molecular layer of PPY. It is likely that the pyrrole rings of the polymer lie almost parallel to the inorganic layers when the gallery height is low, while they rise more vertically when the gallery height is high. According to the molecular dimensions of the polymer, which will be discussed in detail in Appendix B, the 4.0 A height is appropriate for the flat orientation of the pyrrole rings. When the gallery is higher and the pyrrole rings tilt up, more polymer chains can be packed in the gallery. Since all pyrrole forms PPY upon oxidation with FeC13, the ratio of pyrrole to MoS2 in the reaction needs to be controlled. When this ratio is high, excess polymer forms, which cannot be accommodated as a single molecular layer in the gallery. As a result, it forms bulk polymer, which makes the product a mixture of phases. On the other hand, deficiency of pyrrole makes the gallery partially filled, which is also not an ideal situation because it gives rise to restacked MoSz. Furthermore, too much or 88 too little PPY hurts the lamellar structure of the nanocomposites. The structural order (Le. stacking order) in the nanocomposites is indicated by the coherence length [13] of the materials along the stacking direction. The coherence lengths of the products clearly show that the stacking of the layers is poor when the ratio of pyrrole is lower than 0.2 or higher than 0.3. When the ratio is 0.25, the reaction produces nanocomposites of the highest coherence length (i. e. sharpest diffraction peaks). The PPY nanocomposites can also be prepared using excess pyrrole and limited FeCl3 (Procedure 2). Contrary to when excess FeCl3 is used, adjusting the amount of pyrrole does not affect the coherence length of the nanocomposites significantly when the FeCl3 is limited or deficient. The resulting nanocomposites always have fairly high coherence length along the restacking direction. The basal spacing and the amount of polymer in the nanocomposites remains unchanged, i.e. ~11.9 A and ~0.5 repeat—unit per MoSz respectively. Figure 2.1(c) shows the powder X-ray pattern of a nanocomposite (PPY)O,51(H20)0.18M0S2, prepared with these conditions. The claim that the nanocomposites from Procedure 2 contain PPY rather than pyrrole is supported by the IR spectra, see Figure 2.4. In addition, it is supported by the fact that the nanocomposites cannot be de- intercalated. Pumping at 60 0C did not cause a shrinkage in the gallery spacing. In Procedure 2, the amount of PPY present in the product is more than what the FeCl3 alone could have produced. The greater amount of PPY is explained by the fact that ambient oxygen takes part in the reaction as an electron acceptor. It has been proven that the ambient oxygen can oxidize and polymerize pyrrole in the presence of FeCl3 [19]. The whole process can be described as follows: the addition of FeC13 causes or initiates polymerization of pyrrole producing radical cationic oligomers which 89 Transmittance IIIIIIIIIIIIIIIIIIIIIIIIIITIIII 2000 1800 1600 1400 1200 1000 800 600 400 Wavenumber (cm' 1‘) Figure 2.4. IR spectra of a nanocomposite (PPY)0.51(H2O)0,18MoS2 prepared by procedure 2. (a) vacuum dried at room temperature only, and (b) heat treated twice up to 300 0C under N2 atmosphere at a heating rate of 5 oC/min. 90 themselves are more reducing than the pyrrole monomer. The IR spectra of the (PPY)xMoS2 from both procedures show polymer stretching vibrations occurring at energies higher than those found in bulk PPY, prepared under similar experimental conditions (minus the MoS2 layers). The average blue-shift for most peaks is ~10 cm-1 and it reflects the lower molecular weight (MW) of the intercalated polymer. Similar observations were made in previous work where the MW of in situ intercalated polymers is always found to be considerably smaller than that of the corresponding bulk materials [13, 1b, 20]. This is a consequence of polymer growth under kinetically restricted conditions. As it has already been mentioned, the interaction among pyrrole, FeCl3 and exfoliated host materials could potentially give rise to a number of undesirable side reactions such as the restacking of the host materials and the formation of bulk polymer. The formation of the single phase nanocomposites is therefore remarkable. Explanations for this include the possible formation of a MoS2.pyrrole complex in solution before the addition of FeCl3, or more likely the strong electrostatic attraction between the negatively charged [MoS2]"' layers to the positively charged polymer chains. Since the polymer is produced in the doped form, an important issue to be addressed here is: what is the dopant anion in PPY? This has not been simple to resolve. There are several possible candidates for counter ions in (PPY)xMoS2, which include [FeC14]', Cl', [MoS2]"' and [OH]’. Elemental analysis found no chlorine in the nanocomposite sample, which excluded the possibility of [FeCl4]', and Cl'. It is possible that negatively charged [MoS2]"‘ acts as the dopant, as is the case with FeOCl and V205. Indirect 91 experiment data to be presented below are consistent with [MoSz]"' being the most likely counterion. 2. WS2/PPY Nanocomposites The WS2/PPY nanocomposites were prepared using excess pyrrole and limited FeCl3. The successful production of a PPY nanocomposite is indicated by the strong reflections in the XRD pattern, which has seven 001 reflections in the 20 range from 2° to 60°, see Figure 2.5. The basal spacing of the (PPY)0,25(H2O)0,53WS2, 11.0 A, corresponds to an interlayer expansion of 4.8 A which is comparable to that of (PPY)0,25(H2O)0,16M0S2 at 4.3 A. The IR spectra of the nanocomposite are shown in Figure 2.6. Contrary to the case of MoS2, single-phase products of WS2/PPY nanocomposites cannot be prepared with controlled amount of pyrrole and excess FeCl3 (Procedure 1). Reactions under those conditions always produced restacked WS2 mixed with the WS2/PPY products. In addition, the nanocomposites were much less ordered with very broad X-ray diffraction peaks. This dissimilarity between the exfoliated MoS2 and WS2 systems could stem from the small differences in the charge density of the layers and their affinity for the polymer. According to ICP analysis, the exfoliated layers carry a charge of 0.15 e'lW S2 [21], a little less than that of M082 layers, 0.18 e'lMoS2. This small charge difference does not seem enough to explain all the differences in reaction and nanocomposite composition. The observation in PPY intercalation reactions and the fact that much higher poly(ethylene oxide) quantities are needed to achieve a WS2/PEG nanocomposite phase [22], may suggest a lower affinity of WS2, than exfoliated MoS2, for organic polymers. 92 6.3 A (a) restacked W82 E? m A g 11.0/R 5 (b) (PPY)0.25(H20)0.53W82 W ""ll"llT"'l""l""|"" 0 10 20 30 4O 50 60 20 (deg) Figure 2.5. XRD patterns of WS2/PPY nanocomposites. / / / Impurity in KBr Transmittance 8 E / I I I I I I I I I I T I I I I If I I I I I I I I I f I I I I 2000 1800 1600 1400 1200 1000 800 600 400 Wavenumber (cm' 1) Figure 2.6. IR spectrum of (PPY)0,25(H2O)0,53WS2 heating treated twice up t0 320 0C under N2 atmosphere at a heating rate of 5 OC/min. 93 3. An Analysis of Polymer Arrangements inside the Interlayer Galleries Based on the Dimensions of the Polymer Molecules The structural parameters of the PPY nanocomposites and the poly(N-methyl pyrrole) (PMPY) nanocomposites, prepared with the two procedures are presented in Table 2.2. The temperatures for the octahedral to trigonal prismatic structural transition, which will be discussed in the next section, are also listed. The chosen nanocomposites prepared with Procedure 1 are those of good structural order, which have smaller (I spacings and lower polymer contents. Table 2.2. Structural data of PPY and PMPY nanocomposites sample reaction d spacing coherence Ttrans d spacing procedure (A) length (0C) after DSC (PPY)0,25(H2O)0.16M0S2 1 10.5 124 192 9 .6 (PPY)0,51(H2O)0,13M0S2 2 l 1.9 88 208 1 1.1 (PMPY)0,23(H2O)0_12MOSZ 1 10.9 95 198 9 . 9 (PMPY)0,475(H2O)0,13M0S2 2 12.7 155 216 12.3 As mentioned above, when the polymer content is low the pyrrole rings lie almost flat with respect to the inorganic layers, while they rise more vertically when the polymer content is high. An analysis of the dimensions of a PPY chain is necessary for a correct insight into the possible polymer chain arrangement in the galleries. According to the geometric calculation described in Appendix B, which is based on the bond lengths and angles, and the van der Waals' radii of the atoms, the PPY chain looks like a ribbon, which has a width of about 7.8 A. The thickness of the ribbon is about 3.5 A, which is decided by the van der Waals' diameter of C atoms. At the edges, the ribbon is a little narrower, 2.8 A, corresponding to the van der Waals' diameter of the H atoms. If the PPY 94 chains lie totally flat, the nanocomposite would have a d spacing of 9.7 A. This arrangement can accommodate a maximum of 0.31 equivalent of PPY. In (PPY)0,25(H2O)0,16MoS2 which has a d spacing of 10.5 A, the polymer chain can tilt at an angle of ~18° to the MoS2 layers. This tilt increases the maximum PPY content to 0.32 equivalent. The 0.25 equivalent content of polymer in the nanocomposite is less than this maximum value, and thus quite reasonable. In the nanocomposite with (I ~ 11.9 A, i.e. (PPY)0,51(H2O)0,18MoS2, the polymer chain can tilt ~36°. This relatively large tilt makes the partial overlap of the polymer chains possible, so the content of PPY can reach as high as 0.51 equivalent. MoS2/PMPY nanocomposites have also been synthesized with the two procedures used to synthesize MoS2/PPY analogs. It is hoped that a comparison of the PPY and PMPY nanocomposites will lead to a better understanding of the arrangement of polymers inside the galleries. A PMPY ribbon is a little thicker than a PPY ribbon due to the bulky methyl group. The thickness of the ribbon becomes about 4.9 A, a 1.3-1.4 A increase from PPY. The width of PMPY is about the same as PPY, 7.8 A. The two PMPY nanocomposites listed in Table 2.2 have d spacings around 10.9 and 12.7 A respectively. The interlayer gallery height of the first nanocomposite, 4.7 A, is just enough for a monolayer of PMPY to lie with its pyrrole rings parallel to the MoS2 layers. This arrangement of PMPY will let the MoS2 nanocomposite accept a maximum of 0.31 equivalent of polymer. The detected amount of polymer in the nanocomposite, 0.23 equivalent, is therefore a reasonable value. The interlayer gallery height of the second nanocomposite, 6.5 A, allows the PMPY chains tilt about 22°. This small tilt angle will not diminish the projection area that an N-methyl pyrrole ring produces. Therefore, the M082 will not accept, in its galleries, 95 an amount of PMPY as much as 0.475 equivalent, unless the PMPY chains partially overlap or arrange themselves in an inter-digitated fashion. 4. Thermal Properties As discussed in Chapter 1, the electronic structure of MoS2 changes after intercalation of lithium, corresponding to the structural transition from the trigonal prismatic MoS2 to the octahedral MoS2 [1d]. In the restacked MoS2, which preserves the octahedral coordination of M0, the layers are similar to the 1T-MoS2 which is metastable [23]. The MoS2 layers in (PPY)xMoS2 are, accordingly, also prone to conversion to the 2H-MoS2- like modification upon standing or with application of temperature or even pressure. The activation energy of this transition in restacked MoS2 is relatively small, ~53 kJ/mol [243], so the transition temperature measured by DSC is only 100 0C at a heating rate of 5 OC/min. The intercalation of polymers raises the activation energy as judged by the higher transition temperature. For example, the (PEO)1,0MoS2 has an activation energy 82.2 kJ/mol and a transition temperature 138 OC. The intercalation of PPY raises the transition temperature higher than any of the other polymers thus far. The value for (PPY)0,25(H2O)0,16M0S2 and (PPY)0,51(H2O)0,1gMoS2 are 185 OC and 208 0C respectively, as listed in Table 2.2. To determine the activation energy for the transition of (PPY)0,51(H2O)0,13M0S2, a series of DSC measurements was done under different heating rates, see Figure 2.7. Different heating rates provided different transition temperatures, although the heat of transition detected by DSC was always around 37 kJ/mol. The heating rate and the transition temperature are related by the Arrhenius equation [25], 96 4o:— 20 °C/min 228.7 °c 30 '— A : 10°C/min 221.2 °C 3 : E 20: 5°C/min 205.9 °C 0 3 o . 0 <8 - 2.5 C/mm 199.9 C 10 — :1°C/min 188.7°C O - I- " I I I I l I I I I L I I I I I I 1 60 1 80 200 220 240 260 Temperature (°C) Figure 2.7. DSC measurements at different heating rates for (PPY)0.51(H20)0.18M082- 3.5 3.0 2.5 2.0 1.5 1.0 0.50 0.0 -0.50 l. Ea = 138 kJ/mol ln(Rate) Ill‘IlIIlIllIlllIIII'IIIIIIle11IT1 O IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII I J I I I I I I I I 41 I I I I I I I I I I I I I 5 2.00 2.05 2.10 2.15 2.20 1000/T (l/K) \o IIII Figure 2.8. Plot of the DSC heating rate versus the temperature of the MoS2 structural transition for (PPY)0,51(H2O)0,13M0S2. 97 5 _ : 292 °C 4 _- e 33 5; 2:— 8 first cycle second cycle Q 1 - N \ C 0 E- ’ ’ ‘7‘— ’ \ / -1 :I I I I I I I I I I I I I I I J_I I I I I I I I I 141 I I I I I I O 50 100 150 200 250 300 350 Temperature (0C) Figure 2.9. DSC measurement for (PPY)0,25(H2O)0_53WS2. The measurement was done in 2 heating cycles up to 320 0C at a rate of 5 OC/rnin. The curve of the first heating cycle is labeled with the arrows. 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 _O I I I I I I I I I I I I I I I I 14 l L I I I 1.65 1.70 1.75 1.80 1.85 1.90 1000/T (1/K) Ea = 136 kJ/mol ln(Rate) IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIJIIIJLL Figure 2.10. Plot of the DSC heating rate versus the WS2 structural transition temperature in (PPY)0,25(H2O)0,53WS2. 98 Rate = A >1< C'Ea/RT where A is the pre—exponential coefficient, R the gas constant and E3 the activation energy. The activation energy can be calculated from the slope of the ln(Rate)~1/T plot. The plot for (PPY)0,51(H2O)0,13MoS2 gives an activation energy of 138 kJ/mol (see Figure 2.8), which is much higher than those of restacked MoS2 or (PEO)1,oMoS2. The large increases in the transition temperature and activation energy may result from the reducing nature of PPY, thus stabilizing [MoSz]x'. The same situation exists in restacked WS2 and (PPY)XWS2 nanocomposite, but the transition temperatures are higher than MoS2 compounds. The values for restacked WS2 and (PPY)0,25(H2O)0,53WS2 are 207 0C [26] and 292 0C respectively. A DSC measurement for (PPY)0,25(H2O)0_53WS2 is shown in Figure 9. The change of the transition temperature, by 85 0C, is the same as in the case of (PPY)0,25(H2O)0,16M0S2. Under heating rates of 1.0, 2.5, 5.0, 10.0 and 20 OC/min, the transition temperatures of (PPY)0,25(H2O)0,53WS2 detected by DSC were 265.3, 280.9, 292.1, 307.0 and 324.5 0C respectively. The activation energy of the transition calculated from the ln(Rate)~1/T plot (see Figure 2.10) is 136 kJ/mol, much higher than the value of restacked WSZ, 82.4 kJ/mol [26], but close to that of (PPY)0,51(H2O)0,13MoS2. The heat of transition detected by DSC was around 20 kJ/mol, slightly higher than the 18 kJ/mol of restacked WS2, but less than 37 kJ/mol found for (PPY)0.51(H20)0.18M052- After the transition, samples of these nanocomposites still have well ordered lamellar structure. The d spacings of the M082 nanocomposites after the DSC experiments are listed in Table 2.2. The d spacing of the 99 WS2/PPY nanocomposite after the DSC measurement is 10.3 A. The IR spectra of the heated nanocomposites show only PPY peaks. 5. Charge Transport Properties Room temperature electrical conductivity values of the materials described here are listed in Table 2.3 [27]. The conductivities of the restacked WS2 and the fresh restacked MoS2 are high, about 100 and 12 S/cm respectively, as measured with pressed pellets. The conductivity of restacked MoS2 diminishes quickly, due to the gradual structural transition of restacked MoS2 at room temperature. According to experiments, the transition proceeded almost completely in about 4 months and the conductivity reached about 0.07 S/cm. The conductivity stayed at about 0.06 S/cm for the subsequent 3 years. Comparing the samples more than 6 months old, the intercalation of PPY raises the conductivity more than 10 times. A higher PPY content gives the material a higher conductivity. In contrast, the intercalation of PPY reduces the conductivity of restacked Table 2.3. Room temperature electrical conductivities of PPY nanocomposites and restacked M082 and WS2 sample R.T. conductivity (S/cm) restacked MoS2 (1 day) 12 restacked M082 (3 years) ~0.06 (PPY)0.25(H20)0.16M032 0.8 (PPY)0.51(H20)0.18M032 2.5 restacked WS2 (10 days) 100 (PPY)0.25(H20)0.53W32 27 100 WS2, because the conductivity of bulk PPY is only 13 S/cm, smaller than that of restacked WS2. In addition, (PPY)XWS2 may have more grain boundaries giving resistance. Variable-temperature electrical conductivity measurements for pressed (PPY)0,25(H2O)0,16M0S2 pellets indicate a considerable higher conductivity with respect to pristine MoS2, as shown in Figure 2.11. In the temperature range 50-300 K, the material exhibits weak, thermally activated behavior, which is common in conductive granular nanocomposites. The plot of conductivity versus reciprocal temperature does not follow a straight line, which indicates that the thermal activation is more likely through the boundary areas between the conductive domains. The conductivity is similar in magnitude to that of (PANI)xMoS2 consistent with the similar nature of the materials. The corresponding thermoelectric power measurements show relatively large and positive Seebeck coefficient values which indicate a p-type conductor, see Figure 2.12. The relatively large values and the curved thermoelectric power plot indicate a deviation from an ideal metallic behavior. Similar to (PPY)0,25(H2O)0_16M0S2, (PPY)0,25(H2O)0,53WS2 has also a thermally activated electrical conductivity. A variable temperature conductivity measurement for (PPY)0.25(H2O)0,53WS2 is shown in Figure 2.13 together with that for restacked WS2. The comparison demonstrates that (PPY)0.25(H2O)0,53WS2 is not as electrically conductive as the restacked WS2, which has been seen in the room temperature conductivity measurements. Thermoelectric measurements for (PPY)0,25(H2O)0,53WS2 show typical p—type metallic behavior, with small positive values which decreases linearly with falling temperatures, see Figure 2.14. The restacked WS2 has similar thermoelectric behavior [28]. 101 10" (E) 10:3 (H2O)0.16M082 @ 3* IE 105 O O 3 o :3 "O 8 o 10‘7 A, 9 m. 1 1 1 . . . 1 1 1 . . 1 1 1 .A. 1 1° 0 5 10 15 20 1000/T (K'l) Figure 2.11. Variable temperature electrical conductivity measurements on pressed pellets for pristine MoS2 and a MoS2/PPY nanocomposite prepared with Procedure 1. 100 _ SE 90 '— 5‘ Z :1. : V 80 — I... .— d.) .. 3 : 8, 70 :- é _ _g 60 _— H I. 50 l I_I_I I I I I I I I I I; I l l I I I I I l I 50 100 150 200 250 300 Temperature (K) Figure 2.12. Thermoelectric power measurement on a pressed pellet for (PPY)0,25(H2O)0,16M0S2 prepared with Procedure 1. 102 101 5 a \mmmmo restacked WS2 ”E? *’ C>00 o o o o o o o o o q o 0 _ _. a 10 g g b F 3 :E C o (PPY)025(H20)0.53WS -1 _ ° _ g 10 E O o o E U E o o 3 10-2 J g1 I I I l I I I I I I I I I J I I I I I I I O 10 20 30 40 50 1000/r (K'l) Figure 2.13. Variable temperature electrical conductivity measurements on pressed pellets for restacked WS2 and (PPY)0_25(H2O)0,53WS2. 5 .- 9 4:— ; - :1. I ‘ v 3 __ H .— qg - g : cu 2 f a _ : o g 1 .— OOWO E-‘ : O - I I l I l I l I I l l l I I I I l J I l l I l I 100 150 200 250 300 Temperature (K) Figure 2.14. Thermoelectric power measurement on a pressed (PPY)0.25(H20)0.53W32 pellet. 103 Concluding Remarks MoS2/PPY, WS2/PPY and MoS2/PMPY nanocomposites have been successfully synthesized using the new in situ oxidative polymerization- encapsulative precipitation reaction. This reaction is kinetically controlled, so some optimization to find the optimum conditions to produce the nanocomposites, or conditions to get the most ordered structure, is crucial. A great effect has been made in this research to optimize the reactions and the best conditions are reported and discussed. The MoS2/PPY, WS2/PPY and MoS2/PMPY nanocomposites synthesized have ordered structures and. large coherence lengths along the stacking direction. The new nanocomposites were characterized and identified. A discussion about the arrangement of polymer chains inside the MoS2/PPY nanocomposites suggests that the polymer chains lie almost flat inside the galleries when the polymer content is about 0.25 equivalent. The chains tilt up and overlap when the polymer content becomes about 0.5. The intercalation of PPY stabilizes the metastable restacked MoS2 and WS2 layers, by raising structural transition temperatures to 208 and 292 0C respectively. These increases in transition temperature are the largest ones observed yet in M082 and WS2 nanocomposites. Both MoS2/PPY and WS2/PPY nanocomposites are good conductors. The development of a synthetic strategy to intercalate an intractable, infusible polymer in non-oxidizing hosts is the most important achievement in this research. This is an advance from prior work where (a) an oxidizing host was required to proceed polymer intercalation, (b) a soluble polymer was necessary for insertion into a non-oxidizing host, or (c) a monomer intercalation was required followed by a second step of polymerization. The methodology developed here can be applied further to 104 other intractable polymers such as polythiophene. This development broadens the field of organic/inorganic nanocomposite materials as it enables interesting new inorganic/polymer combinations for study and development. 105 Appendix B Based on the conjugated structure of PPY, we assume that it is a flat or nearly flat molecule. As shown in Figure 2.15(a), the distance between the farthest two atoms in the direction perpendicular to the polymer chain (two H atoms) is 2.48 A * 2 = 4.96 A. By adding the van der Waals radius of H atom, 1.4 A, the width of the PPY chain should be 7.8 A. The thickness of the chain should be about 3.54 A, which is deduced from the wan der Waals radius of the C atom, 1.77 A. The length of each pyrrole unit is 3.59 A, so the area that each pyrrole unit occupies is 3.59 A * 7.8 A = 27.9 A2 [291. The area of each MoS2 unit is 8.653 A2, according to the a and y of the 2H—MoS2, 3.161 A and 120° [30]. This indicates that the MoS2 nanocomposite can accept a maximum of 0.31 equivalent of PPY in the galleries if the PPY chains lie flat. In the MoS2 nanocomposite of a d spacing 10.5 A, where the gallery is 4.3 A high, the polymer chain can tilt at an angle of ~18° to the M082 layers according to the following calculation: arcsin[(4.3A-2*l.4A)/(7.8A-2*1.4A)] = 17.50 1.4 A is the van der Waals radius of the H atoms which are on the two sides. When a PPY chain tilts 18°, the area of a projection of a pyrrole unit is (4.96 A * cos18° + 2 * 1.4 A) * 3.59 A = 27.0 A2. The M082 can accept a maximum of 0.32 equivalent of PPY arranged this way. When the expansion is 5.7 A, the polymer chains can tilt ~36°: arcsin[(5.7A-2*1.4A)/(7.8A-2*1.4A)] = 355° 106 The projection of a pyrrole unit in a such placed chain will have an area of (4.96 A * cos36° + 2*1.4 A) >t= 3.59 A = 24.4 A2. The PPY content can reach a maximum of 0.35 equivalent if there is no overlapping of the chains. The high PPY content in the nanocomposite, 0.51, can be best explained by the partial overlapping of the PPY chains in the galleries. Since the edge of the PPY chain is only ~2.8 A, the 5.7 A gallery height is just enough for the edges of two chains to overlap. In the case of PMPY, the bulky methyl group makes the PMPY chain thicker, 4.87 A, as obtained in the following calculation [also see Figures 2.16(b) and (c)]: 2*(C-H * sin110° + 1.4 A): 2*(110 .1. sin110° + 1.4 A) = 4.9 A. Again, 1.4 A is the van der Waals radius of a H atom. The (PMPY)0_23(H2O)0_12M0S2 has a d spacing of 10.9 A. Its 4.7 A gallery spacing is just enough for the PMPY chains to lie flat inside. The width of the PMPY chains is the same as that of PPY chains, because the added methyl groups are buried inside the grooves of the PPY chain, see Figure 2.15(b). If the PMPY chains lie flat, each N-methyl pyrrole unit will cover an area approximately the same as a pyrrole unit. Correspondingly, the MoS2 nanocomposite will accommodate a maximum of 0.31 equivalent of PMPY. The PMPY chains must tilt in (PMPY)0,23(H2O)0_12M0S2, which has a d spacing of 12.7 A. The tilt of the chains is calculated by the following approach, which is different from the calculation for PPY. As shown in Figure 2.16(b), the distance between the diagonal H atoms in the two methyl groups on the two sides, d6, is 5.27 A. When a PMPY chain lies flat, the d6 direction of the chain tilts by an angle of arcsin(2.07 A/5.27 A) = 23.1°, see Figure 2.16(c). In a gallery 6.5 A high, the chain can tilt so 107 that the d6 vector tilts a maximum angle of arcsin[(6.5 A - 2* 1.4 A) / 5.27 A] = 446°. The tilt of the chain is the change of the angle of the diagonal plane, 21.5°, see Figure 2.16(d). The 21.5° tilt of the PMPY chain produces a larger projection on the M082 plane, as shown in Figure 2.16(d). The PMPY chains in the nanocomposite must partially overlap to accept a 0.475 equivalent of polymer. Otherwise, according to the calculation, the polymer content would not exceed 0.29 equivalent. The 6.5 A gallery height does not exceed the sum of the thicknesses of the methyl group and H atom, 4.9 and 2.8 A respectively. In order to achieve a partial overlap with nearby polymer chains, a polymer chain must match its edge H atoms or methyl groups with those of the neighbor chains in an inter-digitated fashion. 108 (a) polypyrrole \i xi :1:—-z- \E \f: /i I—Z . \l l! \/: .. a 5 8f \: g =.\ : F8 sea—4+ I. 1 Length of covalent bonds: C-C 1.54 A, C=C 1.34 A, C-N 1.47 A, C=N 1.27 A, N-H 0.98 A, OH 1.10 A. d 1 = N-H + [(C-N + C=N) =1: sin36°] / 2 - [(C-C + C=C) .1 sin 18°] /4 = 1.56 A d2 = C-H * sin126° + [(C-C + C=C) * sin108°] / 2 + [(C-C + C=C) >1: sin 18°] 74 = 2.48 A d3 = 2 ... [(C-N + C=N) * cos36°] / 2 + [(c-c + C=C) .. cos 18°] /2 = 3.59 A (b) poly(N-methyl pyrrole) \ / ° \ / 070% ------------ I" ’///:—/—\\/1’\////:\\\\/‘36/18 ...... din T M i T \n/ ,2 WC\H H \H WC\H / 126\ ___________________ d1 = OH 1. sin20° + C-N + [(C-N + C=N) * sin36°] / 2 — [(C-C + C=C) =t< sin 18°] /4 = 2.43 A d2 = 2.48 A d3 = 3.59 A Figure 2.15. Schematic structures and dimensions of PPY and PMPY. 109 (a) Front view of poly(N-methyl pyrrole) (c) Contour of a flatly lay poly(N-methyl pyrrole) Figure 2.16. Views of PMPY. c/H (b) Side view of poly(N-methyl pyrrole) :‘ d5 '1 d4=2* dl =2*2.43A=4.86A d5 = 2 *1.10A *sin110°=2.07 A d6 = (d42 + d52)”2 A = 5.27 A (d) Tilted poly(N-methyl pyrrole) in a 6.7 A gallery o‘ ”’ 4 6.5 A ‘ 3.7 A /' I" x. 110 References 1 (a) M. G. Kanatzidis, L. M. Tonge, T. J. Marks, H. O. Marcy and C. R. Kannewurf, J. Am. Chem. Soc. 1987, &, 3797. (b) M. G. Kanatzidis, C.-G. Wu, H. O. Marcy and C. R. Kannewurf, J. Am. Chem. Soc. 1989, 111, 4139. (c) L. F. Nazar, X. T. Yin, D. Zinkweg, Z. Zhang and S. Liblong, Mat. Res. Soc. Symp. Proc. 1991, 2_10_, 417. (d) R. Bissessur, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, J. Chem. Soc., Chem. Commun. 1993, 687. (e) M. G. Kanatzidis, R. Bissessur, D. C. DeGroot, J. L. Schindler and C. R. Kannewurf, Chem. Mater. 1993, _5_, 595. (f) G. Matsubayashi and H. Nakajima, Chem. Lett. 1993, 31. 2 F. Croce, G. B. Appetecchi, L. Persi and B. Scrosati, Nature 1998, fl, 456. 3 (a) L. F. Nazar, H. Wu and W. P. Power, J. Mater. Chem. 1995, _5_, 1985. (b) F. Leroux, B. E. Koene and L. F. Nazar, J. Electrochem. Soc., 1996, m, L181. (c) T. A. Kerr, H. Wu and L. F. Nazar, Chem. Mater. 1996, 8, 2005. (d) B. E. Koene and L. F. Nazar, Solid State Ionics 1996, 8_9_, 147. (e) K. Ramachandran and M. M. Lerner J. Electrochem. Soc. 1997, L44, 3739. (f) F. Leroux, G. Goward, W. P. Power and F. Nazar J. Electrochem. Soc., 1997 , 144, 3886. 4 (a) J. A. Wilson and A. D. Yoffe, Adv. Phys. 1969, 18, 193. (b) L. F. Mattheiss, Phys. Rev. B 1973, 8, 3719. 5 F. R. Gamble and T. H. Geballe, in N. B. Hannay ed., Treatise on Solid State Chemistry, V013, Plenum Press 1976, pp 89. 6 (a) R. H. Friend and A. D. Yoffe, Adv. Phys. 1987, x, 1. (b) W. R. McKinnon, in A. P. Legrand and S. Flandrois ed., Chemical Physics of Intercalation (NATO ASI Ser., B: Phys, Vol. 1_72, Plenum Press, New York 1987, pp 181. 7 (a) C.—H. Hsu, M. M. Labes, J. T. Breslin, D. J. Edmiston, J. J. Winter. H. A. Leupold and F. Rothwarf, Nature, Phys. Sci., 197 3, 246 (155), 122. (b) V. M. Chapela and G. S. Parry, Nature, 1979, 2_8_1, 134. (c) W. M. R. Divigalpitiya, R. F. Frindt and S. R. Morrison, J. Mater. Res. 1991, _6_, 1103. (d) E. Ruiz-Hitzky, R. Jimenez, B. Casal, V. Manriquez, A. S. Ana and G. Gonzalez, Adv. 111 10 11 12 13 14 15 Mater. 1993, 5, 738. (c) J. P. Lemmon, J. Wu, C. Oriakhi and M. M. Lerner, Electrochim. Acta 1995, Q , 2245. (f) C. O. Oriakhi, R. L. Nafshun and M. M. Lerner, Mater. Res. Bull. 1996, Q, 1513. (a) L. W. Shacklette, R. R. Chance, D. M. Ivory, G. G. Miller and R. H. Baughman, Synth. Met. 1979, 1, 101. (b) M. Satoh, H. Yageta, K. Amano and E. Hasegawa Synth. Met. 1997 , 8:4, 167. (c) J. Cao, A. J. Heeger, J. Y. Lee and C. Y. Kim Synth. Met. 1996, 82, 221. (d) H. Kuhn, R. Gregory and W. Kimbrell, Int. SAMPE Electron. Conf. 1989, 3, 570. (e) M. D. Imisides, R. John, P. J. Reiley and G. G. Wallace, Electroanalysis 1991, 3, 879. E. P. Giannelis, V. Mehrotra, O. Tse, R. A. Vaia and T. -C. Sung, Mat. Res. Soc. Symp. Proc. 1992, E, 969. (a) Y. Fukushima and S. Inagaki, J. Inclusion Phenom. 1987, _5_, 473. (b) Y. Fukushima, A. Okada, M. Kawasumi, T. Kurauchi and O. Kamigaito, Clay Minerals 1988, Q, 27. Because the polymer exists inside the nanocomposite, it could experience somewhat incomplete oxidation in the process of elemental analysis experiments. This may lead to lower N amount detected. The formula is derived based on the amounts of C and H. An elemental analysis of another sample done in our department at MSU showed: C 7.28%, H 0.85%, N 1.40%; C 7.19%, H 0.82%, N 1.36%. The corresponding formula is (C4H3N)0,23(H2O)0,35MoS2. Calculated: C 7.28%; H 0.83%; N 2.12%. A nanocomposite with the formula above should lose 22.0% of the weight if the final residue is MoO3. This agrees well with the result of the TGA experiment in oxygen which had a loss around 21.0 wt% at 650 0C. This measurement and elemental analysis measurements hereafter were done by Dr. R. Huang in this department. Another analysis showed C 4.11%, H 0.71% and N 1.19%, which suggests a formula (C4H3N)0,235(H2O)0,62WS2: C 4.11%; H 0.71%; N 1.20%. Two T GA experiments in oxygen had losses around 11.5 wt% and 11.7 wt% up to 650 0C. 112 16 17 18 19 20 (a) B. N. Diel, T. Inabe, J. W. Lyding, K. F. Schock, Jr., C. R. Kannewurf and T. J. Marks, J. Am. Chem. Soc., 1983, 1Q, 1551. (b) J. W. Lyding, H. O. Marcy, T. J. Mark and C. R. Kannewurf, IEEE. Trans. Instrum. Meas. 1988, if, 76. (a) P. W. Selwood, Magnetochemistry (2nd Ed.), Interscience Publishers, New York, 1956, pp 78. (b) R. S. Drago, Physical Methods for Chemists (2nd Ed.), Saunders College Publishing, Philadelphia/San Diego/New York, 1992, Chapter 11. (c) For polyaniline: A. J. Epstein, J. M. Ginder, A. F. Richter and A. G. MacDiarmid, in L. Alcacer ed., Conducting Polymers, D. Reidel Publishing Company, Dordrecht, 1986, pp 121. X-ray scattering coherence lengths were calculated from the Scherrer formula Lhk1=KMBcosez (a) D. M. Moore and R. C. Reynolds, Jr., X-Ray Diffraction and the Identification and Analysis of Clay Minerals, Oxford University Press: Oxford/Newyork, 1989, pp 83. (b) H. P. Klug and L. E. Alexander, X-ray Diffraction Procedures for Polycrystalline and Amorphous Materials, John Wiley & Sons, New York, 1962, pp 491. In the formula, K is a constant close to 1. It is assigned 0.9 in the calculation. A is the X-ray wavelength, which is 1.5419A. B is the broadening of the reflection peak (in 20) in the XRD pattern, and is calculated according to the following equation: 9 = [b2 _ 1302 b is the width of the peak at a half height. b0 is the instrumental broadening, which is 014° according to the width of pristine 2H- MoS2. The unit of [3 is in are rather than degrees, so there is a factor of 7t/180. N. Toshima and O. Ihata, Synth. Met. 1996, B, 165. (a) M. G. Kanatzidis, C.-G. Wu, H. O. Marcy and C. R. Kannewurf, Adv. Mater.1990, 2, 364. (b) M. G. Kanatzidis, H. O. Marcy, W. J. McCarthy, C. R. Kannewurf and T. J. Marks, Solid State Ionics, 1989, 32-33, 594. (c) C.-G. Wu, M. G. Kanatzidis, H. O. Marcy, D. C. DeGroot and C. R. Kannewurf, Polym. Mat. Sci. Eng. 1989, 61, 969. (d) M. G. Kanatzidis, C.-G. Wu, H. O. Marcy, D. C. DeGroot, C. R. Kannewurf, Chem. Mater. 1990, _2_, 222. 113 21 22 23 24 25 26 27 28 29 30 Freshly restacked WS2 was carefully prepared for the inductively coupled plasma (ICP) analysis by stirring and washing the LiBH4, LiH and LiWS2 mixture in water 6 times, similar to the procedure used for the LixMoSZ ICP sample described in Chapter 1. ICP showed Li 0.383 wt% and S 23.0 wt%, which suggest a formula L101 5MoS2. The amount of W was not analyzed because of the lack of W standard. H.-L. Tsai, L. Wang and M. G. Kanatzidis, paper in preparation. F. Wypych and R. Schollhom J. Chem. Soc., Chem. Commun. 1992, Q, 1386. (a) R. Bissessur, Synthesis and Characterization of Novel Intercalation Compounds of Molybdenum Trioxide and Molybdenum Disulfide, Ph. D. Dissertation, Department of Chemistry, Michigan State University, 1994, pp 186. (b) idib, pp 183 and 193. Annual Book of ASTM Standards 1979, E-698, pp 601-608. H.-L. T sai, J. Heising, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Chem. Mater. 1997, 9, 879. The room temperature electrical conductivity data listed in Table 3 were measured at Michigan State University. Although the data are a little different from the room temperature values of the variable temperature conductivity measurements, because of the differences in sample, instrument and manipulation, the set of data in this table should be comparable among themselves. J. Heising, Synthesis and Characterization of Novel Intercalation Compounds of Molybdenum Trioxide and Molybdenum Disulfide, Ph. D. Dissertation, Department of Chemistry, Michigan State University, 1999. In the calculation to obtain the value 27.9 A2, 7.76 A rather than 7.8 A was used to avoid too much approximation and too few significant figures. 2H-MoS2 has a hexagonal structure (P63/mmc): a=3. 161 A, c=12.295 A; a MoS2 unit occupies 8.653 A2: P. Villars and L. D. Calvert, 114 Pearson 's Handbook of Crystallographic Data for Intermetallic Phases, 2nd Ed., Vol. _4, 1991, ASM International, Materials Park, OH 44073, pp 4445. (cited from: B. Schonfeld, J. J. Huang and S. C. Moss, Acta Crystallogr., Sec. B, 1983, _3_9__B, 404.) 115 Chapter 3 LAMELLAR TaSz/POLYMER NAN OCOMPOSITES THROUGH ENCAPSULATIVE PRECIPITATION OF EXFOLIATED LAYERS Introduction Two-dimensional inorganic/polymeric nanocomposites represent an important and growing class of hybrid materials with interesting physical properties. The research and development activity is concentrated mainly in two major fields, nanocomposites as structural materials [11 and as electro-active materials [2]. To synthesize two-dimensional nanocomposites, the insertion of polymers inside the layered hosts is the critical step. Polymers can be sandwiched between the galleries of the layered materials through one of four currently known approaches: (a) by monomer intercalation and subsequent polymerization in the galleries, (b) by in situ intercalative polymerization, (c) by direct polymer insertion and (d) by encapsulative precipitation of polymers from solutions of exfoliated lamellar solids. The latter two methods produce nanocomposites with preformed and well characterized polymers, and are capable of producing large varieties. The direct polymer insertion approach works well with organically modified aluminosilicates, or clays 11]. On the other hand, the encapsulative precipitation method, which has been developed in our laboratory, has been successful in making nanocomposites with electro- active layered hosts [31. In this method, the layered host exfoliates to give monolayers in solution, which interact with dissolved polymer chains and encapsulate them during a restacking process. This approach has led to the synthesis of MoS2 [3a], M003 [3b] and NbSe2 [3C] nanocomposites with various 116 polymers in our group, M003/poly(ethylene oxide) (MoO3/PEO) nanocomposites in Nazar’s group [4], and MoS2, MoSe2, TiS2, TaS2, M003 and MPS3 nanocomposites with PEO [53] and/or polyethylenimine (PEI) [5b] in Lerner’s group. Here we report the synthesis of TaS2/polymer nanocomposites through the encapsulative precipitation method [6’ 7]. The encapsulation of PEO, PEI and poly(vinylpyrrolidinone) (PVP) into TaS2 were examined in detail and the nanocomposites were characterized by a large variety of techniques. Most of all, our discovery about the different characters of various forms of LixTaS2 in exfoliation and polymer intercalation is extremely valuable to the synthesis of intercalative TaS2 complexes. The difference between PEI and other polymers, and the subtleties of the exfoliation and encapsulation of exfoliated TaS2 will be elucidated in this chapter. The exfoliation of TaS2 has long been known [8» 9» 101 and cations of many sizes have been encapsulated between the layers of this material. Early work in this area showed that electrolytically prepared HxTaS2 could be monodispersed in aqueous surfactant solutions and formed adsorption complexes with cationic dyes [8]. Later, chemically prepared NaxTaS2 was dispersed in water or NMF/water mixtures and cluster cations were included inside to form ordered intercalation compounds [10b]. Cation intercalation is driven more or less by cation exchange, which does not require an exfoliated solid [11]. Neutral polymer intercalation, however, needs a different driving force, for example, polymer-LP interactions and/or polymer-TaS2 van der Waals interaction. The intercalation of polymers most probably takes advantage of the affinity of 117 the polymer chains for the interlayer surface of the host and for the cations (i.e. LP“) in the interlayer galleries. Given that some metal dichalcogenides become superconductors at low temperatures, it is intriguing to consider polymer nanocomposites containing superconductive components. Such materials would combine the superconducting properties of inorganic solids with the processible properties of polymers giving rise to new forms of superconductors such as polymer matrix-based wires and free standing films, thus enabling new kinds of applications. We have made a first step in this direction by inserting polymers into NbSe2 and TaS2 to produce lamellar inorganic/polymer superconducting solids with plastic-like characteristics. This work is an outgrowth of our studies of intercalative polymer nanocomposites of MoS2 [38] using the exfoliation procedure [12]. Remarkably, the flexible metallic TaS2/polymer nanocomposites display bulk superconductivity. Experimental Section 1. Reagents PEO (5,000,000), PEO (100,000), PEI (25,000) and PVP (10,000) were purchased from Aldrich Chemical Company, Inc. After the polymers were dissolved, the polymer solutions were filtered to purify from insoluble polymer residues. LiBH4 (95%), LiOH-H2O (98%) and Ta (99.9%, 325 mesh) were also purchased from Aldrich. Sublimed sulfur was from Spectrum Chemical Mfg. Co. Anhydrous ether (99.0%), acetonitrile (99.5%) isopropanol (99.9%) and carbon disulfide (100%) were from Columbus Chemical Industries Inc., EM Science Inc., 118 Mallinckrodt Chemical Inc. and J. T. Baker Inc. respectively. No further purification was applied to the chemicals above. The water was distilled and degassed by bubbling nitrogen through it for 30 min before use. 2. Synthesis of 2H-TaS2 2H-TaS2 was synthesized with a modified literature procedure [131. An amount of 3.619 g of tantalum (20 mmol) and 1.300 g of sulfur (40.5 mmol) sealed in a quartz tube was heated at 450 0C for 12 h and then 950 0C for 36 h. The quartz tube was 13 mm in diameter and ~13 cm long with a volume of about 12-14 ml. The tube was quenched from 950 CC in cold water and excess sulfur was deposited on the tube walls. The lT-TaS2 formed was ground and washed with CS2 to remove any excess sulfur. The purified lT-TaS2 was transformed to 2H-TaS2 by an annealing procedure of slow cooling from 910 to 450 0C in two weeks, according to the protocol shown in Table 3.1. Table 3.1. The annealing procedure to produce 2H-TaS2 temperature (0C) 910 750 650 550 450 time to reach (h) 12 16 20 20 20 time to hold (h) 12 24 48 72 72 3. Synthesis and Exfoliation of LixTaS2 An amount of 4.00 g 2H-TaS2 was reacted with n equivalents of LiBH4 (n = 0.1, 0.2, 0.3, 0.4, 0.5, and 1.0) in 100 ml of anhydrous ether for 3 days under a nitrogen atmosphere to obtain LixTaS2. LixTaS2 was 119 black when n = 0.1, 0.2 and 0.3, while it was reddish brown when n 2 0.4 [14], LixTaS2 was exfoliated in degassed water, in a concentration of 1.0 g/L, by 30 min of sonication in an ultrasonic cleaner under a nitrogen atmosphere. When n = 0.1, not much of LixTaS2 was exfoliated in water. The suspension was slightly yellowish black. When n = 0.2 or 0.3, most of the LixTaS2 went into water. Especially in the case of n = 0.2, the amount of unexfoliated LixTaS2 was very small (~2.6%) and the colloidal suspension had an intense greenish yellow color. When n 2 0.4, the LixTaS2 became increasely difficult to exfoliate. The color of the resulting suspension ranged from brownish yellow to reddish brown. The higher the value of n, the lighter the color of the suspension. Therefore, LixTaS2 from reactions with n=0.2 was chosen as the host material for polymer intercalation. Lio2TaS2 is used to present this form of LixTaS2 for convenience [15]. 4. Encapsulative Precipitation of Polymers In a typical reaction, 0.40 g of Li02TaS2 was exfoliated in 400 ml of degassed H2O by 30 min of sonication. The resulting suspension was mixed with 100 Hi] of polymer solution (5 or 10 times in excess by equivalents of repeat units) and stirred for 2 days under a nitrogen atmosphere. The supernatant of the reaction mixture was centrifuged. A half part of the polymer intercalated TaS2 was separated from the solution by centrifugation and is referred to as batch-I nanocomposite. The centrifuged supernatant was pumped to remove most of the water. A corresponding solvent (acetonitrile for PEO and isopropanol for PVP and PEI) was added to the concentrated supernatant to precipitate the intercalated TaS2 120 remaining in the supernatant. This precipitate is referred to as batch-II nanocomposite. The products were pumped to dryness and vacuum-sealed in glass ampules. 5. Instrumentation and Measurements a. Instrumentation The instrumentation in the measurements such as thermal gravimetric analysis (TGA), differential scanning calorimetry (DSC) and infrared (IR) spectroscopy were the same as described in Chapter 1. Room temperature conductivity measurements, variable temperature direct— current electrical conductivity measurements and thermopower measurements, and magnetic susceptibility measurements were conducted as described in Chapter 2. Generally, X-ray powder diffraction (XRD) patterns were collected as described in Chapter 1. For one-dimensional electron density calculations, we obtained XRD data (in the region 2° S 20 S 135°) from highly oriented samples and a stepwise scanning mode with 0.1° per step. Variable temperature solid state 7Li NMR spectra were taken on a 400 MHz Varian Nuclear Magnetic Resonance Instrument. Samples were loaded in a glove box under a nitrogen atmosphere. Scanning electron microscopy (SEM) and energy dispersive X-ray microanalysis (EDS) were done with JEOL-JSM 35 CF microscope at an accelerating voltage of 15 kV and 20 kV respectively, at the Center for Electron Optics at MSU. Samples were mounted on the sample stab with a conductive tape. Electron diffraction was done on a transmission electron microscope (TEM) JEOL-100CX at 120 kV, at the Center for Electron Optics. Samples were ground and suspended in water or acetone before 121 deposition on copper grids. Gold film doposited on a copper grid was used as standard in the calibration of camera length. b. Composition of nanocomposites The amount of polymer in the nanocomposites was estimated by TGA in oxygen flow. For comparison, the weight loss of Li0,2TaS2 in oxygen flow was checked by TGA up to 800 OC and was found to be about 8.2% [16]. The amount of polymer and water in the nanocomposites was calculated assuming that the Lio,2TaS2 in the nanocomposites lost the same amount of weight. The final product from the TGA in an oxygen flow was mainly Ta2Os, which was confirmed by XRD and IR spectra, however, it probably also contains some Li2SO4 [17]. In TGA measurements under either oxygen or nitrogen flow, the nanocomposites barely lost weight at a temperature lower than 190 0C (< 0.5% for batch-I nanocomposites and < 2.5% for batch-II nanocomposites), suggesting they contain almost no water. The major weight loss step for all the nanocomposites occurred between 200 0C to 350 OC and corresponds to polymer decomposition. c. One-dimensional electron density measurements One-dimensional electron density (l-D ED) maps were calculated according to XRD data collected from highly oriented film samples in a stepwise scanning mode. These film samples were made by casting aqueous nanocomposite solutions under a nitrogen flow, so that the basal planes of the TaS2 layers restacked parallel to the substrate. Several layers of a film were loaded in the sample plate so as to obtain maximum diffraction intensity. The XRD experiments were carried out under a nitrogen 122 atmosphere. In the measurements, slits for different beam width, as well as different data collection times, were used for different 20 ranges, in order to achieve a compromise between peak broadening, peak intensity and experiment time. Specifically, 0.5° slits and a data collection time of 3 s per step were chosen for measurements in the range from 2° to about 40°; 10° slits and 3 8 per step between 26° and 70°; 20° slits and 60 s per step between 59° and 120°; 4.0° slits and 60 s per step from 101° to 135°. The step width was kept the same (01°) in the entire 20 range. A full range XRD pattern (2° S 20 S 135°) was obtained by merging the data from the different 20 ranges and normalizing (scaling) them based on the overlapped regions. The overlapped data regions had at least two peaks in common. The full data set was put into an XRD analysis program, PEAKOC [131, to calculate the integrated peak area of each 001 reflection. In pattern analysis, 001 peaks were fit with the split-pseudo-Voigt function with a linear background subtraction for each peak. In the case that an 001 peak overlapped a little with another peak, or was cut off at the end of the pattern, it was fit with the pseudo-Voigt function (which is symmetric). The integrated peak area of the 001 peaks was used as the intensity of the peaks, and put in a FORTRAN program to calculate the 1-D ED map. The absolute value of the structure factor F of the 001 reflections was calculated from the intensities and the Lorentz-polarization factor (LP) according to (1). |F(l)| = (1/L,,)“2 (1) The Lorentz-polarization factor (LP) is defined as: 123 = (1 + cosz20)/(sin20cos(9) (2) The signs (phases) of the structure factors were assigned based on the signs of the corresponding structure factors of the TaS2 framework alone. This reasonably assumes that the scattering contribution from the intercalated polymer is relatively small compared to that of TaS2 component. The 1-D ED map was calculated from the structure factors of the 00l reflections according to (3). p(z) = (l/L)[ F0 + 22 F(|)cos(27c|z) ] (3) I In (3), z is the fractional coordinate on the c axis, L is the basal spacing of the layered structure, and F0 is the zeroth-order structure factor. To obtain their signs, the structure factors of the TaS2 framework were calculated from: N Flcal =212£11coslj2ttlz ) (4) The scattering factor fj(l) for atom j was calculated according to (5) with ai, bi and c for the element obtained from literature [19], and 0 corresponding to the reflection position. 2 ' 2 bléa-epr b—i—xineh elem-35%) (51 After the model for the structure was established, the sign of the F(l) was checked by recalculating F0310) from the scattering of the atoms including those of the polymer. 124 Results and Discussion 1. Exfoliation Properties of LixTaS2 N axTaS2, as reported previously, did not form ideal suspensions in water, so formamide (FM), N-methylformamide (NMF) or NMF/H2O 1:1 mixture were used to produce concentrated suspensions [9. 10b]. FM and NMF are not preferred for the intercalation of polymers, because not only they are expensive and toxic, they also compete with the polymers and co- intercalate in the host. In order to be able to use TaS2 as a polymer host we needed a material which exfoliates well in water. Therefore, we explored the swelling and exfoliation properties of LixTaS2 as a function of x. LixTaS2 with different degrees of lithiation was prepared under different conditions, through (a) reaction with butyllithium in hexane, (b) reaction with sodium dithionate in water followed by ion exchange with Li+ ions, and (0) reaction with LiBH4 in ether in various stoichiometric ratios and conditions. The LiBH4 method [20] was most successful. We found that the best LixTaS2 suspension was obtained with x ~ 0.2. Namely, Lio2TaS2 formed concentrated and stable colloidal single layer solutions in water, superior to those formed with N axTaS2 in NMF/H2O solution. In all cases, there was always a fraction of LixTaS2 which did not exfoliate, regardless of how dilute the solution was. Usually, the more readily the LixTaS2 exfoliates, the more concentrated suspension it forms and the less residue it leaves behind. To demonstrate how the amount of the LiBH4, in the lithiation reaction, affected the exfoliation of LixTaS2, a semi-quantitative experiment was performed with 0.050 g LixTaS2 with n = 0.1, 0.2, 0.3 and 0.4 [14]. After LixTaS2 was sonicated in 200 ml degassed water for 30 min, the precipitates were collected, dried and weighed. Based 125 on the LixTaS2 used, the unexfoliated fraction was 53%, 2.6%, 21% and 21% respectively for LixTaS2 with n = 0.1, 0.2, 0.3 and 0.4. These results showed that 0.1 equivalent of LiBH4 was not enough, while 0.3 or 0.4 equivalent of LiBH4 were over the optimum ratio. It needs to be pointed out that the charge density of Li02TaS2, ~ 48 AZ/e' [21] , is close to the up limit of the 40 - 120 AZ/e' range [22], in which stable colloidal dispersions are most likely to form. LixTaS2 used by Oriakhi et al [5b] to prepare the PEI/TaS2 nanocomposite was lithiated by reacting with LiOH [23]. Our investigation into that lithiation reaction reveals that control of the amount of LiOH is equally important and the optimum stoichiometric ratio is approximately the same. 2. Polymer Encapsulation The success of polymer encapsulation depends on the nature of both LixTaS2 and polymer. Table 3.2 shows which attempts succeeded or failed in the preparation of nanocomposites. PEO was intercalated only in “Li0,2TaS2”. PVP was intercalated in LixTaS2 prepared with 0.2 < x < 0.4. PEI has the most affinity for LixTaS2 and intercalated in all LixTaS2 samples [24]. This must due to the fact that PEI is a strong organic base which could be protonated in solution. The protonated PEI has positive charges which facilitate (through electrostatic attraction) the interaction with the [TaS2]"' layers. In this sense, the insertion of PEI in [TaS2]"' is an ion-exchange process where Li+ ions are replaced with positively charged PEI molecules. Other polymers such as polyacrylamide (PAM) and methyl cellulose (MCel) were also tried, with partial intercalation achieved with PAM and no intercalation with MCel. 126 Table 3.2. Polymer Intercalation in LixTaS2 intercalation of polymers PEI PVP PEO n=0.1 Yes No n=0.2 Yes Yes Yes n=0.3 Yes Yes n=0.4 Yes Yes No n=0.5 Yes No n=1.0 Yes No * "Yes" indicates that the intercalation was successful, while "No" shows that the intercalation failed. 3. Characterization of LixTaS2/Polymer N anocomposites The formation of lamellar LixTaS2/polymer nanocomposites was evident in the XRD patterns of the products. The presence and position of 00l reflections indicated the separation of the TaS2 slabs, which is evidence for the insertion of polymer chains. The pristine TaS2 and the hydrated Lix(H2O)yTaS2 show interlayer spacings of 6.0 A and 7.4 A respectively, which can be readily distinguished from a polymer intercalated phase. The d-spacings of batch-I and batch-II for both PEO and PVP intercalated compounds were comparable within :1 A. The XRD patterns of the two types of products were similar, except that in the former type (batch-I) we could observe some weak diffraction peaks from residual un—intercalated solid. In the case of PEI, the d-spacing of batch-I product was about 10.0 A, while that of batch-II product was about 37 A indicating the encapsulation of multiple layers of polymer. Some typical XRD patterns of 127 (a) E? .5 L 5 100A :3 ' (C) .fiL_—4— - _SL_ 36.8A (d) 1L L1 IlIIIIIIIIIIIIIIIIIIIIIIIIIfi 0 10 20 30 40 50 60 20(deg) Figure 3.1. XRD patterns of LixTaS2/polymer nanocomposites. (a) Li0,2(PEO)yTaS2 (batch-II), (b) Li0,2(PVP)yTaS2 (batch-II), (c) Lix(PEI)yTaS2 (batch-I) and (d) Lix(PEI)yTaS2 (batch-II). 128 Table 3.3. Properties of Li02TaS2 and polymer nanocomposites room sample d- coherence composition temperature spacing length conductivity (A) (A) (SIcm) LiO,2TaS2 6.0 >103 Hydrated Lio,2Ta82 7 .4 Lio_2(PEO)yTaS2 15 .6 93 Li0.2(PE0)1.36(H20)0.00Ta32 19 (100K) (batch I) Li0.2(PE0)yTa32 15.2 83 Li0.2(PEO)1.51(H20)0.23Ta32 2.2 (100K) (batch H) Li0,2(PEO)yTaS2 15.1 106 Li0.2(PEO)0.75(H20)0.06TaS2 17 (5M) (batch I) Li0.2(PEO)yTaS2 15 .5 83 Li0.2(PEO)l.68(H2O)0.00T352 3 .0 (5M) (batch H) Li0.2(PVP)yTa82 3 1 . 1 256 Li0.2(PVP)0.95(H20)0.08T332 3 1 (10K) (batch I) Lio,2(PVP)yTaS2 30.7 186 Li0.2(PVP)1.28(H20)0.35T332 0.5 (10K) (batch II) Lix(PEI)yTaS2 10.0 93 Lix(PEI)0.82(H20)o.07Ta82 125 (25K) (batch I) Lix(PEI)yTaS2 36.8 - Lix(PEI)3.2(H2O)0.56TaSZ 0.18 (25K) (batch H) * The X-ray scattering coherence length is determined from the half-width of the peak using Scherrer formula Lhk, = K A/Bcose, see Reference 18 of Chapter 2. 129 the nanocomposites are shown in Figure 3.1. More information about d- spacings and coherence lengths in the direction of layer stacking is listed in Table 3.3. From these, one can estimate the ordered domains in particles of the nanocomposites to be between 5 and 10 TaS2 layers thick. As the number of layers in the nanocomposite particles is considered, Lix(PEI)y(H2O)zTaS2 (batch I) has the largest value (9.3), Li0.2(PVP)y(H2O)zTaS2 has medium values (6. 1-8 .2), and Lio,2(PEO)y(H2O)zTaS2 has the smallest values (5.4-7.0). Table 3.3 also lists the compositions of the nanocomposites. It is obvious that a batch-H nanocomposite contains a little more polymer than its batch-I analog. This is understandable because batch-II product was well dissolved in water and reacted more completely with the dissolved polymer. DSC measurements on a Li0,2(PEO)1,o4TaS2 showed no phase transition between room temperature and 300 0C, see Figure 3.2. 3 I 330 °c " .9 6 ‘ E Z <1) A . 5 _ O E 4 a: if I 1st cycle 2nd cycle 0 .9 8 2 _ 308 c g a: b / 2 ‘6' 0 - 8 _ (D _2 .- I I J I I I I I I I I I I I I I I I L I I I 0 100 200 300 400 Temperature (°C) Figure 3.2. DSC of a Li0,2(PEO)yTaS2 nanocomposite. 130 Figure 3.3. SEM photograph of a Lio_2(PEO)yTaS2 film (batch-II; M.W. = SM). 131 The melting peak of PEO at 66 0C is not observed in the nanocomposite. A small exothermic peak at 308 0C (2.21 J/g) and a large one at 330 0C (46.18 J/g) correspond to the decomposition of the PEO in the nanocomposite and are associated with massive weight loss, detected by TGA under nitrogen flow. Pure PEO shows an exothermic decomposition peak around 370 0C. Mass spectra of the Lio,2(PEO)1,04TaS2 in a temperature increase ramp showed that the major decomposition fragments around 350 0C had molecular masses of 45, 58, 73, 87, 88, 89, 103, 120, 133 and so on, which were all PEO related fragments. There was no obvious evidence that the decomposition fragments combined with sulfur from the TaS2 layers [25]. The LixTaS2/polymer nanocomposites described here can be cast into free-standing films readily. The films of batch-II nanocomposites are stronger than that of batch-I products and can be folded without breaking. This is attributed to the higher polymer content of the batch-II materials. Figure 3.3 shows a photo of a folded Li0.2(PEO)1,6gTaS2 film ( batch-II, Mw = 5,000,000) taken on SEM. The film was folded before it was pressed onto the conductive tape. The folding edge did not crack under pressure. 4. Structural Studies: The Conformation of PEO in Lix(PEO)yTaS2 There are many known PEO chain conformations. The most common is the helical conformation [26] which exists in PEO spherulites [271. Planar zigzag conformation was obtained in stretched PEO samples [281. Two kinds of conformations were found in PEO-HgCl2 complexes [29’ 301. More PEO conformations have been found in other PEO complexes in literature and they will be described in Chapter 5. Models for the most important PEO conformations will be shown in Figure 5.11. 132 The conformation and orientation of polymer chains in the interlayer galleries of the layered nanocomposites has always been an important issue. We calculated 1-D ED maps for (PEO)XV2Os.nH2O [31], and determined that the helical conformation of PEO in that nanocomposite was not possible. A bilayer planar zigzag structure was proposed. In the present case, fihns of batch-II Li0,2(PEO)yTaS2 nanocomposite had well defined sharp XRD patterns with seventeen 001 reflections corresponding to a resolution of 0.85 A, see Figure 3.4. This intense XRD pattern offered a good starting point for 1-D ED calculation for Lio,2(PEO)yTaS2 nanocomposite, which provides information for the internal structure of the intercalated species projected on the c axis. As mentioned in the experimental section, the phases for calculating this l-D ED map were obtained from the positions of the Ta and S atoms which are taken to be known. A 1-D ED map is shown in Figure 3.5d. Intensity 0012 0013 0015 0017 0010 J\ 0014 0016 , 0011 IIIIIITIIIIII 0 20 40 60 80 100 120 140 20 (deg) Figure 3.4. Powder XRD patterns of a folded Li0,2(PEO)yTaS2 film (batch- II; M.W. = 100K). 133 ——:—-—' A E'- $ Electron density 5 I I l I I I I L I J I I I 0 0.2 0.4 0.6 0.8 l c-axis Figure 3.5. One-dimensional electron density maps for the Lix(PEO)yTaS2 nanocomposite. (a) Model with two layers of planar zigzag PEO chains while the zigzag planes are perpendicular to the TaS2 layers, (b) Model with two layers of planar zigzag PEO chains while the zigzag planes are parallel to the TaS2 layers, (c) Model with two layers of PEO in the type II PEO-HgCl2 complex conformation, and (d) Computed from experimental data. 134 888%¢8888&EW.A 8888888é%84 ' ~3V%®%?Wwv:mfi fikai-fo-Jkfi sh}- ‘91-‘84?» «— S atoms o——Ta atoms Figure 3.6. Structural model for the Li0,2(PEO)yTaS2 nanocomposite. The oxygen atom region in the middle of the gallery accommodates the Li+ ions. 135 The profile of 1-D ED map of Li0,2(PEO)yTaS2 shows clearly the presence of substantial electron density between the TaS2 peaks. This density is due to the organic polymer and it is distributed away from the center of the gallery, peaking in two separate locations symmetrically above and below the gallery central plane. The peak shape in each location is asymmetric. This immediately excludes the helical conformation of PEO, which must have one symmetric envelop of electron density in the central region of the gallery. The observed profile in the Li0,2(PEO)yTaS2 is also different from that of the (PEO)XV2Os.nH2O system which displays two symmetric bumps with two maxima on each of them, so the planar zag-zig model does not fit here. Figures 3.5a and 3.5b present the profiles of the 1-D ED map calculated for two layers of planar zigzag PEO chains arranged parallel and perpendicular inside the gallery space of TaS2 [321. As expected, they have two symmetric bumps and their maxima are too far apart to match the profile calculated from XRD data of Li0,2(PEO)yTaS2. The structural model best matching the experimental data is set up with two layers of PEO with the conformation found in the type H PEO- HgCl2 complex. This conformation has oxygen atoms on one side and carbon and hydrogen atoms on the other side of the molecule. By putting two layers of PEO inside the gallery so that the oxygen atoms face each other in the middle of the gallery, as shown in the model in Figure 3.6, two asymmetric bumps appear in the electron density map and the positions of the atoms can be decided by fitting maxima of the bumps (see Figure 3.5c). In this model, the atoms of PEO occupy reasonable positions in the gallery. In addition, the orientation of the PEO chains is chemically plausible because the hydrophobic part of the PEO (—CH2-CH2- groups) forms van 136 der Waals contacts with the sulfur atoms in the TaS2 layers and the oxygen atoms form a more polar, hydrophilic environment in which, presumably, the small Li+ cations reside. This model not only matches the experimental data very well, it also makes good chemical sense. The type II PEO-HgCl2 complex conformation proposed here is probably brought about by coordinating interactions of PEO with the Li+ ions, just as for the HgCl2 complex. That the PEO conformation in Lix(PEO)yTaS2 is different from that of (PEO)XV2Os-nH2O [29] is attributed to the lack of coordinating ions (e.g. Li+) in the latter. 5. Superconductive State The magnetic properties of LixTaS2 and nanocomposites were measured with a SQUID. As observed earlier with the NbSe2 system [3°], the lamellar nanocomposites reported here undergo superconducting transitions at temperatures higher than that of the pristine 2H-TaS2, 0.6 K (or 0.8 K according to different publications). It is known that the To of TaS2 is pushed down by the periodic lattice distortion-charge density wave (PLD-CDW) of TaS2. If the PLD-CDW is suppressed, the Tc could be raised to 4.1-4.5 K. In fact 0.08 equivalents of electrons are enough to suppress the PLD-CDW in TaS2 [33, 341. Therefore in the LixTaS2 we expect a higher To. The Cooper pairs in the superconductive state have a coherence length in the magnitude of a micron. This spatial correlation length is much longer than the gallery space in the LixTaS2/polymer nanocomposites which is occupied by the insulating polymer chains. The Cooper pairs can penetrate the barrier of the polymer layers and move around in the nanocomposites, so the intercalation of polymers (at least up to a certain expansion of the TaS2 layers) does not destroy the 137 Table 3.4. Properties of the superconductive state for TaS2 intercalates sample xmolar percentage of sample State To (K) (2K,5G) superififliuaa LixTaS2 powder 3.7, 4.3 -2.64 82 (n=0.2) LixTaS2 powder 4.1, 4.6 -0.128 4.0 . (n=0.4) LixTaS2 powder 2.7, 4.3 _1.1*10-3 3.41.10:2 (n=0.5) Lix(PEI)yTaS2 powder 2.9 -0.60 14 (batch I) Li0,2(PVP)yTaS2 powder 2.5 -0.82 5.8 (batch I) Lio,2(PEO)yTaS2 film 2.7 -0. 17 (batch 1) parallel film 2.6 -2.7 42 perpendicular Li0,2(PEO)yTaS2 film 2.9 -5 .6* 10'3 (batch 11) parallel film 2.9 .53410-2 0.9 perpendicular * Densities of the materials are needed in the calculation. The density of 2H-TaS2, 6.075 g/cm3 is used in the calculation of LixTaS2 (n=0.2, 0.4 and 0.5). The densities deduced from the compositions and structural parameters of the nanocomposites are used in the calculation of the percentage of superconductive state in the nanocomposites. 138 superconductive state. Measurements of Tc and Meissner effect are summarized in Table 3.4. A pure superconductor has the magnetic susceptibility of x=-1/471:. The percentage of the superconductive state in the samples is calculated by comparing the susceptibility of the samples at 2 K with -1/41t. The high percentages found in many of the nanocomposites indicate that we are observing bulk superconductivity. The TC values were determined by the point of intersection of the extrapolations from the linear magnetization of the superconductive state and the normal-state magnetization. The LixTaS2 (n=0.2, 0.4 and 0.5) samples exhibit two Tc’s, so they are mixed phase superconductors. This can be compared to the multiphase samples prepared from reaction of 2H-TaS2 with less than 0.5 equivalent of NaOH reported by Biberacher et a1 [23]. After Li02TaS2 is intercalated with polymers, the second phase disappears or is not detectable because Tc moves to lower temperature. From Table 3.4, it can be seen that the more the TaS2 is reduced, the lower the percentage of superconductive state in the sample. From LixTaS2 (n=0.2) to LixTaS2 (n=0.5), the value decreases three orders of magnitude. This shows the enormous effect of reduction to the electronic structure of TaS2. It explains why the control of the amount of LiBH4 in the lithiation reaction is so important to the exfoliation and intercalation properties of the LixTaS2. The intercalation of polymer causes the percentage of superconductive state to drop, but not by orders of magnitude. This is shown by the powder samples of Lix(PEI)yTaS2 and Li0,2(PVP)yTaS2, and the film sample of batch-I Li0,2(PEO)yTaS2, see Table 3.4. The low percentage values of batch-II Lix(PEO)yTaS2 films could result from the smaller dimensions of TaS2 slabs in the nanocomposite. 139 (a) 0.5[IIIIIIIIIIIIIIIIIIIIIIIIIIlfir 1 Zero field cooling IIIIII_IIIIIIIIIIIIIIIIIIIIIIII 2.5 3.0 3.5 4.0 4.5 5.0 Temperature (K) -0.5 -1.0 X molar -1.5 -2.5 JJIIIIJIIIIIJIIIJIJIIILIIIIII III -3.0 N llllllIITIIIIIIIIIIIIIIIIIIIIIIIII 'o (b) 0.5-rIIIIlllllIIIIIIIIIIIIIIIFIIr] O 3 Field cooling x molar I p—I LII IIIIIIIIIIIIIIIIIIIIIIIIIIIIII Zero field cooling IIIIIIIIIIIIIIIIIIIIIIIIIIIIII 2.0 2.5 3.0 3.5 4.0 4.5 mIIIIIIIIIIIIIIIIIIIIIIIIIIIIIJ_LIJ Temperature (K) Figure 3.7. Variable temperature magnetic susceptibility (at 5 Gauss) for (a) Li02TaS2 (powder), and (b) Li0,2(PEO)yTaS2 (batch-I; M.W. = 100K; films perpendicular to the applied magnetic field). 140 (a) (b) X molar X molar 0.5 -0.5 -1.0 -1.5 p—I 1.5 IIII lllI IIII I'll IIII IIIT I I I I I Field cooling Zero ield coolin IIIIIIIII IIII IIIIII IIIJIIII in IIIIIIIITII1IIIIIIIIIIIIIIIIIIIIII P IIIIIIIIIIIIIIIIIIIIIIIIIIlII III 2.0 2.5 3.0 3.5 Temperature (K) 4.0 IIIIIITIIIIIIIIIIIIIIIIIIIIII1IIII IIITIIIIIIIIIIIIIIIIIIIIIIITI IILII ILIILIIJJILLI III1 JIIIIJIIIIIIIIIIIIIIIIIIIJIIL P IIII L1] 2.0 2.5 3.0 3.5 4.0 Temperature (K) Figure 3.8. Variable temperature magnetic susceptibility (at 5 Gauss) for Li0,2(PEO)yTaS2 films (batch-I; M.W. = 100K). (a) films perpendicular to the applied magnetic field, (b) films parallel to the applied magnetic field. 141 Field dependent magnetic measurements on LixTaS2 and the nanocomposites show that the Meissner effect decreases with increasing magnetic field. Variable temperature magnetic susceptibility measurements show that the Meissner effect decreases gradually with increasing temperature. Furthermore, the field cooling curve diverges from the zero- field cooling curve in these materials, which is consistent with a type—II superconductor [35]. In the variable temperature magnetic susceptibility measurements, displayed in Figure 3.7, the field cooling curve was very different from the zero-field cooling curve in Li0_2TaS2 [Figure 3.7(a)], while the two curves were much closer together in batch-I Li0_2(PEO)yTaS2 [Figure 3.7(b)] . The two curves of batch-H Li0,2(PEO)yTaS2 were also close together. The drift of the field cooling curve toward the zero-field cooling curve, which was also found in the PEI and PVP intercalated nanocomposites, indicates that these materials are less able to pin the electromagnetic vortices present in the material. This is typical of granular superconductivity which appears in materials composed of tiny grains of superconductor particles surrounded by insulating layers [35]. In contrast, NbSe2 nanocomposites do not show this effect [30]. A dramatic difference is observed with sample orientation in the magnetization curves below To, see Figure 3.8. The Meissner effect is much stronger when the film is placed perpendicular to the magnetic field. In both batch-I and batch-II LixTaS2/PEO samples, the xmolar was > 10 times larger when the film was perpendicular to the field (i. e. TaS2 layers perpendicular to field) than when it was parallel. This observation suggests that the Cooper pairs are moving predominantly in a circular way within the TaS2 layers without passing through the polymer layers. This is a 142 beautiful demonstration of the two-dimensional nature of these nanocomposites. The magnetic properties of LixTaS2 and LixTaS2/polymer nanocomposites were also measured at temperatures above TC. In this temperature range the magnetic susceptibility of LixTaS2 (n=0.2 and 0.4) was Pauli-like and temperature independent with molar XTIP around 1.0*10'4 cm3/mol. 6. Electrical Transport Properties Room temperature electrical conductivity measurements show that Lio_2TaS2 and its nanocomposites are good conductors, see Table 3.3. Li0_2TaS2 itself is highly conductive, with a conductivity higher than 1000 S/cm. The intercalation of polymers brings the conductivity down by several orders of magnitude. The batch-I products of nanocomposites are l to 2 orders of magnitude less conductive, while the batch-II products are 3 to 4 orders of magnitude less conductive. This is probably because the batch-II products contain more polymer, and have TaS2 slabs of smaller dimensions. Variable temperature measurements for batch-I and batch-II of Lio,2(PEO)yTaS2 showed that the conductivity of both increases with falling temperature, consistent with metallic charge transport in these materials, see Figure 3.9. The conductivity reached a broad maximum at 75 K in batch-I Li0,2(PEO)yTaS2 and at about 130 K in batch-II Li0,2(PEO)yTaS2 before it started to drop. The superconductive behavior was not detected because our experimental set up could not reach the low temperature at which the metallic to superconductive transition occurs. The metallic character is further confirmed by thermopower data shown in Figure 3.10. The Seebeck coefficient is very small and negative suggesting n-type 143 _7 —' 7' _”_*%Fi T, 7—; 2O L - t - L - ”a“ 15 — 5 2 ' . re : 3 b u .. IE 10 T i g _ _ _d - . g ' 1 c.) 5 T 7. 0 - I I I I I I I I I I I I I I I I LL I I I l I I I I I I I - 0 50 100 150 200 250 300 Temperature (K) Figure 3.9. Variable temperature electrical conductivity measurements for pressed pellets of Li0,2(PEO)yTaS2 (batch-I product). 9 -: E. -1;- —; 8 ~25- 5 g 3 : 8‘ '35‘ r: E -4; _- J: : ' : I" _-_ (b I -5 r __ C _6bl I I I I I I I I I I I I I I I I I I I IJ I I I I I I I 0 50 100 150 200 250 300 Temperature (K) Figure 3.10. Variable temperature thermopower data of pressed pellets of Li0,2(PEO)yTaS2 (batch-I product). 144 (a) . Figure 3.11. Electron diffraction pattern and superstructure of L102TaS2. The sample was exfoliated in water before deposition on carbon-coated copper grids. The / 3 x,/ 3 superlattice reflections are shown with an arrow. 145 metallic behavior. The n-type transport is somewhat surprising since one might expect them to be hole conductors, given that 2H-TaS2 has a half- filled conduction band and injection of a small amount of electrons should result in a band which is more than half-filled. The failure of this prediction is due to the formation of a 3/5 x\/3 superlattice in the TaS2 layers, which was detected by electron diffraction, see Figure 3.11. This superlattice is believed to be due to Ta-Ta bonding interactions within the TaS2 layer and leads to a modification of the electronic band structure which changes the charge transport properties. There is no doubt from the above data that the charge transport within the layers is metallic. However, when the carriers go through the boundary areas and polymer layers, activation energy is needed. This contributes a thermally activated contribution to the conductivity, especially when the TaS2 slabs become more separated as the polymer content increases. 7. Solid State N MR Spectroscopy In order to probe the behavior of lithium ions in the two different gallery environments, (i. e. gallery with and without polymer) we measured the variable temperature static solid state 7Li NMR spectra for Li02TaS2 and Li0_2(PEO)yTaS2 (batch-II). Examination of the temperature dependence of the 7Li N MR line-width provides an independent perspective on the cation mobility in inhomogeneous solid, especially in (Li salt)/polymer complexes [36, 371 and Li/polymer/inorganic nanocomposites [381. Both Li02TaS2 and Lio,2(PEO)yTaS2 showed just one resonance peak due to a first order quadrupolar transition with no significant satellite peaks. 146 The spectra of Li02TaS2 at -80 0C and 100 0C are presented in Figures 3.12(a) and 3.12(b). At -80 0C, Li02TaS2 exhibits a broad peak with the maximum situated at a chemical shift of 5.5 ppm higher than the solid LiCl reference. The peak was somewhat asymmetric with a barely noticeable shoulder on the low field side. With rising temperature the position of the peak maximum did not change, but the shoulder protruded more and moved towards lower field. At 100 0C, the shoulder became a second peak at a chemical shift of 15 ppm higher than that of LiCl. The shape and position of the peak changed reversibly over many cooling and heating cycles. The change of the linewidth (the width at half-height) of the whole peak, versus temperature, is presented in Figure 3.13A(a). The broadening of the 7Li resonance peek is mostly due to many overlapping signals corresponding to slightly different lithium ion sites. The slow decrease of the linewidth with increasing temperature suggests that there is a large activation barrier for Li+ to hop from site to site. The presence of the shoulder in the spectrum indicates a different chemical environment for Li+ ions, which might be related to the multiphase observed in the superconductive state discussed above. The low temperature (-80 0C) static solid state 7Li NMR spectra of Li0,2(PEO)yTaS2 (batch-II) showed a broad symmetric resonance peak with a chemical shift almost identical to that of solid LiCl, see Figure 3.12(c). With increasing temperature, the position of the peak did not change, but its shape became asymmetric, see Figure 3.12(d). The peak base on the low field side extended a little farther than on the opposite side. The linewidth did not change appreciably in the temperature range from -80 0C to -40 0C but it narrowed dramatically from -40 0C to 60 0C and then continued to narrow albeit at a lower rate. At 100 0C, the linewidth of the peak was only 147 120 I. I I I I I I I I I I I I .- 100 '— fio (a) 4' I O I .- . O -1 1' . 00 '2 b t g 0 : .o-mq _— O. 0 1 11:): I 0 C22) I E r ’o ': H : .‘~ : (b) 1 _20 I I I I I L I I I I I I I‘ 60 20 0.0 -20 -40 Field (ppm) 120 '- I l .I l I I I I I I I I ‘- Z ’ I 100 — o _ : (d) : 80 3- ‘ 0 —‘ 3;. I o I a 60 :' 1 8 : 9 ’ : E 4": s . '5 0.0 -20 —40 Field (ppm) Figure 3.12. Static solid state 7Li NMR spectra of (a) Lio,2TaS2 at -80 0C, (b) at 100 0C, (c) Li0,2(PEO)yTaS2 (batch-H; M.W. = 100K) at -80 0C and (d) at 100 0C. 148 8000 . .44 T , . . . . , . . . . , . . . . , . 7000 (A) 6000 5000 4000 3000 Linewidth (Hz) 2000 1000 IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII O O O IlllIlllLIlILLIIIlIIIIlLLIIIIIIIILIIIII O IIIIIIIIJIIAIIIIIIIII -100 —50 O 50 100 Temperature (0C) 8000 7000 6. L110 CO CO CO 80.) o8 CO 1000 Linewidth (Hz) .h 8 O IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII I I I I I I I I I I I I I I I I I I I I I I I I I I I I I %L5 31) 215 41) 4L5 51) 1000/T (K'l) A 0' v 0 Sn IIJIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII Figure 3.13. Temperature dependence of the linewidth of 7Li NMR resonance peaks for (a) Lio,2TaS2 and (b) Li0,2(PEO)yTaS2 (batch-II; M.W. = 100K). 149 28% of that at -80 0C, see Figure 3.13A(b). This behavior is called dynamical motional narrowing and has been intensively studied in many solid polymer electrolytes [36. 37] and some PEO [or poly(ethylene glycol) (PEG)] nanocomposites [381. The change of line—width is caused by the averaging of magnetic couplings over the local magnetic field associated with other spins such as other Li nuclei, proton nuclei and unpaired electrons, due to dynamical motion. In most solid polymer electrolytes and PEO [or (PEG)] nanocomposites, dynamical motional narrowing is caused by the dissociation of a dipole-dipole coupling between pairs of nuclear spins, specifically 7Li and proton of the PEO [36,37, 38] and reflects the onset of the dynamical motion of either the polymer chains or the Li+ ions. In some materials with unpaired electrons, which produce much larger spin fields than nuclei, the line narrowing is caused by the averaging over the dipole coupling in the electron spin field [38]. Lio_2(PEO)yTaS2 is close to Li/PEO/fluorohectorite [38b], a PEO intercalated nanocomposite without paramagnetic centers in the inorganic layers. One might expect that the broadening of the 7Li resonance peak mainly results from the dipole-dipole coupling between 7Li and 1H of the PEO, as in the case of Li/PEO/fluorohectorite [38b]. This assignment would be supported by the similar linewidths before narrowing in these two nanocomposites, ~5 kHz, and the appearance of the plateau on the low- temperature as well as high-temperature side. If this is the case, the dynamical motional narrowing should correspond to the onset of the segmental motion of the polymer chains. However, the temperature range of the narrowing, -40 0C ~ 60°C, is more close to the onset temperature of the Li+ ion motion obtained in Li/PEO/montmorillonite, -20 0C ~ 400C [3321’ bl, rather than that of the polymer segmental motion observed in 150 28% of that at -80 0C, see Figure 3.13A(b). This behavior is called dynamical motional narrowing and has been intensively studied in many solid polymer electrolytes [36, 37] and some PEO [or poly(ethylene glycol) (PEG)] nanocomposites [381. The change of line-width is caused by the averaging of magnetic couplings over the local magnetic field associated with other spins such as other Li nuclei, proton nuclei and unpaired electrons, due to dynamical motion. In most solid polymer electrolytes and PEO [or (PEG)] nanocomposites, dynamical motional narrowing is caused by the dissociation of a dipole-dipole coupling between pairs of nuclear spins, specifically 7Li and proton of the PEO [36,3738] and reflects the onset of the dynamical motion of either the polymer chains or the Li+ ions. In some materials with unpaired electrons, which produce much larger spin fields than nuclei, the line narrowing is caused by the averaging over the dipole coupling in the electron spin field [33]. Lio,2(PEO)yTaS2 is close to Li/PEO/fluorohectorite [38b1, a PEO intercalated nanocomposite without paramagnetic centers in the inorganic layers. One might expect that the broadening of the 7Li resonance peak mainly results from the dipole-dipole coupling between 7Li and 1H of the PEO, as in the case of Li/PEO/fluorohectorite [381?]. This assignment would be supported by the similar linewidths before narrowing in these two nanocomposites, ~5 kHz, and the appearance of the plateau on the low- temperature as well as high-temperature side. If this is the case, the dynamical motional narrowing should correspond to the onset of the segmental motion of the polymer chains. However, the temperature range of the narrowing, -40 0C ~ 600C, is more close to the onset temperature of the Li+ ion motion obtained in Li/PEO/montmorillonite, -20 0C ~ 40°C [33% bl, rather than that of the polymer segmental motion observed in 150 Li/PEO/fluorohectorite, -100 0C ~ 40°C [38b]. In LiClO4(PEG)9, LiClO4(PEG)25 and LiBF4(PEG)9, where the motion of Li‘r ions is essentially governed by the segmental motion of the polymer chains so that the line narrowing corresponds to both the onset of the correlated motions of Li+ ions and polymer segments, the narrowing occurs in the range —50 OC ~ 500C [3631. Considering the line narrowing of Lio2TaS2, which happens in the same temperature range as that of Li0.2(PEO)yTaS2, we believe that this dynamical motional narrowing observed in Li0,2(PEO)yTaS2 is caused by the onset of the Li+ ion hopping in the gallery. This is because the [TaS2]"' anions are massive and only the motion of Li+ ions can cause the line narrowing in Li02TaS2. On the other hand, the linewidth after the narrowing is much wider in Li02(PEO)yTaS2 (~1500 Hz) than in LiClO4(PEG)9, LiClO4(PEG)25 and LiBF4(PEG)9 (60 - 90 Hz), which also suggests that the broadening of the 7Li resonance peak in Lio,2(PEO)yTaS2 is not caused only by the dipole-dipole coupling between 7Li and the polymer protons. For comparison with literature data, the 7Li NMR linewidths of Li02TaS2 and Li0.2(PEO)yTaS2 are ploted against 1/T in Figure 3.13B. The considerable narrowing of the 7Li resonance peak of Li0,2(PEO)yTaS2 at high temperatures suggests that the lithium ions begin to undergo facile site hopping. The mobility of lithium ions is affected by temperature more readily in the polymer intercalated galleries than in the un-intercalated galleries. This is attributed to the more disordered state of the Li lattice sites in the nanocomposites which lowers the activation barrier for hopping, relative to the more ordered, crystallographically well defined sites in Li02TaS2. The narrowing of linewidth in solid state 7Li 151 100 : I E (a) Parallel to field [I linewidth = 2893 Hz —)H - chemical shift = 3.14 ppm I I : (b) Perpendicular to field I - linewidth = 4618 Hz I 00 O O’\ O Intensity N A O O O I- _20IIJIIIIIIIIIIIIIIIIIJIIIIIIIIIIIIIILIII 200 150 100 50 0 -50 -100 -150 -200 Field (ppm) Figure 3.14. Room temperature 7Li NMR spectra for a Li0,2(PEO)yTaS2 film (batch-I; M.W. of PEO, 100K). (a) film parallel to the magnetic field and (b) film perpendicular to the field. NMR has also been observed in LixV2O5 [31] and LixMoOg [3b], as well as their PEO nanocomposites. A high degree of anisotropy was observed in the solid state 7Li NMR spectra of Li0,2(PEO)yTaS2 films (both batch-I and batch-II). As shown in Figure 3.14, both the resonance peak position and peak—width vary with the change in the orientation of the film of batch-I relative to the direction of applied magnetic field. When the film is perpendicular to the magnetic field, the peak width is broadened considerably more than when it is parallel. The mechanism of the change of the linewidth corresponding to the orientation is not yet clear. However, the high electron mobility in the two-dimensional TaS2 slab probably couples to the external applied magnetic field and causes local magnetic field distortions which affect the 152 Li nuclei and consequently the 7Li NMR signals. This orientation effect is less pronounced in films of batch-II Li0,2(PEO)yTaS2 probably due to the fact that the dimensions of TaS2 slabs are smaller and the slabs are less well stacked and more separated by PEO. The results of these experiments are summarized in Table 3.5. Table 3.5. Effect of film orientation on the 7Li-N MR spectrum of L10,2(PEO)yT382 L10.2(PEO)yT382 (batch-I) L10,2(PEO)yTaSZ (batch-II) parallel perpendicular parallel perpendicular chemicalshift of 3.14 12.57 10.99 9.42 peak p081t10n(ppm) lineWidth 2893 4618 2924 2902 (H2) * "parallel" and "perpendicular" refer to the film with respect to the instrument's applied magnetic field. Concluding Remarks The exfoliation properties of LixTaS2 were systematically explored and it was found that LixTaS2, prepared from controlled lithiation with 0.2 equivalent LiBH4, exfoliates well in water and has high affinity for various polymers. Lamellar nanocomposites of PEO, PEI and PVP were thus obtained through the encapsulative precipitation method. The nanocomposites dissolve in water and are easily cast into free-standing films. These plastic like films convert into superconductors at temperatures below their Tc’s, which raises the possibility of developing flexible superconductors with these or other materials in the future. Solid state 7Li 153 NlVIR measurements indicate that Lix(PEO)yTaS2 provides a more facile hopping environment for Li ions. According to the 1-D EM calculation for Lix(PEO)yTaS2, the nanocomposites are probably constructed with two sheets of PEO chains inserted in each gallery. The PEO chains adapt a conformation similar to that found in type II PEO-HgCl2 complex and are arranged with -CH2- groups facing the TaS2 layers and -O- atoms towards the center of the gallery where the Li+ ions seem to be located. 154 References 10 (a) E. P. Giannelis, Adv. Mater. 1996, 8, 29. (b) R. Krishnamoorti, R. A. Vaia. and E. P. Giannelis, Chem. Mater. 1996, 8, 1728. (a) E. Ruiz-Hitzky, Adv. Mater. 1993, 8, 334. (b) E. Ruiz-Hitzky, P. Aranda, B. Casal and J. C. Galvan, Adv. Mater. 1995, Z, 180. (a) R. Bissessur, M. G. Kanatzidis, J. L. Schindler and C. R. Kannewurf, J. Chem. Soc., Chem. Commun., 1993, 1582. (b) L. Wang, J. Schindler, C. R. Kannewurf and M. G. Kanatzidis, J. Mater. Chem. 1997, Z, 1277. (Also Chapter 4 of this dissertion.) (c) H.-L. Tsai, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Chem. Mater. 1997, 9, 875. L. F. Nazar, H. Wu and W. P. Power, J. Mater. Chem. 1995, 5, 1985. (a) J. P. Lemmon, J. Wu, C. Oriakhi and M. M. Lerner, Electrochim. Acta 1995, Q , 2245. (b) C. O. Oriakhi, R. L. Nafshun and M. M. Lerner, Mater. Res. Bull. 1996, 3_1, 1513. A TaS2/PEI nanocomposite by almost the same preparation approach has been reported by Oriakhi et a1 [5b]. TaS2/Polymer(oligomer) nanocomposites were also synthesized by other methods. (a) TaS2/Poly(4-vinylpyridine) nanocomposite was synthesized by in situ monomer intercalation and interlayer gallery polymerization: C.-H. Hsu, M. M. Labes, J. T. Breslin, D. J. Edmiston, J. J. Winter, H. A. Leupold and F. Rothwarf, Nature, Phys. Sci., 197 3, _2_4_6_(155), 122. (b) Polypeptides were intercalated into 2H-TaS2 by direct insertion: V. M. Chapela and G. S. Parry, Nature 1979, £1, 134. D. W. Murphy and G. W. Hull Jr., J. Chem. Phys. 1975, 6_2, 973. A. Lerf and R. Schdllhorn, Inorg. Chem. 1977, m, 2950. (a) A. J. Jacobson, Mater. Sci. Forum 1994, 152-153 (Soft Chemistry Routes to New Materials), 1. (b) L. F. Nazar and A. J. Jacobson, J. Mater. Chem. 1994, 4, 1419. (c) A. J. Jacobson, in G. 155 ll 12 13 14 15 l6 17 18 Alberti and T. Bein ed., Comprehensive Supramolecular Chemistry, Vol. 1, Elsevier Science Ltd., 1996, pp 315. A. Lerf, E. Lalik, W. Kolodziejski and J. Klinowski, J. Phys. Chem. 1992, %, 7389. (a) P. Joensen, R. F. Frindt and S. R. Morrison, Mater. Res. Bull. 1986, 2_1, 457. (b) M. A. Gee, R. F. Frindt, P. Joensen and S. R. Morrison, Mater. Res. Bull. 1986, Q, 543. (c) D. W. Murphy and G. W. Hull, Chem. Phys. 1975, Q, 973. J. F. Lomax, Intercalation Chemistry of Layered Transition Metal Dichalcogenides with Organic Bases, Ph. D. Dissertation, Department of Chemistry, Northwestern University, 1986. The equivalent of LiBH4 used in the lithiation reaction, 11, directly affects the amount of Li, x, in LixTaS2. Since no experiments were arranged to determine the x values for different forms of LixTaS2, n is used here and in the discussion. When n is small (< 0.5), x, the amount of Li in LixTaS2, is expected close to the value of 11. However, this expectation is not checked with experiments. Therefore, “Li02TaS2” is used to present the form of LixTaS2 prepared under the stoichiometry n=0.2 only for convenience. Under oxygen flow, Li02TaS2 gained weight at first and then lost weight at about 600 0C. The net loss up to 750 0C was 9.18%, 8.55%, 7.63% and 7.37% respectively in 4 measurements due to the transformation to Ta205 and Li2SO4. For Li02TaS2 to become Li2SO4 and Ta2Os, the weight loss should be 5.89%. Because Li compounds can be volatile at high temperatures, the weight loss of the sample in an oxygen flow could be somewhat higher than the theoretical value. In addition, a trace of small molecules such as water and diethyl ether in the Li0,2TaS2 can also cause a slightly higher weight loss than the theoretical value. PEAKOC is an XRD powder pattern analysis program provided by Inel Inc. (Mail Adress in USA: P. O. Box 147, Stratham, NH 03885.) 156 19 20 21 22 23 24 25 26 27 28 International Tables for X-ray Crystallography , Kynoch Press, 1974. M. G. Kanatzidis and T. J. Marks, Inorg. Chem. 1987, 26, 783. While we expect the n values to be close to the x values in LixTaS2 when n is < 0.5, no attempt was made to determine the exact x in the samples. A. Jacobson, Mater. Sci. Forum 1994, I, 152. W. Biberacher, A. Lerf, F. Buheitel, T. Butz and A. Hfibler, Mat. Res. Bull. 1982, 1_7_, 633. PEI was also intercalated in the LixTaS2 prepared from either 2H- TaS2 or 1T-TaS2 through a solid state reaction with 3 equivalent of LiBH4 in the temperature range from 300 0C to 525 OC. The Lio2TaS2 and Li0,2(PEO)1_04TaS2 were heated to 350 0C under nitrogen atmosphere and kept for 10 minutes at this temperature to check if there was any loss of sulfur in the TaS2 layers. The sulfur content in these two samples as well as the untreated samples was checked with EDS, using synthesized 2H-TaS2 as a standard. The stoichiometric number for sulfur, m (as TaSm), was 1.96, 2.01, 1.98 and 1.96 for Li02TaS2, L10,2(PEO)1.04T332, treated Lio2TaS2 and treated Lio,2(PEO)1,o4TaS2 respectively. This indicates that almost no sulfur is lost in the lithiation, intercalation and heating treatment of Li02TaS2 and Li0_2(PEO)1,04TaS2 up to 350 0C under nitrogen. This is consistent with TGA which showed that Li02TaS2 had no weight loss in nitrogen up to 530 0C. (a) Y. Takahashi and H. Tadokoro, Macromolecules 197 3, _6, 672. (b) J. Brandrup and E. H. Immergut, Polymer Handbook, 3rd Ed., John Wiley & Sons, New York, 1989, pp VI/72. F. P. Price and R. W. Kilb, J. Polym. Sci. 1962, 5_7, 395. Y. Takahashi, I. Sumita and H. Tadokoro, J. Polym. Sci., Polym. Phys. Ed., 1973, fl, 2113. 157 29 30 31 32 33 34 35 36 37 38 R. Iwamoto, Y. Saito, H. Ishihara and H. Tadokoro, J. Polym. Sci., A-Z, 1968, _6, 1509. M. Yokoyama, H. Ishihara, R. Iwamoto and H. Tadokoro, Macromolecules 1969, 2, 184. Y.-J. Liu, J. L. Schindler, D. C. DeGroot, C. R. Kannewurf, W. Hirpo and M. G. Kanatzidis, Chem. Mater. 1996, 8, 525. The positions of the atoms of PEO were decided according to the structural data available from the PEO crystal structure with planar zigzag conformation [28]. When the ac plane of the unit cell is parallel to the layers, the PEO zigzag plane is almost parallel to the TaS2 layers; when the bc plane of the unit cell is parallel to the TaS2 layers, the PEO zigzag plane is almost perpendicular to the layers. P. Garoche, P. Manuel, J. J. Veyssie and P. J. Molinié, Low. Temp. Phys. 1978, E, 323. R. H. Friend and A. D. Yoffe, Adv. Phys. 1978, E, l. V. Z. Kresin and S. A. Wolf, Fundamentals of Superconductivity, Plenum Press, New York/London, 1990, pp 95. (a) S. Panero, B. Scrosati and S. G. Greenbaum, Electrochim. Acta 1992, 40, 1533. (b) W. Gorecki, E. Belorizky, C. Berthier, P. Donoso and M. Armand, Electrochim. Acta 1992, 3_7, 1685. (a) S. H. Chung, K. R. Jeffrey and J. R. Stevens, J. Chem. Phys. 1991, 94, 1803. (b) J. P. Donoso, T. J. Bonagamba, P. L. Frare, N. C. Mello, C. J. Magon, H. Panepucci, Electrochim. Acta 1995, J), 2361. (a) S. Wong, S. Vasudevan, R. A. Vaia, E. P. Giannelis and D. B. Zax, J. Am. Chem. Soc. 1995, 1_l7_, 7568. (b) S. Wong and D. B. Zax, Electrochim. Acta 1997, $2, 3513. (c) D.-K. Yang and D. B. Zax, J. Chem. Phys. 1999, m, 5325. 158 Chapter 4 LAMELLAR LixMoO3/POLYMER NAN OCOMPOSITES VIA ENCAPSULATIVE INSERTION Introduction The investigation of inorganic/polymer nanophase composites is motivated by many reasons, including the need for novel electronic anisotropic materials, better performing battery cathode materials, functionalized structural materials with superior mechanical properties, hierarchical materials, and systems in which to study polymer orientation, epi- and endo-taxy and polymer/inorganic surface interactions [“4]. Polymer-based nanocomposites have been reported with layered silicates (e.g. montmorillonite, hectorite, etc.) [11, FeOCl 12], V205 [3, 4], M003 [5. 6t 71, layer metal phosphates [3], MS2 (M=Mo [9’ 10, 11, 73], Ti [1131), NbSe2 [12], layered metal phosphorus chalcogenides (MPS3) [13, 7] and layered double hydroxides [14]. The most common methods of preparing inorganic/polymer nanocomposites are (a) by monomer intercalation followed by polymerization, (b) by in situ intercalative polymerization, (c) by direct insertion and (d) by encapsulative precipitation from solutions of exfoliated lamellar solids. The last two methods give inorganic/polymer nanocomposites in which the molecular masses and nature of the polymers can be decided before intercalation. Encapsulative precipitation has been applied with V205 133-Ct 4], MoO3 153-0’73], MoS2 [10131 C, 11, 7a], TiS2 [11a] and NbSe2 [12], in combination with various polymers. M003 is one of the layered metal oxides which shows reversible Li ion insertion properties which are relevant to rechargeable Li batteries [15]. 159 In intercalative electrodes of rechargeable batteries, ion conductivity is very important. Other solid-state ionic applications such as electrochromic devices also need good ion conductivity. Inorganic/Polymer nanocomposites should exhibit fast ion conduction 11%] and introduction of a polymer with affinity for Li ions between the sheets of MoO3 could improve its performance as an intercalative electrode. Polymer insertion into MoO3 has been reported previously, namely, with a polymeric ionomer [53. b], with PEO [56, 7a] and with polyaniline l5d~ 6’ 7b]. This research develops further the polymer intercalation chemistry associated with M003 and introduces a new family of LixMoOglpolymer nanocomposites. Nanocomposites with poly(ethyleneoxide) (PEO), poly(ethyleneglycol) (PEG), poly(propyleneglycol) (PPG), poly(vinylpyrrolidinone) (PVP), methylcelulose (MCel), polyacrylamide (PAM) and nylon-6 (PA-6) are reported here. These nanocomposites were characterized by thermal gravimetric analysis (TGA), differential scanning calorimetry (DSC), powder X-ray diffraction, FTIR and solid-state UV- VIS spectroscopy, variable-temperature solid-state 7Li and 13C NMR spectroscopy, magnetic susceptibility measurements and electrical conductivity measurements. Experimental Section 1. Reagents PEO (5,000,000), PEO (100,000), PEG (10,000), PEG (2,000), PEG (400), PPG (1,000), PVP (10,000), MCel (63,000), PAM (5,000,000), PA-6 (10,000) and LiBH4 (95%) were purchased from Aldrich Chemical Company, Inc. After the polymers were dissolved, the 160 polymer solutions were filtered to remove insoluble polymer residues. MoO3 (99.95%) was purchased from Johnson Matthey Catalog Company. Anhydrous diethyl ether (99.0%), 2,2,2-trifluoroethanol (99%), acetonitrile (99.5%), isopropynol (99.9%) and 200 proof ethanol were from Columbus Chemical industries Inc., Lancaster Synthesis Inc., EM Science Inc., Mallinckrodt Chemical Inc. and Quantuum Chemical Company respectively. No further purification was applied to the chemicals above. Water used in the reactions was distilled water provided by the Department of Chemistry, and was degassed by bubbling nitrogen for 30 min before use. 2. Synthesis of LixMoOg (0.30 < x < 0.40) Commercial MoO3, used as the starting material, was fired in an open quartz vial in the air at 600 0C for 36 h during which the crystal size of MoO3 remarkably increased. This M003 was used to react with LiBH4 to prepare the lithium molybdenum bronze [16]. In a typical reaction, 0.1 mol of MoO3 was reacted with 0.04 mol of LiBH4 in 80 ml diethyl ether for 24 h, under a nitrogen atmosphere. The product was collected by filtration, washed with ether and dried in vacuo. The yield was >98%. The lithium molybdenum bronze was thereafter stored in a nitrogen dry-box. The X-ray powder diffraction pattern can be indexed on the basis of an orthorhombic cell similar to that of MoO3 but expanded along the stacking (a-axis) direction, with a: 16.528 A, b=3.775 A and c=3.969 A. The dw- spacings (A) are: 8.26200 (vs), 4.132400 (s), 3.578201 (In), 3.434210 (m), 2.755600 (w), 2.453311 (m) and 2.372411 (m). The amount of lithium in the bronze was analyzed by TGA under oxygen flow, heating up to 650 0C, and by elemental analysis which was 161 accomplished by Oneida Research Services, Inc., Whitesboro, New York. Elemental analysis of Li in the molybdenum bronze was done by ion chromatography, while Mo was measured by X-ray fluorescence. 3. Preparation of LixMoOglPolymer Nanocomposites The LixMoOg was exfoliated in degassed water by 5 minutes of sonication, to form a suspension with a concentration of 5 g/L. This suspension was added dropwise into a stirred polymer solution of the same volume, which contained five times of excess polymer (per repeat unit) to M003 and the mixture was stirred for 2 days under a nitrogen atmosphere. The nanocomposites formed were isolated in different ways according to their behavior in solution. Nanocomposites of methylcelulose, polyacrylamide and nylon-6 precipitated and were collected by filtration. Those containing MCel and PAM were washed with water, while the nanocomposites of PA-6 were washed with trifluoroethanol. Nanocomposites with poly(ethylene oxide) (PEO), poly(ethylene glycol) (PEG), poly(propylene glycol) (PPG) and poly(vinylpyrrolidinone) (PVP) remained in colloidal form and were isolated by pumping off the water to dryness. The dried material was stirred in an appropriate solvent for several hours to dissolve the extraneous polymer. The solid product was filtered and was washed again with the solvent. The solvents used to process the products are list in Table 4.1. The products were pumped to dryness and stored in a nitrogen atmosphere. 4. Instrumentation The instrumentation in the measurements such as X-ray powder diffraction (XRD) patterns, thermal gravimetric analysis (TGA), 162 differential scanning calorimetry (DSC) and infrared (IR) spectrosc0py were the same as described in Chapter 1. Room temperature conductivity measurements, variable temperature direct-current electrical conductivity measurements and thermopower measurements, and magnetic susceptibility measurements were conducted as described in Chapter 2. Variable-temperature solid-state 7Li and 13C NMR measurements were taken on a 400 MHz Varian Nuclear Magnetic Resonance Instrument. Samples were loaded in a glove-box under a nitrogen atmosphere. Electronic transmission spectra were recorded with a Shimadzu UV- 3101PC UV-VIS-NIR scanning spectrophotometer. Samples were dried as thin films on quartz slides. Results and Discussion 1. Synthesis and Characterization of LixMoO3 The lithium molybdenum bronze exfoliates readily in water to form stable colloidal solutions, and this makes it an appealing candidate for polymer intercalation. The LiBH4 route to LixMoO3 is the best one to date in providing the material conveniently and in high yield. The previously reported method for Lix(H2O)yMoOg [17] involves one step to prepare N ax(H2O)yMoOg and two steps to accomplish ion-exchange and gives only 26% yield [18]. An additional advantage of the LiBH4 method is that it is conducted in diethyl ether and so provides the anhydrous form of the molybdenum bronze. The LixMoOg product prepared in this fashion, though still crystalline, shows broader diffraction maxima than the precursor M003. The Li insertion into M003 is topotactic as evidenced by 163 our ability to fully index the X-ray powder diffraction pattern of the product. The amount of Li in LixMoOg was determined both by TGA and by direct elemental analysis. When the material is heated in an oxygen atmosphere to 650 0C, it gains 1.71 mass%, owing to a change from LixMoOg to Li2O and MoO3. This gain of mass is reproducible and corresponds to an x value of 031(3). On the other hand, the elemental analysis showed that the molybdenum bronze consisted of 1.64% Li and 56.39% Mo. This gives a ratio of Li to M0 of 0.4:1 corresponding to the molar ratio of LiBH4 used in the lithiation reaction. It is of course possible that x varies slightly from sample to sample in the range 0.3 < x < 0.4. The lithium molybdenum bronze is metastable and undergoes an intense, irreversible and exothermic phase change at 356 0C, as detected by DSC (Figure 4.1) and X-ray powder diffraction. The product appears to be 5 >— : 356°C 4 :- .9. 3 ~ g A : o 3 2 :. >< E 2nd cycle 0 V E_lst cycle 8 1 .9. Q :/ g _4 0 Q g j e -1 _ ~ i £3 _ 0 I -2 L- 1 1 1 l l 1 1 1 1 l 1 1 1 1 l 1 m 1 l l l I 0 100 200 300 400 Temperature (0C) Figure 4.1. DSC diagram of LIXMOO3. 164 a mixture of at least two unknown phases. As this mixture is heated to higher temperatures it undergoes additional phase changes yielding other new phases. 2. Exfoliation and Polymer Encapsulation Chemistry The lithium molybdenum bronze described above can be readily exfoliated in water after several minutes of sonication. The exfoliated product forms supramolecular complexes with most water-soluble polymers. The complexes form solutions or precipitates in water, depending on the type and molecular mass of the polymers. We have encapsulated PEO, PEG, PPG, PAM, MCel and PVP inside the lithium molybdenum bronze to obtain lamellar nanocomposite materials. Poly(ethylenimine) (PEI) could not be successfully intercalated. Instead, the blue Lix(H2O)yMoO3 monolayer suspension decolorized and totally dissolved in the aqueous PEI solution. The same phenomenon occurred when ammonia was introduced in the Lix(H2O)yMoO3 suspension, suggesting that the PEI solution is too basic and attacks the M003 lattice. Water-insoluble polymers were also tried, however of these, only PA6 was successfully intercalated. Details of the reactions are given in Table 4.1. The existence of polymer chains between the layers of the lithium molybdenum bronze was verified by IR spectroscopy and X-ray powder diffraction. Figure 4.2 shows a comparison of the IR spectra of a nanocomposite Lix(H2O)y(PVP)zMoOg and its components; from this, it is obvious that the vibrational peaks of Lix(H2O)y(PVP)zMoO3 is a combination of the vibrational peaks of PVP and that of Lix(H2O)yMoO3. The positions of the vibrational peaks arising from the encapsulated PVP are close to those of pure PVP, while the positions of the peak due to the 165 Table 4.1. Chemical and structural characteristics of LiXMoO3/polymer nanocomposites nanocom- polymer solu- washing d- expansion coherence posite Mw bfljty a solvent spacing of gallery len (A) (A) < ) LixMoO3 5x 106 yes MeCN 1 6.6 9 .7 64 /PEO LixMoO3 100,000 yes MeCN l 6 .6 9 .7 121 /PEO LixMoO3 10,000 yes MeCN 1 6.8 9 .9 108 /PEO L LixMoO3 2,000 yes MeCN l 3 .8 6.9 72 /PEG 126-14.7 5.7-7.8 - b LixMoO3 400 yes MeCN 1 3 .5 6 . 6 1 1 1 /PEG 1 2.6 5 .7 97 LixMoO3 1,000 yes 17 .2 10.3 1 13 /PEG EtOH l 1 .8 4 9 108 18.0&1 1.5 - _ b LixMoO3 63,000 no H20 27 .6 20.7 92 lMCel LixMoO3 10,000 yes PriOH 3 8 .6 3 1 .7 154 /PVP LixMoO3 5x 106 no H20 3 8 .0 3 1 . 1 69 /PAM 33.8 26.9 80 4 1 .2 34. 3 - b LixMoO3 10,000 no CF3CH2OH 22.1 15 .2 39 /PA6 l 6. 8 9 .9 26 a. Nanocomposites dissolved in water, except for LixMoOglPA6, dissolved in 222-trifluoroethanol. b. Peaks too broad to obtain estimate. 166 % Transmittance (d) MO'O IIIIIIIIIIIIIIIIIII I II II I I I I I l I 2000 1800 1600 1400 1200 1000 800 600 400 Wavenumbers (cm'l) Figure 4.2. IR spectra of (a) poly(vinylpyrrolidinone), (b) Lix(H2O)y(PVP- 10,OOO)ZMOO3, (c) hydrated lithium molybdenum bronze, and (d) M003. 167 (b) Lix(I-IZO)yMoO3 Absorbance (c) Lix(HzO)y(PEO-100,000)zMoO3 286 nm ((1) LixG-IZO)y(PAM—5,000,000)zMoO3 200 700 1200 1700 2200 Wavelength (nm) Figure 4.3. Solid—state optical absorption spectra of the LixMoO3/polymer nanocomposites. The polymer and its molecular mass are indicated on each spectrum. 168 13.5 A PEG(400) ' "Turf ' ‘ “—-‘3‘*+M.... 13.8 A PEG(2,000) 54. 16.8 A 'FB :1 3 PEG(10,000) .5 16.6A PEO(100,000) 16.6A PEO(5,000,000) T I r I I I ' l I IfT II I I I I Ivl‘ri ffitIAIIifi-IJA-ggh 0 10 20 30 40 50 60 20 (deg) Figure 4.4. Typical XRD patterns of nanocomposites with poly(ethylene glycol) and poly(ethylene oxide) of different molecular mass. The polymer and its molecular mass are indicated on each spectrum. 169 17.2 A PPG(1,000) 27.6 A MCel(63,000) 38.6 A >~. H a C: 8 3 PVP(10,000) 38.0 A PAM(5,000,000) PA6( 10,000) I I f I T I I I I I I I T I r I I I I I I I O 10 20 3O 40 50 60 20 (deg) Figure 4.5. Typical XRD patterns of the various LixM003/polymer nanocomposites. 170 Mo=O stretch [for Lix(HgO)yMoO3] is shifted to higher wavenumbers, suggesting that the M003 layers in the nanocomposite are slightly more oxidized. The optical transmission absorption spectra of these macromolecular intercalates were examined. The dark-blue color of these systems arises from the intense intervalence transitions associated with the MOM/Mo6+ couple. These electronic transitions are broad and range from the IR region to the visible (Figure 4.3) and are responsible for the electrical conductivity of these materials. The absorption at 286 nm arises from excitations across the band-gap from the 02’ p-band to the Mo6+ d-band and is present in all compounds including pristine M003 and Lix(H20)yMoO3. This is consistent with the expectation that host metal oxide structure is not disturbed upon intercalation. The encapsulation of polymers inside the interlayer galleries of M003 is also demonstrated by X-ray powder diffraction, which shows an expansion of the gallery space. Figures 4.4 and 4.5 show typical XDR patterns of some of the nanocomposites. The sharp and intense (001) reflections indicate that the M003 layers are well stacked. X-Ray scattering coherence lengths, which are calculated from the Scherrer formula Lhkl = KNBcose (see Reference 18 of Chapter 2), and the gallery spacings are given in Table 4.1. The basal spacing of some nanocomposites depends on the preparation procedure. For example, Lix(H20)y(PVP)zMoO3, which is water soluble, showed a d-spacing of 59.0 A before washing with isopropyl alcohol and 38.3 A after washing. Washing these materials may not only remove extra-lamellar polymer, but could lead to polymer loss from the galleries changing their composition. This behavior makes it difficult to 171 determine at what polymer loading we begin to saturate the intralamellar space. Similar phenomena were described in polymer-V205 xerogel systems [33]. Nevertheless, the observed d-spacings were consistent to within :1 A, as long as the preparation procedures were not altered significantly from batch to batch. The d-spacings and the degrees of lamellar stacking of products which contained polymers at the very extremes of the molecular mass range (very high or very low) were hard to control, as for example in PEO(5,000,000), PEG(2,000) and PPG(1,000). The Lix(H20)y(PEO- 5,000,000)zMoO3 had a consistent d-spacing ca. 16.6 A, but the peaks were broad. Lix(H20)y(PEG-2,OOO)ZM003 showed broad peaks in the range 126-147 A. Lix(H20)y(PPG-l,OOO)zMoO3 sometimes showed a peak in the range 11.8 131-17.2 A, while other samples showed two peaks in this range suggesting a mixture of phases. Evidently, for low molecular masses the polymers are mobile enough in the galleries to form several different arrangements leading to multiple phases. The effect of polymer molecular mass on product formation was examined, particularly with PEO and PEG, and was found to be significant. The high molecular mass PEO(5,000,000) immediately formed a precipitate with lithium molybdenum bronze in water upon mixing while this phenomenon did not occur with PEO of lower molecular mass. The molecular mass also affects the structure of the nanocomposites. Table 4.1 shows that the Lix(H20)y(PEO-5,OOO,OOO)zMoO3 sample has a lower coherence scattering length than its lower molecular mass analogues, which is attributed to the fact that it is kinetically unfavorable for extremely long polymer chains to align in an ordered structure. When the molecular mass is extremely low, 1'. e. in oligomer range, the gallery expansion of the intercalate is lower, almost one half of that of the long polymers. For PEO 172 and PEG, a molecular mass of 2,000 is about the upper limit of this situation. The data listed in Table 4.1 show that the Lix(H20)y(PEG— 2,OOO)zMoO3 sample has a much shorter coherence length than its analogues with higher or lower molecular masses. An analogous Lix(H20)y(PEG-2,OOO)ZM003 sample, prepared under similar conditions, had a broad X-ray basal peak which corresponded to a d-spacing varying from 12.6 to 14.7 A, suggesting that the local conformation of the polymer is important. Annealing the Lix(H20)y(PEG)zMoO3 and the Lix(H20)y(PEO)zMoO3 samples at 150 OC and then gradually cooling them to room temperature tends to improve their lamellar order, especially when the starting coherence length is short. For example, after annealing, the Lix(H20)y(PEG-2,OOO)zMoO3 sample whose XRD pattern had a broad peak centered at 13.4 A gave a pattern with a sharp peak at 12.7 A, (Figure 4.6). The water—insoluble nanocomposite Lix(H20)y(PAM- 5,000,000)zMoO3 gave samples with d-spacings of 38.0, 33.8 and 41.2 131. The Lix(H20)y(PA6-10,000)zMoO3 showed d-spacings of 22.1 and 16.8 A. Occasionally, in the intercalation of PA-6 a mixed-phase material with basal spacings of 12.8 and 9.8 A was obtained. This shows the difficulty to control the reaction when quick precipitation is used to obtain a specific phase. To prepare nanocomposites of this type, a very dilute Lix(H20)yMoO3 suspension of < 0.5 mass% is recommended. The polymer composition of the nanocomposites were determined by TGA in an oxygen atmosphere and are listed in Table 4.2. The water in the nanocomposites was estimated by the mass loss step observed below 230 oC, and the amount of polymer was determined by the mass loss steps observed at higher temperatures. Despite drying under vacuum, the nanocomposites 173 13.4 A before annealing 12.7 A Intensity after annealing I I I I I I I I I I I r I I I T I T I 12 16 20 24 28 29 (deg) —1 -I>—l 00 Figure 4.6. XRD patterns of Lix(H20)y(PEO-2000)ZM003 showing the effect of annealing on the stacking regularity of the layered structure of a nanocomposite. contain some water in the galleries. The water-soluble nanocomposites, Lix(HzO)y(PEO)ZMoO3, Lix(H20)y(PEG)zMoO3, Lix(H20)y(PPG)zM003, and Lix(H20)y(PVP)zMoO3, usually contain 2-4 mass% water which is very difficult to remove. This water is thought to be coordinated to Li+ ions. The water-insoluble nanocomposites Lix(H20)y(MCel)zMoOg and Lix(H20)y(PAM)zMoO3, however, contain much less water. Compared to the corresponding M082/polymer intercalates [10b], most LixMoOg/polymer intercalates have much higher polymer contents and larger gallery spacings. As in the case of M082/polymer intercalates, PVP and MCel give the largest expansions. A marked difference is found in PAM which gives LixMoO3/PAM an expansion as large as 27-35 A, while M082/PAM has only an expansion of 9.4 A (see Chapter 1). This confirms 174 Table 4.2. Composition and physicochemical properties of the LixMoO3/polymer nanocomposites limit of limit of electronic nanocom- polymer d— composition thermal thermal conducti- posite Mw spacing (according to TGA) stability stability ‘ vity (A) in m /Scm‘1 N 2/ 0C (Dz/QC pure 8.27 1.3 10'2 LixMOO3 X pure 6.93 3.3 10'5 M003 X LixMOO3 5X106 16.6 LiX(H20)0.20PEOO.33MOO3 260 220 2.4X10-5 /PEO LixMoO3 100,000 16.6 Lix(H20)0_32PE01.04MoO3 260 220 2.9x 1 0-5 /PEO LixMoO3 10,000 16.8 Lix(H20)0,29PEG0,75M003 260 220 5.2x 1 0-5 /PEG LixMoOg 2,000 13 .8 Lix(H20)0,28PEG0.57MoO3 260 220 2.2x 10-4 /PEG LixMOO3 400 13 . 5 Lix(H20)0.38PEGO.33MOO3 260 220 3 . 1x10-4 /PEG LixMOO3 1,000 17 .2 Li X(H20)0. 18PPGQ99MOO3 220 200 - /PPG 1 1~8 Lix(H20)o.52PPGo.14M003 5.3x 10'4 LixMoO3 63 ,000 27 .6 LixMCCIOJOMOO3 180 170 2.0,(10-6 /MCel LixMoO3 10,000 3 8.6 Lix(H20)0,43PVP1.17MoO3 220 220 1.8x10'7 /PVP LixMoO3 5X106 3 8 .0 LixPAM32MOO3 150 150 6.3x 10'7 /PAM LixMoOg 10,000 22. 1 Lix(H20)0_44PA60.32MoO3 280 270 - /PA6 16.8 Lix(H20)0.48PA60.23M003 2.11110”4 175 that multiple layers of this polymer can enter the accessible space of LixMoO3. The water-soluble LixMoO3/polymer nanocomposites can be cast into films and other shapes, which may provide opportunities for applications. The nanocomposites with high molecular mass polymers are strong, though their mechanical properties depend on the polymer. For example, the nanocomposite of PEO(5,000,000) is tough, while that of PAM(5,000,000) is hard. Lix(H20)y(PEO-5,000,000)zMoO3 can be swollen by acetonitrile and becomes resilient and plastic with the consistency of unsulfurized rubber. When the Lix(H20)y(PAM—5,000,000)ZM003 is swollen by water, it is not as plastic, but is stronger and tougher. Precise mechanical measurements have not been taken. 3. Solid State NMR Spectroscopy In order to probe the effect of the polymer on the behavior of the lithium ions in the gallery, variable-temperature solid-state 7Li NMR static spectra were measured for LixM003 and Lix(H20)y(PEO—100,000)zMoO3. In both cases a broad peak was observed with a chemical shift very similar to that of the solid LiCl standard. At —80 0C the lineshape in the two spectra differs, with that of the Lix(H20)y(PEO—100,000)ZMoO3 spectrum being slightly more asymmetric, Figure 4.7. This suggests that the presence of PEO causes a distribution of Li ions over several, slightly different sites in the gallery. Some of the sites may involve coordination of water molecules while other site are associated with the ether—like oxygen atoms in PEO or even those in the M003 layers. The linewidth (width at half maximum) of the resonance peak is greater in the PEO intercalated sample than in the host LixMoOg material and this too is consistent with a distribution of the 176 80 Lix(H20)y(PEO-1OO,OOO)ZM003 60 - E? 40“— 273 i I: B - a 201— r—4 1- I .. IlllllllllLllllllLlllllllllllllllllllll 0 200 150 100 50 0 -50 -100 -150 -200 5(ppm) Figure 4.7. Static 7Li NMR spectra of LixMoO3 and Lix(H20)y(PEO- 100,000)ZM003 at -80 0C. The broader assymetric line in the spectrum of the nanocomposite is evident in the upfield region. 7000 DO 6000 8 9 O Lix(H20)y(PEO-100,000)ZM003 OD 5000 ' 4000 3000 Linewidth (Hz) 2000 1000 jI—IIIIIIIIIIIIIIIIIIIIIIIIIIIIIII O 0Llllllllllllllllllllllll -100 -50 O 50 100 150 Temperature (C) Figure 4.8. Temperature dependence of the linewidth of the resonance peak in solid-state 7Li MNR spectra of LixMoO3 and Lix(H20)y(PEO- 100,000)ZM003. 177 Li+ ions over several sites in the former. The data in the low-temperature region suggest a more well defined coordination environment for the Li+ ion in LixMoO3 as would be expected in this crystalline solid. The linewidth of 2300 Hz in LixMoO3 at 23 0C is much narrower than that of ~12000 Hz observed for Li2M0204 indicating substantial degree of ion motion in the lattice of LixMoOg relative to that of Li2M0204 [19]. This is rationalized by the fact that in the latter the Li+ ions fully occupy well defined crystallographic positions in the lattice [20] while the non-stoichiometric nature of LixMoO3 gives rise to Li mobility via vacant crystallographic sites. The resonance peak in both samples narrows dramatically as the temperature is increased from -80 0C to 100 0C, Figure 4.8. This is attributed to rapid motion of Li+ ions between the M003 layers. At 100 0C the peak linewidth in the spectra of Lix(H20)y(PEO-100,000)zMoO3 is comparable to that of Lio,5(H20)1,3Mon4 [19]. The onset temperature of the transition from a wide to a narrow peak in Lix(HzO)y(PEO-100,000)zMoO3 and LixMoO3 is similar. The spectra of both materials end up with an equally narrow peak at 100 OC indicating comparable rates of hopping of Li ions between different positions in the interior of the two materials. 4. Magnetism Because the materials are formally mixed-valence compounds and exhibit intense intervalence Mos‘i-Mo6+ optical transitions, we expect unpaired electrons to be delocalized over the d-orbitals of the Mo atoms. Magnetic susceptibility measurements were carried out as a function of temperature for LixMoO3 and Lix(H20)y(PEO-100,000)ZMoO3 and the data are displayed in Figure 4.9. Surprisingly, the susceptibility of both 178 compounds is rather small with substantial contributions from temperature independent paramagnetism (XTIP)- Correcting for the latter 1/(Xmolar‘XT1P) vs. T plots for LixMoOg and Lix(H20)y(PEO-100,000)ZM003 are linear in the temperature ranges 5-120 K and 2-300 K, respectively. The Curie constants estimated from the slope of the plots yield [Jeff of 0.15 and 0.09 respectively, however, the weak susceptibilities and the large diamagnetic and XTIP corrections make Ileff values unreliable. It is interesting that the magnetic properties of LixMOO3 are more similar to those of Li0,9Mo6017 [211 than of other molybdenum bronzes such as the blue bronze A03MoO3 (A=K,Tl) and purple bronzes A0,9Mo6017 (A=K, Na, T1) [22]. In the latter bronzes the temperature dependence of the magnetic susceptibility shows transitions at low temperature associated with charge density waves. Such phenomena were not observed in the LixMoO3 samples reported here. Xmolar IITWTI I I I I I I I I I I I I I IW r I I I T I I 100 150 200 250 300 350 Temperature (K) Figure 4.9. Temperature dependence of magnetic susceptibility of LixMoOg (circles) and Lix(H20)y(PEO-100,000)ZM003 (triangles). The measurements were conducted at 1000 G with powder samples. 179 10'4 8 l IIIIIIII l Illlllll I I IIIIIII l I lllllll Conductivity (S/cm) I I IIIIIII I Llllllll 10'7llllllllgllrllliilllilll 50 100 150 200 250 300 Temperature (K) (b) Conductivity (S/cm) S Temperature (K) Figure 4.10. Variable-temperature electrical conductivity measurements for (a) LixMOO3 and (b) Lix(H20)y(PEO-100,000)ZM003. 180 5. Electrical Conductivity Electrical conductivity values for all the nanocomposites are listed in Table 4.2. The room-temperature electrical conductivity of the LixMoO3 was 0.013 S/cm, which is slightly lower than values reported in related materials [23]. As shown in Table 4.2, the conductivity of the nanocomposites is significantly lower than that of LixMoOg and decreases with increasing layer expansion. For materials with extremely large gallery spacing, the electrical conductivity is lower than that of M003 itself. When the measurements are performed under vacuum the conductivity decreases concomitantly with the loss of water from the galleries. This suggests that water is contributing to charge transport in these materials probably via proton mobility. Figure 4.10 shows the temperature dependence of the electrical conductivity of LixMoOg and Lix(HzO)y(PEO-100,000)zMoO3, which are thermally activated. Mixed ionic/electronic conducting nanocomposites are of current interest [43]. Concluding Remarks A new family of polymer-molybdenum bronze nanocomposites have been synthesized. The host material, LixMoO3, was synthesized via a LiBH4 route which is different from the conventional approach. This material exfoliates in water and has affinity for a large variety of polymers. Polymers such as PEG, PEO, PPG, PVP, MCel, PAM and PA6 give well stacked lamellar nanocomposites. Depending on the nature of the polymer and its molecular mass, some of the nanocomposites are soluble and can be processed into films. Electronic transmission spectra show the broad transition associated with the MoS+-Mo6+ couple in these nanocomposites 181 ranging from the IR to the visible region. Solid-state 7Li NMR spectroscopy indicates a more versatile chemical environment in the nanocomposites than in the host which may lead to high ionic conductivities in these lamellar systems. The electrical conductivity of these materials is thermally activated and ranges from 10'2 to 10'7 S/cm. 182 References l (a) Y. Fukushima and S. Inagaki, J. Inclusion Phenom. 1987, 8, 473. (b) E. Ruiz-Hitzky and P. Aranda, Adv. Mater. 1990, _2_, 545. (c) A. Okada and A. Usuki, Mat. Sci. Eng. C-Biomin. 1995, _3_, 109. (d) E. P. Giannelis, Adv. Mater. 1996, 8, 29. (e) Z. Wang, T. Lan and T. J. Pinnavaia, Chem. Mater. 1996, 8, 2200. (f) J. J. Tunney and C. Detellier, Chem. Mater. 1996, 8, 927. (g) J. C. Hutchison, R. Bissessur and D. F. Shriver, Chem. Mater. 1996, 8, 1597. 2 (a) M. G. Kanatzidis, L. M. Tonge, T. J. Marks, H. O. Marcy and C. R. Kannewurf, J. Am. Chem. Soc. 1987, _10_9_, 3797. (b) M. G. Kanatzidis, H. O. Marcy, W. J. McCarthy, C. R. Kannewurf and T. J. Marks, Solid State Ionics 1989, 32-33, 594. (C) M. G. Kanatzidis, C.-G. Wu, H. O. Marcy, D. C. DeGroot, C. R. Kannewurf, A. Kostikas and V. Papaefthymiou, Adv. Mater. 1990, _2_, 364. (d) C.- G. Wu, H. O. Marcy, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf, W.-Y. Leung, M. Benz, E. LeGoff and M. G. Kanatzidis, Synth. Met. 1991, 41-43, 797. 3 (a) Y.-J. Liu, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Chem. Mater. 1991, 8, 992. 0)) Y.—J. Liu, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Adv. Mater. 1993, 8, 369. (c) Y.-J. Liu, J. L. Schindler, D. C. DeGroot, C. R. Kannewurf, W. Hirpo and M. G. Kanatzidis, Chem. Mater. 1996, 8, 525. (d) C.-G. Wu, M. G. Kanatzidis, H.O. Marcy, D. C. DeGroof and C. R. Kannewurf, in R. M. Metzger, et al, ed., Lower—Dimensional Systems and Molecular Electronics, Plenum Press, New York, 1991, pp 427. (e) Y.-J. Liu, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, J. Chem. Soc., Chem. Commun., 1993, 593. (f) C.-G. Wu, D. C. DeGroot, H. O. Marcy, J. L. Schindler, C. R. Kannewurf, Y.-J. Liu, W. Hirpo and M. G. Kanatzidis, Chem. Mater. 1996, 8, 1992. 4 (a) G. M. Kloster, J. A. Thomas, P. W. Brazis, C. R. Kannewurf and D. F. Shriver, Chem. Mater. 1996, 8, 2418. (b) F. Leroux, B. E. Koene and L. F. Nazar, J. Electrochem. Soc. 1996, $9), L181. 5 (a) L. F. Nazar, X. T. Yin, D. Zinkweg, Z. Zhang and S. Liblong, Mat. Res. Soc. Symp. Proc. 1991, 2_10_, 417. (b) L. F. Nazar, Z. Zhang and D. Zinkweg, J. Am. Chem. Soc. 1992, _11_4, 6239. (c) L. F. Nazar, H. Wu and W. P. Power, J. Mater. Chem. 1995, _5_, 1985. 183 10 ll 12 13 14 15 16 (d) T. A. Kerr, H. Wu and L. F. Nazar, Chem. Mater. 1996, 8, 2005. R. Bissessur, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, J. Chem. Soc., Chem. Commun., 1993, 687. (a) C. O. Oriakhi and M. M. Lerner, Chem. Mater. 1996, 8, 2016. (b) P. G. Hill, P. J. S. Foot and R. Davis, Synth. Met. 1996, _7_6, 289. (a) G. Cao and T. E. Mallouk, J. Solid State Chem. 1991, 9_4, 59. (b) J. E. Pillion and M. E. Thompson, Chem. Mater. 1991, 8, 777. (c) A. Clearfield and C. Y. Ortiz-Avila, in T. Bein ed., Supramolecular Architecture ( ACS Symposium £9), American Chemical Society, 1992, pp 178. (d) Y.-J. Liu and M. G. Kanatzidis, Inorg. Chem. 1993, 8;, 2989. (e) Y. Ding, D. J. Jones, P Maireles-Torres and J. Roziere, Chem. Mater. 1995, _7_, 562. W. M. R. Divigalpitiya, R. F. Frindt and S. R. Morrison, J. Mater. Res. 1991, Q, 1103. (a) M. G. Kanatzidis, R. Bissessur, D. C. DeGroot, J. L. Schindler and C. R. Kannewurf, Chem. Mater. 1993, 5_, 595. (b) R. Bissessur, M. G. Kanatzidis, J. L. Schindler and C. R. Kannewurf, J. Chem. Soc., Chem. Commun., 1993, 1582. (c) R. Bissessur, J. L. Schindler," C. R. Kannewurf and MG. Kanatzidis, Mol. Cryst. Liq. Cryst. 1993, 84_5, 249. (d) L. Wang, J. L. Schindler, J. A. Thomas, C. R. Kannewurf and M. G. Kanatzidis, Chem. Mater. 1995, _Z, 1753. (a) E. Ruiz-Hitzky, R. Jimenez, B. Casal, V. Manriquez, A. S. Ana and G. Gonzalez, Adv. Mater. 1993, 8, 738. (b) J. P. Lemon and M. M. Lerner, Chem. Mater. 1994, Q, 207. Hui—Lien Tsai, Jon L. Schindler, Carl R. Kannewurf and Mercouri G. Kanatzidis, Chem. Mater. 1997, 2, 875. I. Lagadic, A. Léaustic and R. Clément, J. Chem. Soc., Chem. Commun., 1992, 1396. P. B. Messersmith and S. I. Stupp, Chem. Mater. 1995, 1, 454. (a) N. Margalit, J. Electrochem. Soc. 1974, 1_28, 1460. (b) J. Desilvestro and O. Haas, J. Electrochem. Soc. 1990,12, 5C. M. G. Kanatzidis and T. J. Marks, Inorg. Chem. 1987, 86, 783. 184 17 18 19 20 21 22 23 D. M. Thomas and E. M. McCarron III, Mat. Res. Bull. 1986, 2_1, 945. Yield obtained by reproducing the experiment in Reference 17. S. Colson, J. M. Tarascon, S. Szu and L. C. Klein, Mat. Res. Soc. Symp. Proc. 1991, _2_18, 405. J. M. Tarascon and S. Colson, Mat. Res. Soc. Symp. Proc. 1989, 188, 421. Y. Matsuda, M. Sato, M. Onoda and K. Nakao, J. Phys. C: Solid State Phys. 1986, _l_9, 6039. (a) M. Greenblatt, in Low-Dimensional Electronic Properties of Molybdenum Bronzes and Oxides, ed. C. Schlenker, Kluwer Academic Publishers, 1989. (b) L. F. Schneemeyer, F. J. DiSalvo, R. M. Fleming and J. V. Waszczak, J. Solid Staste Chem. 1984, g, 358. (c) G. H. Bouchard, Jr., J. Perlstein and M. J. Sienko, Inorg. Chem. 1967, 8, 1682. (d) D. C. Johnston, Phys. Rev. Lett. 1984, 88, 2049. (a) C. Julien and G. A. Nazri, Solid State Ionics 1994, 88, 111. (b) J. O. Besenhard, J. Heydecke, E. Wudy and H. P. Fritz, Solid State Ionics 1983, 8, 61. 185 Chapter 5 OL-RuCl3/POLYMER NANOCOMPOSITES: THE FIRST GROUP OF INTERCALATIVE NANOCOMPOSITES WITH TRANSITION-METAL- HALIDES Introduction Intercalation compounds with polymers as the guest species are a rapid growing class of nanocomposite materials [11. The combination of the two extremely different components, at the molecular level, provides an excellent opportunity to modulate the properties of one or more of the component materials [2]. In some circumstances, it is also a unique way to generate materials which have special properties that are unknown in the individual components [3]. Intercalative polymer hybrids have advantages over their small molecular analogs in compositional stability and mechanical strength, which make them more suitable for applications. As a result of recent efforts, a large variety of intercalative nanocomposites have been prepared using hosts from most major classes of layered inorganic compounds, i.e., clays, layered transition metal chalcogenides, metal oxides, metal phosphates and metal thiophosphates [41. Layered transition metal compounds such as oxides, chalcogenides, halides and thiophosphates have interesting electronic, magnetic, optical and catalytic properties [5]. These properties arise from the close-packed intralayer atomic arrangement and the special atomic interactions which results in partially filled d—electron bands. For example, oc-RuClg shows intralayer ferromagnetism, interlayer antiferromagnetism [6a] and photoconductivity [6b]. On the other hand TaSz and NbSe2 exhibit charge 186 density waves and superconductivity [71. One can expect that the physical properties of these materials can be tuned or changed by means of intercalation. Layered transition metal halides, in contrast to layered transition metal chalcogenides [3], oxides [9] and thiophosphates [10], have not received much attention in the study of intercalation, because most of them are not very stable under even mild conditions for soft chemical reactions. Hydrolysis, dissolution and other decomposition reactions are often common properties of this group of materials. a-RuClg is an exception, and is very stable under these conditions. oc-RuCl3 has a lamellar structure of defect Cdlz structure type [63], see Figure 5.1. The intercalation chemistry of oc-RuC13 has been reported with the insertion of simple cations and neutral polar molecules [11]. Cations can be intercalated through a reduction reaction of oc-RuCl3 or through an ion-exchange reaction, while neutral polar molecules can be incorporated through solvent exchange. These properties are similar to those of layered chalcogenides and oxides such as TiSz, 2H-TaSz and M003, suggesting the possibility of intercalating polymers in this material. In fact, oc-RuCl3 behaves as an excellent polymer-intercalation host. It has affinity for various polymers and is suitable for several of the existing polymer intercalation methods. For example, oc-RuCl3/polyaniline (Ct-RuClg/PANI) nanocomposites were synthesized with in situ redox intercalative polymerization [12], while soluble polymers such as poly(ethylene oxide) (PEO), poly(vinyl pyrrolidone) (PVP) and polyethylenimine (PEI) are intercalated through encapsulative co-precipitation [13] of polymers and on- RuC13 from solutions of exfoliated Ot-RuCl3. A modification of the second method, in situ polymerization coupled with encapsulative precipitation [14] 187 View down the 001 plane 3! ‘3 ‘5 1! ‘3 3' S 3! S :‘$‘$‘$‘¥‘$‘z :‘a.g‘$‘¥;‘3“£‘z . m. 07- .4- .4. View down the 100 plane ‘95;§9‘g§ O 1‘ '5 O 99~~§29~9~ «i ‘1 C Figure 5.1. The structure of oc-RuClg. 188 from suspensions of exfoliated (X-RUC13, can be used to intercalate intractable polymers such as polypyrrole (PPY). This chapter describes the insertion of polymers into (X-RUC13 producing a new group of lamellar nanocomposites with various physicochemical properties. Experimental Section 1. Reagents Ru (99.95%, 325 mesh) was purchased from Cerac Inc. C12 (99.5%) and CO (99.5) gases were from AGA Specialty Gases. LiBH4 (95%), PEO (100,000), PVP (10,000) and PEI (25,000) were purchased from Aldrich Chemical Company, Inc. After the polymers were dissolved, the polymer solutions were filtered to remove insoluble residues. Aniline and pyrrole were purchased from Aldrich and Mallinckrodt Inc. respectively, and were distilled before use. Br2 (reagent grade) and anhydrous FeCl3 (purified) were from Fisher Scientific. Anhydrous ether (99.0%), acetonitrile (99.5%) and isopropanol (99.9%) were from Columbus Chemical Industries Inc., EM Science Inc. and Mallinckrodt Chemical Inc. respectively. 2. Reactions and Sample Preparations a. Preparation of oc-RuCl; oc-RuCl3 was synthesized as described in the literature [15]. The set-up for the reaction, as shown in Figure 5.2, was first purged with nitrogen overnight. An amount of 10 g Ru spread in an alumuna boat was then reacted with mixed C12 and CO gases at 800 0C for 8 h. The ratio of C12 to CO gas was adjusted to about 1:2 or 1:3 and the flow was set at 189 C12 regulator \ Er— Fumace f Clz Alumina boat To hood N aOH Concentrated solution H2304 Figure 5.2 Experimental set-up for the synthesis of (x-RuCl3. 190 approximately 60—120 ml/min. The gas was set to flow before heating and was turned off after the reaction tube cooled to about 50 0C. It is important to keep the gas under constant flow, otherwise a negative presure could build up and the H2SO4 (aq.) could be pushed into the reaction tube. b. Synthesis of (PANI )1R_uC_31_ An amount of 0.1 g oc—RuCl3 in 10 ml of 4% aniline/acetonitrile solution was stirred in open air for 1 week. The product was washed with copious acetonitrile and dried in vacuum. Elemental analysis, done by Quantitative Technologies Inc., gave: 15.33% C, 1.43% H, 3.11% N, which suggests the formula (PANI)0,57(H2O)0,50RuC13 (calculated: 15.3% C, 1.43% H, 2.97% N). The above formula contains 22.7 wt% intercalated species, which agrees with the thermogravimetric analysis (TGA) measurement indicating 22.8 wt% organics and water inside this nanocomposite [16]. c. Synthesis of LilRu_3Cl_ and preparation of RuCl3 monolayer suspensions In a typical reaction, 18 mmol oc-RuCl3 was mixed with 3.6 mmol LiBH4 in 40 ml anhydrous ether under N2 atmosphere for 3 days. The product was collected by filtration, washed with anhydrous ether and pumped to dryness. It was stored in a glove box under N2 atmosphere. The product is designated as LixRuClg, where x z 0.2. An aqueous 0.5 wt% LixRuCl3 monolayer suspension was prepared simply by stirring LixRuCl3 in water for half an hour. 191 d. S nthesis of Li PEO l(RuCli, LiXIPEIIXRuClg and LiXIPVPIXRuCla In the preparation of PEO and PVP nanocomposites, an amount of 40 ml 0.5 wt% LixRuCl3 monolayer suspension (containing 0.96 mmol LixRuCl3) was mixed with 40 ml of polymer solution containing 3 mmol repeat-units of polymer. The mixture was stirred for 2 days in ambient atmosphere. The nanocomposites of PEO and PVP are water-soluble, and have to be collected from very concentrated solutions. Therefore, the reaction mixture was condensed to less than 20 ml under vacuum before it was poured in an organic solvent to precipitate the product. Acetonitrile and isopropanol were used as solvents for PEO and PVP respectively. In the preparation of PEI nanocomposite, the amount of 40 ml 0.5 wt% LixRuCl3 monolayer suspension was mixed with 40 ml of PEI solution containing 7 mmol repeat-units of polymer. The PEI nanocomposite precipitated out from water immediately upon mixing. It was collected with centrifugation after having been stirred as a suspension for 2 days. The nanocomposites were dried under vacuum. The compositions of the three nanocomposites were Lix(PEO)1,5RuCl3 (x~0.2), Lix(PVP)2.2RuCl3 (x~0.2) and Lix(PEI)4,6RuC13 according to T GA measurements under oxygen flow [161. e. Synthesis of (PPYARuClg An amount of 0.45 g pyrrole (6.7 mmol) was dissolved in 40 ml of water and mixed with 40 ml of aqueous LixRuCl3 suspension which contained 0.20 g LixRuClg (0.96 mmol). The mixture was cooled in an ice bath before the dropwise addition of 10 ml aqueous solution of 0.11 g FeC13 (0.67 mmol) which was also cooled in an ice bath. The reaction was carried out by stirring in a stoppered, ice-cooled flask for 24 h. The 192 product was collected by centrifugation, washed with copious water, and dried first in air and then under vacuum. Elemental analysis showed that a (PPY)xRuC13 sample contained 13.88% C, 1.32% H and 3.95% N. This corresponds to a formula (PPY)0,77(H2O)0,60RuC13 (calculated: 13.78% C, 1.31% H and 4.01% N). TGA measurements showed that the material lost 50.2% (theoretical 50.4%) of weight in oxygen flow up to temperatures higher than 450 0C. X-ray powder patterns of the final residue indicated RuO2. f. Oxidation of Li_xRuCl_3 and nanocomposites with Br2 The oxidized RuCl3 was obtained by reacting exfoliated LixRuCl3 (x ~ 0.2) with 10 equiv. of Br2 in water for one day. The RuCl3, which flocculated, was collected by centrifugation, washed with copious water and dried under vacuum. The oxidized “(PEO)yRuCl3” was prepared by reacting directly the nanocomposite solutions, produced by the intercalation reaction, with 10 equiv. of Br2 in water for one day. g. Preparation of oriented Li1(PEO)¥RuCl_3 sample for one-dimensional electron density map calculations One-dimensional electron density (l-D ED) map calculations need highly oriented samples. Lix(PEO)yRuC13 samples for this purpose were cast films. Immediately before the X-ray experiments, the samples were pumped at 75 0C for at least 3 days to remove excess water from the nanocomposite. This procedure ensures that the whole sample has only one phase with a single basal spacing. The X-ray patterns were collected under a nitrogen atmosphere to prevent absorption of moisture from the air, which causes peak broadening and drifting. 193 3. Instrumentation The instrumentation in the measurements such as thermal gravimetric analysis (TGA) and infrared (IR) spectroscopy were the same as described in Chapter 1. X-ray diffraction (XRD) powder patterns were obtained as described in Chapter 1, except the increment for the continuous scanning mode was 005° for general purpose spectra. Room temperature conductivity measurements, variable temperature direct-current electrical conductivity measurements and thermopower measurements, and magnetic susceptibility measurements were conducted as described in Chapter 2. Scanning electron microscopy (SEM) and energy dispersive X-ray microanalysis (EDS) and electron diffraction were done as described in Chapter 3. XRD patterns for one-dimensional electron-density calculations were done as described in Chapter 3. Slits for beam width control and time of data collection were as follows: 0.5° slits and a data collection time of 12 8 per step were chosen for measurements in the range from 2° to about 29°; 10° slits and 60 s per step between 16° and 71°; 20° slits and 60 s per step between 57° and 835°; 4.0° slits and 90 8 per step from 72° to 136°. The step width was kept the same (01°) in the entire 20 range. Results and Discussion 1. Preparation of (PANDXRuClg by in situ Redox Intercalative Polymerization The in situ redox intercalative polymerization reaction is the most direct method to intercalate conductive polymers. Its topotactic character least disturbs the crystalline structure of the host. In the case of 194 FeOCl/polyaniline, even single crystals of the nanocomposite could be obtained [17]. This type of reaction requires a strongly oxidizing host to provide a driving force to pull electrons from the monomers and oxidize them into polymers. In addition, the host should be able to distribute efficiently those electrons through out their structure. Because of the scarcity of such highly oxidizing and conducting hosts, the reaction has been limited to FeOCl [128], V205 [12b] and VOPO4 [120]. We discovered that oc-RuCl3 also happens to be a suitable such host, and can form intercalative nanocomposites with polyaniline. The reaction of an aniline CH3CN solution with (X-RUCI3 in air results in the formation of polyaniline (PANI) within the gallery space of RuCl3. The reflection-mode powder X-ray diffraction (XRD) patterns of the product show a 6.2 A increase in the separation of the RuC13 layers, see Figure 5.3 [18]. This expansion is reasonable for insertion of a monolayer of PANI molecules, and comparable to the 5.94 A observed in (PANI)xFeOCl [171, and 5.2 A in (PANI)XV2OS [12b]. The transmission-mode powder XRD patterns show that the hk0 reflections of (PANI)xRuCl3 remain the same as those of OL-RuClg, indicating that the structure of oz- RuC13 is preserved, see Figure 5.4. This result is supported by electron diffraction experiments which provide the same hk0 diffraction patterns for (PANI)xRuCl3 and oc-RuCl3. The formation of polyaniline between the RuCl3 layers is supported by IR spectroscopy. Almost all peaks in the IR spectra of (PANI)xRuC13 are associated with polyaniline (emeraldine salt) and only a few with anilinium, see Figure 5.5(a). The peaks corresponding to the polymer are much more intense than those of anilinium. Heating the product at 120 0C in air for 5 days removes the peaks corresponding to anilinium at 744 cm-1 195 (X-RUCI3 3L - 1L89A. I (PANI)xRuCl3 E? r422A1 g -—I uyPEokhuog E L JLA 28.78 A UJPVPkRUCk A (PPYhRucg IIIIIIIIIII'IIIIIIIIIIerI 20 (deg) Figure 5.3. X-ray diffraction patterns of oc-RuClg and nanocomposites. 196 O O O 1- O O 0') 0t-FluCI3 - (O O O O O O N 1- N .29 u-4 m 8 E 8 1—4 '- (PANI)XF1uC|3 S O l I I 1 0 10 20 (deg) Figure 5.4. Transmission X-ray diffraction patterns of OL-RuCl3 and (PANI)xRuC13 with indexing. 197 (PAN|)xRuC|3 PAWJI Transmittance ArHHne 1 I I I I I I I I I I I fiI I I r I I I I I r I I I I 1800 1600 1400 1200 1000 800 600 400 Wavenumbers (cm’ 1) Figure 5.5(a). Infrared spectra of (PANI)xRuC13, PANI and aniline. (The peak at 1385 cm-1 is due to an impurity in KBr.) 198 (PPYkRUCQ WW Transmittance PPY j—IIIIIIIITIIIIIIIIIIITIIIIII 1800 1600 1400 1200 1000 800 600 400 Wavenumbers (cm'l) Figure 5.5(b). Infrared spectra of (PPY)xRuCl3 and PPY. (The peak at 1385 cm-1 is due to an impurity in KBr.) 199 Lix(PEO)yRuCI3 Transmittance PEI) III—IjIfiIIIIIIIIIIIIIIIIIIII 1800 1600 1400 1200 1000 800 600 400 Wavenumber (cm' 1) Figure 5.5(c). Infrared spectra of Lix(PEO)yRuCl3 and PEO. (The peak at 1385 cm-1 is due to an impurity in KBr.) 200 Lix(PVP)yRuCI3 Transmittance PVP ITIIIIIIIIIITIIIIIIIIIIIIII 1800 1600 1400 1200 1000 800 600 400 Wavenumber (cm' 1) Figure 5.5(d). Infrared spectra of Lix(PVP)yRuC13 and PVP. (The peak at 1385 cm-1 is due to an impurity in KBr.) 201 and 687 cm'l. The appearance of such monomer peaks indicates that the polyaniline formed in the galleries does not have very high molecular weight. This is reasonable since both the concentration and mobility of the aniline in the gallery are limited and the polymer is expected to be similar to the ~5,000 daltons found in the (PAN I)xFeOCl system. In the process of intercalation, a fraction of Ru3+ is reduced to Ru2+a giving a mixed valence compound. The Ru2+ centers are very stable because they are low spin diamagnetic (16 systems. This probably acts as a powerful driving force for the oxidation of aniline by virtue of the maximized Ligand Field Stabilization Energy (LFSE) in low spin (16 metal centers. Similar to the intercalative polymerization of aniline in FeOCl [123] and V2O5 [12b], the presence of oxygen is key to a successful outcome of the reaction. This was verified by control experiments where in the absence of air or oxygen no intercalation reaction occurred in 23 days. 2. N anocomposites of oc-RuCl3 with Water Soluble Polymers The method of encapsulative precipitation from solutions of exfoliated lamellar solids provides a most convenient way to prepare nanocomposites with water-soluble polymers. With a stable and concentrated aqueous LixRuCl3 monolayer suspension, the intercalation is performed with water-soluble polymers: PEO, PVP and PEI. The reflection XRD patterns of the products show that the interlayer spacings of the Lix(PEO)yRuC13, Lix(PVP)yRuC13 and Lix(PEI)yRuC13 nanocomposites increase by 8.5, 23.0 and 3.6 A respectively, see Figure 5.3. The existence of PEO, PVP and PEI in the nanocomposites is proven by IR spectroscopy. (See Figures 5.5(c) and (d) for IR spectra of Lix(PEO)yRuCl3 and Lix(PVP)yRuC13.) 202 (é) Figure 5.6. Electron diffraction patterns for (a) oc-RuC13 and (b) LixRuClg. 203 Electron diffraction experiments show that the exfoliated RuCl3 has the same hk0 diffraction pattern as OL-RuCl3, see Figure 5.6. The same electron diffraction pattern was seen in (PANI)xRuCl3, suggesting no intralayer structure change during intercalation. 3. oc-RuC13/Polypyrrole Nanocomposites The exfoliation property of oc-RuCl3 offers the possibility to synthesize nanocomposites with insoluble polymers, by in situ polymerization-encapsulative precipitation [14]. In this method, a solution of a monomer (such as pyrrole) is mixed with an exfoliated inorganic host. When an initiator is added, the polymerization causes the co-precipitation of the polymer and the host monolayers, forming a nanocomposite with a certain lamellar thickness. As described in the Experimental section, (PPY)xRuCl3 was produced this way. The production of (PPY)xRuCl3 nanocomposites is indicated by XRD patterns of the layered products with basal spacings around 11.7 A, see Figure 5.3. The existence of the conductive form polypyrrole inside the galleries is confirmed by the characteristic IR spectrum, which exhibits peaks at 1541, 1314, 1150 1043 and 963 cm-1 [191, see Figure 5.5(b). The expansion between the RuC13 layers, 6.0 A, corresponds to a layer of polypyrrole chains arranged almost 40 degrees to the RuCl3 layers. This behavior is similar to that of (PPY)xMoS2 when pyrrole is used in excess in the reaction [14C]. The packing density of polypyrrole in (PPY)xRuCl3, 0.77 pyrrole-unit per RuCl3, is comparable to that in (PPY)xMoS2 prepared under similar conditions, 0.50 pyrrole—unit per MoS2 [201. 204 In the reaction to form (PPY)xRuCl3, only 0.7 equivalent FeC13 was used, which could oxidize at maximum 0.35 equivalent pyrrole to polypyrrole. Therefore, the amount of polypyrrole produced is more than the amount of pyrrole that FeC13 could have oxidized. This is explained by the fact that the ambient oxygen takes part in the reaction as an electron acceptor. It has been proven that ambient oxygen can oxidize and polymerize pyrrole and its oligomers when FeC13 is present as a catalystm]. A control experiment, which was conducted under the same experimental conditions but without the addition of FeC13, produced a lamellar phase with a basal spacing of 10.1 A (expansion of 4.4 A). This phase cannot be de-intercalated by dynamic pumping. It must correspond to a RuC13 intercalation compound with one layer of pyrrole or oligomers lying with the pyrrole rings parallel to the RuC13 layers. This compound lost 48.0 wt% in a TGA experiment in oxygen flow up to 650 0C, which is comparable with that of (PPY)0,77(H2O)0,60RuC13, 50.2 wt%. Another control experiment showed that no polymerization occurred when both FeC13 and RuCl3 were absent. As an alternative method, we also tried the in situ redox intercalative polymerization in order to prepare a RuC13/PPY nanocomposite. When on- RuCl3 was stirred in an aqueous pyrrole solution [22] in open air, intercalation occured in two weeks to form a product with 11.2 A basal spacing. However, IR spectra showed that the polymer formed inside the gallery spacing was not ordinary polypyrrole and further investigation was not performed. 205 4. Charge Transport Properties The intercalation of polymers causes large changes in the properties of oc-RuC13, as expected. Because of the formation of Ru2+ centers which provide free hopping electrons, the electrical conductivity of LixRuCl3 (x ~ 0.2) at room temperature is ~ 0.3 S/cm, about three orders of magnitude higher than that of oc-RuCl3, 5x10"4 S/cm [23]. Intercalation of insulating polymers reduces the conductivity, with those of PEO and PVP nanocomposites being 4.5x10'3 and 1.7x10'3 S/cm respectively. Nanocomposites with conductive polymers have conductivities higher than that of LixRuCl3. (PANI)xRuCl3 has a room temperature conductivity of ~ 1 S/cm, while (PPY)xRuC13 has one of 23 S/cm. It is obvious that the conductivities of these two materials are substantially enhanced by the presence of conductive polymers. Variable temperature measurements on pressed pellets reveal that the electrical conductivities of all oc-RuCl3, LixRuCl3 (x~0.2), Lix(PEO)yRuC13, (PANI)xRuC13 and (PPY)xRuC13 are thermally activated, see Figure 5 .7. In the case of oc-RuCl3, LixRuCl3 and Lix(PEO)yRuC13, the log(o) versus 1/T plots are almost linear. The activation energies, which are calculated according to the formula 0 = (50 e-AE/Zkt, are 0.36, 0.30 and 0.35 eV respectively. Since the pellets used in measurements have inter-particle boundaries, the activation energies do not necessarily correspond to the intrinsic band gaps. In (PANI)xRuC13 and (PPY)xRuCl3, the data do not form straight lines in the log(o) versus l/I‘ plots, which suggests that several kinds of electrical barriers exist in these materials, including electron hopping barriers associated with transport through and across the chains of the conjugated polymer. At temperatures higher than 100 K, the data of (PANI)xRuCl3 and (PPY)xRuC13 almost fall in straight lines. The 206 (PPYkRuCg 00°00: 10'1 V ., (PANl)xRuC|3 10'3 V v v E A V: a E DC] 3 3:“ 10-52— DB > E U 8 : 0. E :E Du LIXRUC|3 o I - U _7- LIX(PEO)yRuCI3 .. k 10 _E 151% @q% :1 10‘9 0” oc-RuCl D 3 O OO 0 10‘1111:11|11|111111I11111111111 2 4 6 8 10 12 14 16 1000frtkrh Figure 5.7. Variable temperature electrical conductivity mesurements for pressed pellets of oc-RuCl3 and nanocomposites. 207 150 I I I I I I I I I I I I I I I T I I l I I I I I 100 _ I. E. : H .- 0.) g - - Q 50 — — é ' : D . - F5 - (PANl)xRuC|3 - 0: " (PPY)xRuCI3 L I. L I I I I I l l l I I L l l I l I 1_L I l l I Id 50 100 150 200 250 300 Temperature (K) Figure 5.8. Thermopower measurements for pressed pellets of LixRuCl3 (x ~ 0.2) and nanocomposites. 208 .. —_—— apparent activation energies are 0.18 ev for (PANI)xRuCl3 (> 125 K) and 0.10 ev for (PPY)xRuC13 (> 160 K). The thermopower, which is less affected by grain boundary effects, was measured for LixRuCl3 (x ~ 0.2), (PANI)xRuCl3 and (PPY)xRuC13, see Figure 5.8. All three materials have positive Seebeck coefficients suggesting that the dominant carriers are holes. The p—type charge- transport behavior is consistent with the fact that the reduced RuC13 layers, polyaniline (emeraldine salt) and polypyrrole are all p-type conductors. Both the high Seebeck coefficient value of ~ 60 ItV/K for LixRuClg and its decreasing trend with rising temperature indicate that LixRuClg is a semiconductor. The hole type transport in the [RuCl3]"' layer arises from the partially empty band composed of t2g type orbitals while electron configuration is somewhat between t2g5 and t2g6. (PANI)xRuCl3 and (PPY)xRuCl3 have lower thermopower values which increase as the temperature increases. This trend is usually seen in metallic conductors. Considering that the reduced RuC13 layers are semiconductors, the bulk metallic-like conductivity as suggested by the small thermopower indicates that charge transport in (PPY)xRuC13 and (PANI)XRuCl3 is controlled by the conductive polymers. 5. Magnetic Susceptibility Studies In a-RuCl3 the intralayer Ru3+ ions are ferromagnetically coupled, however, in adjacent layers the Ru3+ ions are antiferromagnetically (AF) coupled 163’ 24]. The interlayer AF coupling becomes strong at low temperatures (2-20K), and causes a remarkable AF ordering at about 15 .6 K [68’ 241, see Figure 5 .9(a). At high temperatures (50-300K), the interlayer AF coupling is so weak that it has no significant effects on the magnetic 209 susceptibilities. The magnetic susceptibilities of OL-RuCl3 in this range follow Curie-Weiss law with a positive Weiss constant 0, which reflects the weak effect of the intralayer ferromagnetic coupling. In LixRuCl3 (x ~ 0.2), the conspicuous AF ordering disappears and the magnetic susceptibility follows Curie-Weiss law to temperatures as low as 2 K, see Figure 5.9(b). The Weiss constant 0 becomes negative, indicating a change from the weak overall ferromagnetic coupling to a weak overall AF coupling. These changes in magnetism are explained by the introduction of diamagnetic Ru2+ centers in the Ru sublattice, which alters the magnetic couplings both in the layers and between the layers. The magnetic susceptibility data of Lix(PEO)yRuCl3, (PANI)xRuCl3 and (PPY)xRuC13 are similar to those of LixRuCl3, see Figures 5.9(c) and 5.9(d). Since the reduction of the RuCl3 layers has already disrupted the original interlayer AF coupling, the subsequent insertion of the polymers does not have any additional effect on the magnetic susceptibility. The ”eff for (X-RUC13 is 2.32 It}; [251, while those for LixRuCl3 and nanocomposites Lix(PEO)yRuCl3, (PANI)xRuCl3 and (PPY)xRuC13 range from 1.60-1.74 I113. The drop in paramagnetic moment is due to the decrease in the number of unpaired electrons in RuC13 layers because of the presence of diamagnetic low spin Ru2+ centers. The derived paramagnetic moments Ileff and Weiss constants 0 for these compounds are listed in Table 5.1. The xdia and xm, used in the manipulation of the magnetic data are also listed in Table 5.1. As mentioned above, the intercalation of polymers does not bring much change in the magnetic susceptibility of the reduced RuCl3 layers, because the interlayer magnetic coupling has already been broken down by the generation of Ru2+ centers. If the RuCl3 layers were not reduced, i. e., 210 Xmolar 0.020_ _ 500 : ,DE . —400 0.015 : '_' 4 - -300 : ' I x 0.010_— - 8‘ : {200 ‘3" - 1 0.005- - - :100 @0000 " ".,.’:.1...11.11.11.11111..1...11' O'°°°0 50 100 150 200 250 300 ° Temperature (K) Figure 5.9(a). Magnetic susceptibility measurements for a sample of oc- RUCI3. Xmolar 0.020 , ,cJ .0 . 6 ’ I /O’ .— 0015 , 9’ —> - 9 H - 28 ’ L 0.010 — 7 0.005 - _‘ f) l l l I l l I l I I I l I I l l l l l l l l l I l l l T: - 0'0000 50 100 150 200 250 300 Temperature (K) 1 000 Figure 5.9(b). Magnetic susceptibility measurements for LixRuCl3 (x~0.2). 211 Xmolar 0.035 . 1000 0.030 ,j j 800 0.025 a 0.020 1‘ 600 I X I 8 0.015 _ 400 93. 0.010 j ‘ 200 0.005 . 0.000 "I I 1 l I l I l l I l l I l I L L l I I 1 I l I I 1 l l l I - O 0 50 100 150 200 250 300 Temperature (K) Figure 5.9(c). Magnetic susceptibility measurements for Lix(PEO)yRuCl3. xmolar 0.020 0.015 0.010 0.005 0.000 0 IIIIIIIrIIIII I I l l l I l I l l l I l l l I I l l l I I l l l I I I I I I l I I I l l l I l l 1 l I I l l l l 50 100 150 200 250 300 Temperature (K) 1000 0 Figure 5.9(d). Magnetic susceptibility measurements for (PANI)xRuCl3. 212 Table 5.1. Magnetic properties of oc-RuCl3 and nanocomposites “eff 0 Xdia*106 XTIP*106 sample (uB/mol Ru) (K) (l/mol Ru) (1/mol Ru) 0L-RuCl3 2.32 11 -101 0 LixRuC13 (x~0.2) 1.60 -18 -101 176 (PAN I)xRuCl3 1.63 -20 -133, 250 Lix(PEO)yRuC13 1.73 -10 -167 667 (PPY)xRuC13 1 .74 - 15 -142 247 RuC13 (0Xid.) 2.33 6.2 -101 450 (PEO)XRUCI3 1.82 0.2 -143 450 (0Xid.) the magnetic coupling among the Ru3+ ions existed, the weakening or elimination of the interlayer Ru3+ ion magnetic coupling by the polymer intercalation should be observed. Such a phenomenon is expected in the “Br2 oxidized Lix(PEO)yRuC13” (i.e. “(PEO)xRuC13”), in which the Ru2+ centers in the reduced RuC13 layers are oxidized back to Ru3+. In this material the intralayer Ru3+ ferromagnetic coupling should be restored, while the interlayer Ru3+ AF coupling should still be retarded by the increased interlayer separation. An investigation of the magnetic properties of “Br2 oxidized LixRuCl3” (i.e. RuC13) was first carried out, because this material is structurally simpler than the oxidized nanocomposites yet intimately related to them. The oxidized product indeed reestablished the weak overall ferromagnetism at temperatures higher than 50 K. However, the significant AF ordering present in pristine a-RuClg did not recover. Instead, a gradual change from overall ferromagnetism to overall antiferromagnetism at low temperatures was observed, see Figure 5.10(a). It is obvious that 213 0.025 _ 500 0.020:- 0’ ’0 j 400 : I O I .. 0.015:- — 300 ,... é : 3 Ex X 0.010 3- — 200 3 0.0050:— —_ 100 f I I I I I I I I I I I I I I I I I J I I I I I I I I I I I 0’00 50 100 150 200 250 3000 Temperature (K) Figure 5.10(a). Magnetic susceptibility measurements for “Br2 oxidized LixRuCl3” (i. e. RUCI3). 0.25 800 I ’0’ "j 700 D - 0.20 ’0. ’ _: 600 ’0’ : , 0 I -_ 500 h 0.15 [(7 : I—l g D’ 1 400 P? E o ’ . E x 0.10 ,0, ’ _j 300 m o’ - ’0 I 'E 200 0.050 09’ 2 o '1 100 o I 0.0 I I I I I I I I I I I I I I I I I I I I I I; I I I I I I + O 0 50 100 150 200 250 300 Temperature (K) Figure 5.10(b). Magnetic susceptibility measurements for “Br2 oxidized Lix(PEO)yRuC13” (i.e. “(PEO)xRuC13”). 214 ’;,¢; y-ynnzylunu 100 IIIIIIIIITIIIIIIITIIIIIIIIIII 50 ('11 o IIIIIIIIIIIIIIIII IIJIIIIIIIIIIIIIIII Magnetization (Gauss-cm3/mol) _100 7 I I I I I I I I I I I I I L J l 1 1 1 I r I l I I I l I I -3000 -2000 —1000 0 1000 2000 3000 Magnetic Field (Gauss) Figure 5.10(c). Magnetic hysteresis measurements for “Br2 oxidized LIXRUC13” (i.e. R11C13). _ I I I I I I I I I I I I I I I I I I I I I I I I I I I I I :3 400 _r j E _ - 5 200 r -' a) I— _ m - - a - . g 0:- -- s: _ - .9 t . ‘5 -200 - - d.) " - c: , - g0 -400 — — 2 I I I I LI I I I I I I I I I I I I I I I I I I I I I l I - -3000 -2000 -1000 0 1000 2000 3000 Magnetic Field (Gauss) Figure 5.10(d). Magnetic hysteresis measurements for “Br2 oxidized Lix(PEO)yRuC13” (i.e. “(PEO)xRuCl3”). 215 ‘j;g‘;§ run-v exfoliation and restacking causes stacking disorder ( e.g. turbostratic) of the RuC13 layers, so that optimum interlayer AF coupling in oxidized RuCl3 is not possible. The overall ferromagnetic coupling for the oxidation derived “(PEO)xRuCl3” at temperatures higher than 5 K is slightly weaker than that of corresponding RuC13 (> 50 K), as is suggested by the less positive 0 value. The intercalation of polymer causes the AF coupling to almost disappear, as demonstrated by the persistence of the overall ferromagnetism to as low as 5 K, see Figure 5.10(b). The magnetic hysteresis measurement, which is characterized by a hysteresis loop, confirms that the “(PEO)xRuC13” produced by Br2 oxidation of Lix(PEO)yRuCl3 is ferromagnetic at 5 K, see Figure 5.10(d). In contrast, RuCl3 produced by Br2 oxidation of LixRuCl3 does not show a hysteresis loop, see Figure 5.10(c). The latter material has peg and 0 values close to that of oc-RuClg, which means that almost all Ru2+ centers are converted to the Ru3+ state. The [Jeff and 0 of “(PEO)xRuCl3” produced by Br2 oxidation of Lix(PEO)yRuCl3 are between those of un-oxidized nanocomposites and oc- RuC13, which might suggest that some Ru2+ centers remain. The incomplete conversion of Ru2+ to Ru3+ in the nanocomposite could be attributed to the presence of polymers in the interlayer galleries, which stabilize the reduced layers by offering coordination to Li+ ions. Other reasons could include slower oxidation kinetics in the polymer intercalated systems. The paramagnetic moment ueff and Weiss constant 0 of the oxidation derived RuCl3 and “(PEO)xRuC13” are shown in Table 5.1. 216 6. One-Dimensional Electron Density Calculations and Arrangement of Polymer Chains in Lix(PEO)yRuCl3 In inorganic/polymeric nanocomposites, the arrangement of the atoms in the inorganic layers is well defined. However, the arrangement of polymer chains in the interlayer galleries is not so clear, yet the structural issue is critical to a full understanding of the properties. Because the interlayer gallery provides a two-dimensional space for polymer chains, their arrangement in the interlayer gallery also stimulates a lot of interest in fundamental polymer physics. Many researchers have explored this type of situation with different approaches [26]. Because of the simple repeat structure and flexible chain of PEO, its intercalative nanocomposites have been the subject of most investigations [26th 0, d, 6’ H. In addition, many structures for PEO and PEO-complexes are known. The most common conformation of bulk PEO is the helical one [27], which exists in pure PEO spherulites [23]. The planar zigzag conformation has been observed in stretched PEO samples [29]. Two more types of conformations have been found in PEO-HgClz complexes [30’ 31]. There are many PEO—(alkali metal salt) complexes with versatile morphologies and the determination of their structures has been attempted. The conformation of PEO in these alkali metal salt complexes is thought to be either a double helix [32], a waving single helix, which accommodates itself around the alkali ion lattice [33], or one similar to that in the type II PEO-HgClz complex [34]. In addition, many PEO complexes form with organic molecules: PEO-urea complex [35], PEO-thiourea complex [36], PEO-para- dibromobenzene complex [37] and PEO-para-nitrophenol complex [38]. Several common PEO conformations are shown in Figure 5.11. 217 Films of oriented Lix(PEO)yRuCl3 have well defined sharp XRD patterns with up to fourteen 001 reflections corresponding to a resolution of 0.98 A, see Figure 5.12. The l-D ED maps for Lix(PEO)yRuC13, which can provide information about the projected structure on the c axis, were calculated from the X-ray powder pattern of oriented films (see Figure 5.13 [39]), with formulas described in Chapter 3. In the calculation for 1-D ED map, the phases of structure factors are needed. The phase information for the calculation was obtained from the positions of the Ru and C1 atoms. Due to the minor contribution of the diffracted intensity from the polymer, the inorganic layer plays a critical role here in phasing the structure factors and consequently in the computation of the electron density map. The 1-D ED map profile for Lix(PEO)yRuCl3 shows clearly that the electron density due to the intercalates forms two asymmetric peaks placed symmetrically away from the center of the gallery, similar to that of Li0.2(PEO)xTaSQ [1301. As in the case of Lio,2(PEO)xTa82, the 1-D ED map favors best the structural model with two layers of PEO in the gallery in the conformation of type II PEO-HgClz complex [40]. (The model is shown in Figure 5.14.) The wide valley in the l-D ED map in the center of the gallery immediately excludes the possibility of a layer of helically formed PEO, which must have appreciable electron density in the central region of the gallery. Considering that the basal spacing of the Lix(PEO)yRuC13 film used is 13.76 A [411, which gives an interlayer expansion of 8.0 A, the accommodation of two layers of helical PEO in such a gallery is physically impossible, because the Van der Waals diameter of one PEO helix is ~8.0 A [27]. Furthermore, neither the arrangement of PEO in the zigzag conformation (which was proposed in the (PEO)XV205 [26d]) nor that of the 218 (a) % ”Ev/“M" (b) 8 op\’b’\p‘o’b"op‘o"o’. w L?" (d) ’3’ ‘of‘ohbh’oh'nhtf‘rf‘o’ Figure 5.11. Possible PEO conformations: (a) single helix (7/2) [27], (b) zigzag [29], (c) type I PEO-HgC12 [30b], ((1) type II PEO-HgC12 [31], (e) double helix [32]. 219 4‘ “ O; 001 002 Intensity § 0013 005 0011 003 0012 0014 x1000 xlm A [IrrlfiTIIIIP['1-fi11lll[1ll 0 20 40 6O 80 100 120 140 20 (deg) Figure 5 .12. X-ray powder diffraction pattern of oriented Lix(PEO)yRuCl3 film used for the calculation of one-dimensional electron density maps. <——Ru atoms ' <—— Cl atoms , b ‘. .' '2‘ '- PEO I :3 t : I: | I / \ |' l 8 : : .' '. E I' ' I, ‘\ : 'l L“ : ". : '1 I I I I 1 V I’\ , I \__‘l I ‘ ’II' V . . . . . . l . . . I . I . . . . 0 02 0.4 . 06 08 1 c-ax1s Figure 5.13(a). One-dimensional electron density map for Lix(PEO)yRuC13 using the model shown in Figure 5.14. (Solid line, calculated from experimental data; dash line, from model.) 220 <—-———Ru atoms ' <—— Cl atoms . >‘ I .5.) I. PEO I I: | I <0 I I I Q I ‘ / \ II I: s: I | g l' I :\| "‘I II I o I . , I . I I l 2 ' ' |I ' I I. l LL] ‘| " I I I I | "\ ' I I s I v ‘ ' ' ‘ II ‘ I 'I U s \ _ I ‘ l \ I 41 I I I I I I I I I I I I I O 0.2 0.4 0.6 0.8 1 c-axis Figure 5.13(b). One-dimensional electron density map for Lix(PEO)yRuC13 using a model with intragallery double helical PEO that takes an orientation as shown in Figure 5.11(e). (Solid line, calculated from experimental data; dash line, from model.) <—— Ru atoms <——— Cl atoms >5 3:: PEO m G 0) Q / \ C: I O I *5 I I‘ \ ' i3 ' .' . LU 1“ I. ’1' ‘0 L ‘\ : U, I I ‘ "' \I II I I I I I I I I I I I I I I I I 0 0.2 0.4 . 0.6 0.8 C-aXlS 1 Figure 5.13(c). One-dimensional electron density map for Lix(PEO)yRuC13 using an intragallery double helical PEO model that assumes an orientation of 450 rotation relative to the one shown in Figure 5.11(e). (Solid line, calculated from experimental data; dash line, from model.) 221 <——Ru atoms I I I <—— Cl atoms 2." '5 C: a) Q CI 0 b o 2 LL] 0 0.2 0 8 1 ° c-axis Figure 5.13(d). One-dimensional electron density map for Lix(PEO)yRuCl3 using a model with randomly rotated intragallery double helical PEO. (Solid line, calculated from experimental data; dash line, from model.) I <——Ru atoms 3* g I <——— Cl atoms PEO " Q I / I I I c: e ‘. \ .' IE I I I‘ ’ “ \ : I |I It ’ ‘ I, ‘I g "’ I \ ~ I- \ I I I I I I I I P I _;I I I I I I I l 0 0.2 0.4 0.6 0.8 l c-axis Figure 5.13(e). One-dimensional electron density map for Agx(PEO)yRuCl3 using a model similar to that shown in Figure 5.14. (Solid line, calculated from experimental data; dash line, from model.) 222 ‘- 1.43A 3.17A :108A :2. 40 A 13.76 A is: .)® O ; ,2 -. 0 a ‘1 5 4-— Cl atoms a— Ru atoms Figure 5.14. Structural model for Lix(PEO)yRuCl3. For clarity, in the two galleries the directions of the PEO are drawn orthogonal to each other. 223 type I PEO-HgClz complex (which was proposed for the Na0_3(PEO)xMoO3 [260]) is consistent with the observed l-D ED map profile. In the model of Figure 5.14, the PEO chains are placed at a Van der Waals distance from the RuC13 layers, i.e., it is in touch with RUC13. The Van der Waals thickness of a layer of OL-RuCl3 is 5.73(3) A [6a]. According to the crystal structure of zigzag PEO [291, the distance of the outmost C atom in PEO to the Van der Waals boundary of the molecule should be about 1.73 A. Therefore, the distance of the C atoms closer to RuCl3 should be 4.60 A from the Ru layer, and the distance of the O atoms to the middle of the gallery is 1.20 A. As shown in Figure 5.13(a), the 1-D ED map profile of the model matches well the profile calculated from experimental data in all regions except the center of the gallery. In the center of the gallery, the electron density of the model is obviously lower than that from the experiment due to the fact that disordered water molecules could accumulate around the lithium ions in the free space between the PEO layers. The fact that all the O atoms in the PEO face towards the center of the gallery makes this region hydrophilic, suggesting that the Li+ ions may reside there. The width of this hydrophilic region is different in Lix(PEO)yRuCl3 and Lio,2(PEO)xTaSz models, and depends on the amount of water molecules present in it. Electron density due to the Li+ ions themselves is probably not observable in these ED maps due to insufficient quantities of this very light metal in the material. The validity of the argument that the lithium ions could be in the center of the gallery was examined using a control experiment in which the Li+ ions were replaced with Ag+ ions. We reasoned that because both ions are single charged the heavier Ag+ may behave similarly to Li+ in the nanocomposite and at the same time 224 substantially contribute to the diffraction pattern. The 1-D ED map of the Agx(PEO)yRuC13 sample indicated substantially higher electron density in the central region, see Figure 13(e). (In the model for Figure 13(e), 0.2 equivalent of Ag+ ions are put in the center of the gallery. An electron density map calculated from the experimental data using a model without the Ag+ ions also shows the central peak.) This supports the argument that the ions occupy the center of the gallery. There exists some similarity between the present model and the structure of PEO-NaSCN complex, in which PEO also has a conformation similar to that found in type II PEO-HgClz complex [34]. In PEO-NaSCN, the chains are arranged with the CH2CH2 units in proximity with the S atoms of N aSCN, which are less hydrophilic, whereas 0 atoms coordinate with the N a”r ions, which are much more hydrophilic. This environment of PEO is reminiscent of the present Lix(PEO)yRuC13 model. In addition, this particular form of PEO-NaSCN complex exists under high tension and converts to another form with a helical PEO conformation if the tension is released. In the Lix(PEO)yRuC13 case, the polymer chains extend themselves over a 2-D environment, which of course represents a form of tension as well. The proposed Lix(PEO)yRuC13 model matches well the observed l-D ED map and is geometrically and chemically reasonable. This model is different from the one proposed for Li/PEO/montmorillonite nanocomposite in which the Li+ ions are thought to be located near the surface oxides of the silicate layer [421. Presumably the lower affinity of Li+ for C1' causes its migration in the middle of the gallery which presents an oxygen rich coordination environment. Such a PEO arrangement should provide a good diffusion environment for ions in directions parallel to the 225 layers, which could make Lix(PEO)yRuCl3 a good ion conductor, perhaps better than the amorphous PEO-(alkali-metal) complexes that have segments of helical PEO chains in the structure [43]. 7. PEO Conformation from IR Spectra IR spectroscopy may not be used as a decisive technique to probe the PEO conformation in a nanocomposite. The weakness of IR spectroscopy can be seen from the uncertainty of the predictions. The region from 800- 1000 cm-1 of the IR spectra is supposed to be sensitive to the conformation of PEO chains. A trans O-C-C-O would have absorption peaks around 773 and 992 cm-1, while a gauche O-C-C-O would have absorption around 880 and 944 cm-1 [44]. However, all PEO conformations known have only absorptions around 850 and 950 cm-1, even those in type I PEO-HgC12 complex [30b] and PEO-p-nitrophenol complex [33b], in which the trans O-C- C-O exists. Some controversy in the PEO IR peak assignment arises from the disagreement among the type I PEO-HgC12 IR data reported [303, b]. The peaks observed at 1324, 1309, 1014 and 815 cm-1 [303] were assigned to the trans O-C-C-O conformation [44, 260], because there were no corresponding peaks in the IR spectra of bulk PEO. The more recent and dependable IR spectra of type I PEO—HgClz [30b] did not have peaks at 1309, 1014 and 815 cm-1, while the spectra of type II PEO-HgC12 had peaks at 1312 and 1015 cm-1 [30b]. Recalling the fact that the type I PEO-HgC12 converts to type II PEO-HgClz spontaneously [30b], one might suspect that the type I PEO- HgClz sample used by Blumberg and Pollack [3031 contained type II PEO- HgClz. If this is so, 1309 and 1014 cm-1 do not belong to trans O-C-C-O conformation. Therefore, 1324 cm-1 may be the only peak that can be 226 assigned to the trans O-C-C-O conformation, as used in some references [26h 6’ H. The spectrum of Lix(PEO)yRuCl3 does not have a peak near 1324 cm-1, see Figure 5.5(c), which suggests that no trans O-C-C-O conformation exists. This agrees with the choice of the type II PEO-HgClz conformation proposed in the previous section. Absorption peaks of Lix(PEO)yRuC13 are compared in Table 5.2 with those of helical PEO, type I PEO-HgClz and type II PEO-HgClz read from the spectra in Reference 30b [45]. Because of the poor resolution of the IR spectrum of Lix(PEO)yRuCl3, some nearby peaks may merge and some weak peaks may not show. In regions 11 and VII, the peaks of Lix(PEO)yRuCl3 match those of type II PEO-HgClz. In region IV, the broad peak at 1094 cm-1 matches the peak 1100 cm-1 of the type II PEO- HgClz. The peak is assigned to C-0 stretching and its slight shift to the lower frequency is caused by the coordination of O atoms to Li+ ions. The peak at 1154 cm-1 in type II PEO-HgClz is weak, so it does not appear in the Lix(PEO)yRuCl3 spectrum. Type II PEO-HgClz has a pair of peaks in both regions IH and V, while Lix(PEO)yRuCl3 has only one peak in each of them. In type II PEO-HgClz spectra, the two peaks in each pair have a difference of 31 cm'1 or less. Each pair might merge to give a single peak in the nanocomposite spectrum. The absorption of Lix(PEO)yRuCl3 in regions I and V1 is more different from that of type II PEO-HgClz than in other regions. The peaks of Lix(PEO)yRuC13 at 1465, 1455 and 950 cm-1, which are close to those of the helical PEO, may come from PEO chains dangling outside the layers. (Other absorption peaks of free PEO chains may be buried in the absorption of Lix(PEO)yRuC13.) This explanation satisfies region I. In region VI, the absence of the 892 and 944 cm-1 peak in the Lix(PEO)yRuCl3 227 Table 5.2. A comparison of IR absorptions (cm-1) of PEO Lix(PEO)1_5RuC13 pure PEO PEO.HgC12(I) PEO.HgClz(II) region (x~0.5) 1465 (w) 1462 (s) 1468 (s) 1455 (w) 1456 (w) 1451 (s) I 1445 (w) 1445 (s) 1352 (w) 1355 (w) 1350 (s) 1356 (m) 11 1303 (w) 1338 (s) 1328 (m) 1312 (s) 1278 (s) 1278 (s) 1262 (m) 111 1248 (w) 1246 (s) 1237(m) 1239 (s) 1143 (vs) [46] 1154 (m) IV 1094 (vs) 1110 (vs) [46] 1110 (vs) 1100 (vs) 1060 (sh) 1064 (s) 1057 (s) 1044 (s) 1046 (vs) V 1031 (m) 1032 (s) 1015 (vs) 950 (m) 950 (s) 945 (s) 944 (w) 920 (sh) 923 (5) VI 892 (m) 876 (s) 846 (m) 859 (m) VII 837 (s) [46] 831 (In) 835 (s) 228 spectrum might be explained by its low intensity. The absorption in this region is contributed mostly by the 950 cm-1 peak of the free PEO chains and the 923 cm-1 peak of the type H PEO-HgClz conformation. In any case, the IR absorption profile of Lix(PEO)yRuC13 matches the spectrum derived from a superposition of the absorption peaks of type II PEO-HgClz conformation and free PEO chains. Concluding Remarks The stability of oc-RuCl3 in acidic or basic aqueous solutions as well as reducing and oxidizing conditions makes it a suitable compound for intercalation reactions. This is rare among transition-metal halides. The intercalation of oc—RuCl3 with the conducting polymers PANI and PPY and the water soluble polymers PEO, PVP and PEI gives a new class of lamellar (metal halide)/polymer intercalative nanocomposites. The synthesis of these nanocomposites is enabled mainly by the successful exfoliation of oc-RuClg after controlled lithiation, which allows the use of the exfoliation-encapsulative precipitation method. On the other hand, 0L- RuC13 is also one of the few layered hosts that are suitable for in situ redox intercalative polymerization, which provides an additional approach for new nanocomposites. oc-RuClg nanocomposites contain reduced layers in which free electron hopping, associated with Ru2+/Ru3+ couples, raise the electrical conductivity by 2—3 orders of magnitude. The intercalation of conducting polymers can further increase the conductivity. The dominant carriers in the reduced RuCl3x’ layers are holes residing in a narrow tzg type band. The reduction of oc-RuC13 and its polymer intercalation profoundly affect 229 ‘5;{;fil'gf}1'lvvi‘é" !"‘“!'-"¢" the intralayer and interlayer Ru3+ magnetic couplings, so that new magnetic properties appear in the nanocomposites. These magnetic properties can be adjusted by oxidation or other chemical manipulations. Based on X-ray scattering and IR spectroscopy, a structural model is proposed for Lix(PEO)yRuCl3, in which each gallery contains two layers of PEO chains in a conformation found in type II PEO—HgClz. The Li+ ions seem to reside exactly in the middle of the interlayer space sandwiched between two monolayer of PEO. The model suggests that Lix(PEO)yRuC13 could have good two-dimensional ion conductivity. 230 References 1 (a) D. O'Hare, in D. W. Bruce and D. O'Hare ed., Inorganic Materials, John Wiley & Sons Ltd, 1992, pp 164. (b) G. Alberti and T. Bein ed., Comprehensive Supramolecular Chemistry, Vol. _7_, Elsevier Science Ltd, 1996. 2 (a) F. Leroux, B. E. Koene and L. F. Nazar, J. Electrochem. Soc. 1996, 143(9), L181. (b) F. Leroux, G. Goward, W. P. Power and L. F. Nazar, J. Electrochem. Soc. 1997 , 14_4, 3886. 3 Y. Wang and N. Herron, Science 1996, 27_3 (5275), 632. 4 (a) E. P. Giannelis, Adv. Mater. 1996, 8, 29. (b) E. Ruiz-I-Iitzky, P. Aranda, B. Casal and J. C. Galvan, Adv. Mater. 1995, 1, 180. 5 (a) C. Schlenker (Ed.), Low-Dimensional Electronic Properties of Molybdenum Bronzes and Oxides, Kluwer Academic Pub., Netherlands, 1989. (b) A. Manthiram and J. Gopalakrishnan, Rev. Inorg. Chem. 1984, 6, 1. (c) J. A. Wilson and A. D. Yoffe, Adv. Phys. 1969, 18, 193. (d) B. E. Taylor, J. Steger and A. Wold, J. Solid State Chem. 1973, _7_, 461. (e) R. Brec, D. M. Schleich, G. Ouvrard, A. Louisy and J. Rouxel, Inorg. Chem. 1979, 1_8, 1814. (f) S. Harris, R. R. Chianelli, J. Catal. 1984, 8_6_, 400. (g) J. Brenner, C. L. Marshall, L. Ellis, N. Tomczyk, J. Heising, M. G. Kanatzidis, Chem. Mater. 1998, _5_, 1244. 6 (a) J. M. Fletcher, W. E. Gardner, A. C. Fox and G. Topping, J. Chem.Soc. (A), 1967, 1038. (b) I. Pollini, Phys. Rev. B 1994, 50, 2095. 7 R. H. Friend and A. D. Yoffe, Adv. Phys. 1987, i6, 1. 8 (a) C.-H. Hsu, M. M. Labes, J. T. Breslin, D. J. Edmiston, J. J. Winter, H. A. Leupold and F. Rothwarf, Nature, Phys. Sci., 197 3, A6 (155), 122. (b) M. G. Kanatzidis, R. Bissessur, D. C. DeGroot, J. L. Schindler and C. R. Kannewurf, Chem. Mater. 1993, 5, 595. 9 (a) Y.-J. Liu, D. C. DeGroot, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Adv. Mater. 1993, 5, 369. (b) L. F. Nazar, Z. Zhang and D. Zinkweg, J. Am. Chem. Soc. 1992,1_14, 6239. 231 10 ll l2 l3 14 15 16 I. Lagadic, A. Léaustic and R. Clement, J. Chem. Soc., Chem. Commun., 1992, 1396. (a) R. Schollhom, R. Steffen and K. Wagner, Angew. Chem. Int. Ed. Engl. 1983, 22, 555. (b) R. Steffen and R. Schollhorn, Solid State Ionics 1986, 2_2_, 31. (e) W. Nonte, M. Lobert, W. Muller-Warmuth and R. Schollhorn, Synthetic Metals 1989, 3A, 665. In situ redox intercalative polymerization has been successfully applied to FeOCl, V205 and VOPO4: (a) M. G. Kanatzidis, L. M. Tonge, T. J. Marks, H. O. Marcy and C. R. Kannewurf, J. Am. Chem. Soc. 1987, 102, 3797. (b) C.-G. Wu, M. G. Kanatzidis, H. O. Marcy and C. R. Kannewurf, J. Am. Chem. Soc. 1989, L1, 4139. (c) G. Matsubayashi and H. Nakajima, Chem. Lett. 1993, 31. The method of encapsulative precipitation from solutions of exfoliated lamellar solid has be applied to M082, M003, TaSZ and NbSe2: (a) R. Bissessur, M. G. Kanatzidis, J. L. Schindler and C. R. Kannewurf, J. Chem. Soc., Chem. Commun., 1993, 1582. (b) L. Wang, J. Schindler, C. R. Kannewurf and M. G. Kanatzidis, J. Mater. Chem. 1997, 1, 1277. (c) Chapter 3. (d) H.-L. Tsai, J. L. Schindler, C. R. Kannewurf and M. G. Kanatzidis, Chem. Mater. 1997 , 9, 875. The method of in situ polymerization coupled with encapsulative precipitation has been applied to MoS2, M003 and W82: (a) L. Wang, J. L. Schindler, J. A. Thomas, C. R. Kannewurf, and M. G. Kanatzidis, Chem. Mater. 1995, 1, 1753. (b) T. A. Kerr, H. Wu and L. F. Nazar Chem. Mater. 1996, 8, 2005. (c) Chapter 2. G. Grauer, Handbook of Preparative Inorganic Chemistry, Vol. 2, Academic Press Inc., New York, 1965, pp 1597. TGA indicated a total loss of 48.2% in oxygen flow at temperatures up to 650 0C. By comparing with the TGA results of oc-RuCl3 under the same condition which had losses of 32.85 and 33.07% in two trials, the amount of organics and water inside the nanocomposite was determined. The amount of intercalates inside LiX(PEO)yRuCl3, Lix(PVP)yRuCl3 and Lix(PEI)yRuC13 was also determined this way. 232 17 18 19 20 21 22 23 24 25 26 C.-G. Wu, D. C. DeGroot, H. O. Marcy, J. L. Schindler, C. R. Kannewurf, T. Bakas, V. Papaefthymiou, W. Hirpo, J. P. Yesinowski, Y.-J. Liu, and M. G. Kanatzidis, J. Am. Chem. Soc. 1995, _1_1_7, 9229. Due to the preferential orientation of the layers on the sample holder, the reflection-mode powder XRD patterns show predominantly the 001 reflections. The transmission-mode powder XRD patterns show mainly hkO reflections. Polypyrrole exhibits peaks at 1540, 1300, 1150, 1040 and 900 cm-1: E. T. Kang, K. G. Neoh, T. C. Tan and Y. K. Ong, J. Macromol. Sci. -Chem. 1987, A2_4(6), 631. The (PPY)xMoS2, prepared under similar conditions, has an x about 0.5. Considering that the unit area of RuCl3 is 1.75 times that of MoS2, the amounts of polypyrrole in the galleries are quite comparable. (A M082 unit cell has one MoS2; its a dimension is 3.159 A. A RuCl3 unit cell has 2 RuC13; its a dimension is 5.87 A.) N. Toshima and O. Ihata, Synth. Met. 1996, B, 165. The choice of solvent is important, because a pyrrole/CH3CN solution significantly reduces the rate of the reaction and does not produce single-phase (PPY)xRuCl3 even after one month. L. Binotto, I. Pollini and G. Spinolo, Phys. Stat. Sol. (B), 1971, g, 245. Y. Kobayashi, T. Okada, K. Asai, M. Katada, H. Sano and F. Ambe, Inorg. Chem. 1992, Q, 4570. The literature value was 2.25 LLB, see Reference 24. (a) L. F. Nazar, Z. Zhang and D. Zinkweg, J. Am. Chem. Soc. 1992, M, 6239. (b) P. Aranda and E. Ruiz-Hitzky, Chem. Mater. 1992, 4, 1395. (c) L. F. Zazar, H. Wu and W. P. Power, J. Mater. Chem. 1995, 5, 1985. (d) Y.-J. Liu, J. L. Schindler, D. C. DeGroot, C. R. Kannewurf, W. Hirpo and M. G. Kanatzidis, Chem. Mater. 1996, 8, 525. (e) J. J. Tunney and C. Detellier, Chem. Mater. 1996, 8, 927. (f) Y. Matsuo, K. Tahara and Y. Sugie, Carbon 1997 , 85, 233 27 28 29 30 31 32 33 34 35 36 37 38 113. (g) T. A. Kerr, H. Wu and L. F. Nazar, Chem. Mater. 1996, 8, 2005. (h) G. R. Goward, T. A. Kerr, W. P. Power and L. F. Nazar, Adv. Mater. 1998,18, 449. (a) Y. Takahashi and H. Tadokoro, Macromolecules 197 3, _6, 672. (b) J. Brandrup and E. H. Immergut, Polymer Handbook, 3rd Ed., John Wiley & Sons, 1989, New York, pp VI/72. F. P. Price and R. W. Kilb, J. Polym. Sci. 1962, 5_7_, 395. Y. Takahashi, I. Sumita and H. Tadokoro, J. Polym. Sci., Polym. Phys. Ed., 1973, 1_1,, 2113. (a) A. A. Blumberg and S. S. Pollack, J. Polym. Sci., Part A, 1964, 2, 2499. (b) R. Iwamoto, Y. Saito, H. Ishihara and H. Tadokoro, J. Polym. Sci., A-Z, 1968, _6_, 1509. M. Yokoyama, H. Ishihara, R. Iwamoto and H. Tadokoro, Macromolecules 1969, 2, 184. J. M. Parker, P. V. Wright and C. C. Lee, Polymer 1981, 22, 1305. (a) Y. Chatani and S. Okamura, Polymer 1987, E, 1815. (b) P. Lightfoot, M. A. Mehta and P. G. Bruce, Science 1993, 2Q, 883. (c) P. Lightfoot, J. L. Nowinski and P. G. Bruce, J. Am. Chem. Soc. 1994, _1_1__6_, 7469. (d) P. Lightfoot, M. A. Mehta and P. G. Bruce, J. Mater. Chem. 1992, 2, 379. Y. Chatani, Y. Fujii, T. Takayanagi and A. Honma, Polymer 1990, 3_1_, 2238. H. Tadokoro, T. Yoshihara, Y. Chatani and S. Murahashi, J. Polym. Sci. B, 1964, 2, 363. H. Tadokoro, Macromol. Revs. 1966, _1_, 119. J. J. Point and C. Coutelier, J. Polym. Sci., Polym. Phys. Ed., 1985, Q, 231. (a) J. J. Point and P. Damman, Macromolecules 1992, 25, 1184. (b) P. Daman and J. J. Point, Macromolecules 1994, 2_7, 3919. 234 39 40 41 42 43 44 45 46 Since the calculated l-D ED maps for the models with zigzag PEG and type I PEO-HgC12 conformations are similar to those for their Li0.2(PEO)xTaS2 anologs, Figure 5.13 shows only those for the models with type II PEO-HgC12 and double helical PEO conformations. This conformation was also suggested for poly(ethylene glycol) (PEG) in kaolinite/PEG nanocomposites [263], because its hydrophilic side has higher affinity for the highly polar gibbsite Al(OH)3 side and its hydrophobic side suits the not so polar tetrahedral Si02 side. This value is estimated from a (cos20/sin0 + cos20/0) correction: H. P. Klug and L. E. Alexander, X-ray Diflraction Procedures (for polycrystalline and amorphous materials), 2nd ed., John Wiley & Sons, New York, 1974, pp 594. D.-K. Yang and D. B. Zax, J. Chem. Phys. 1999, 110, 5325. Preliminary measurements done by Dr. Jin-Ho Choi’s group at Seoul National University, Korea, demonstrate that the ion conductivity of Lix(PEO)yRuCl3 at 20 0C is 6.3x10'5 S/cm, equal to or better than the best (lithium salt)/polymer electrolytes: F. M. Gray, Solid Polymer Electrolytes, VCH Publisher, New York, 1991, Chapter 5, pp 83. B. L. Papke, M. A. Ratner and D. F. Shriver, J. Phys Chem. Solids 1981, Q, 493. Reference 30b provides the spectra of two orientations with the electric vector perpendicular and parallel to the fiber axes. An un- oriented spectrum should be close to the superposition of the two spectra. The peaks around 1148, 1112 and 851 cm“1 are characteristic of the helical PEO conformation: H. Matsuura and T. Miyazawa, Spectrochim. Acta 1967, 23_A, 2433. 235 ' .1: .f....f